Reactivity of Triruthenium Furyne and Thiophyne Clusters: Multiple

Feb 22, 2012 - Department of Chemistry, University of Warwick, Gibbet Hill Road, Coventry CV4 7AL, U.K.. Organometallics , 2012, 31 (7), pp 2546–255...
0 downloads 0 Views 742KB Size
Article pubs.acs.org/Organometallics

Reactivity of Triruthenium Furyne and Thiophyne Clusters: Multiple Additions of Thiolato and Selenolato Ligands through Oxidative Addition of S−H and Se−H Bonds Md. Iqbal Hossain,† Md. Delwar H. Sikder,† Shishir Ghosh,† Shariff E. Kabir,*,† Graeme Hogarth,*,‡ and Luca Salassa§ †

Department of Chemistry, Jahangirnagar University, Savar, Dhaka 1342, Bangladesh Department of Chemistry, University College London, 20 Gordon Street, London WC1H OAJ, U.K. § Department of Chemistry, University of Warwick, Gibbet Hill Road, Coventry CV4 7AL, U.K. ‡

S Supporting Information *

ABSTRACT: Reactions of 50-electron furyne and thiophyne clusters Ru 3(CO)7(μ-H)(μ3 -η2-C4H2E){μ-P(C4H3E) 2}(μdppm) (1, 2; E = O, S) with thiols, dithiols, and benzeneselenol leads to the oxidative addition of the E−H bonds followed by concomitant elimination of the alkyne (probably as the alkene) to afford a range of new thiolato and selenolato triruthenium complexes. Addition of PhSH or i PrSH in boiling benzene affords the 48-electron clusters Ru3(CO)5(μ-SR)2{μ-P(C4H3E)2}(μ-dppm)(μ-H) (3−6; E = O, S; R = Ph, iPr) resulting from the addition of 2 equiv of thiol. In contrast, analogous reactions with 1,2-ethanedithiol and 1,3-propanedithiol yield the 50-electron clusters Ru3(CO)3{μ-S(CH2)nS)2{μ-P(C4H3E)2}(μ-dppm)(μ-H) (7−10; E = O, S; n = 2, 3), in which four S−H bonds have been activated. A similar multiple addition reaction is seen upon addition of PhSeH to 1, affording the tetraselenolato complexes Ru3(CO)4(κ1-SePh)(μ-SePh)3{μ-P(C4H3O)2}(μ-dppm)(μ-H) (11) and Ru3(CO)3(μ-SePh)4{μ-P(C4H3O)2}(μ-dppm)(μ-H) (12). Reaction of 2 with PhSeH gave the tetraselenolato complex Ru3(CO)4(κ1-SePh)(μ-SePh)3{μ-P(C4H3S)2}(μ-dppm)(μ-H) (13) together with bis(seleno)-capped Ru3(CO)5{PPh(C4H3S)2}(μ3-Se)2(μ-SePh)2(μ-dppm) (14) resulting from further cleavage of two selenium−carbon bonds and formation of a new carbon−phosphorus bond. The new clusters have been characterized by a combination of analytical and spectroscopic methods, and the molecular structures of 3, 4, 7, 8, and 11 have been determined by single-crystal X-ray diffraction studies. Complexes 7−10 are examples of 50-electron clusters containing three apparent metal−metal bonds; however, DFT calculations carried out for 7 show that the longest metal−metal interaction of 3.119 Å is actually held in place by the bridging thiolato and diphosphine ligands and does not represent a direct metal−metal bonding interaction.



INTRODUCTION The chemistry of chalcogen-containing transition-metal clusters remains an area of intense interest, due to the biological relevance of thiolato and sulfido ligands and also the applications of metal chalcogenides in microelectronics.1 Incorporation of chalcogenide moieties into low-valent clusters can be achieved via several synthetic routes, including the reaction of the elements themselves2 and the oxidative addition of R3PE3 or RE−ER4 reagents to cluster centers. A further route widely utilized toward the incorporation of thiolato ligands is the oxidative addition of thiols.5 As early as 1937 Hieber and Spacu reported6 that oxidative addition of thiols to Fe 3 (CO)12 afforded what are now known to be the bis(thiolato) complexes Fe2(CO)6(μ-SR)2.7 In 1965 Osborne and Stone developed this approach, isolating Re2(CO)8(μSPh)2 upon room-temperature addition of thiophenol to HRe(CO)5 and proposing the formation of the 48-electron © 2012 American Chemical Society

trinuclear cluster Re3(CO)9(μ3-SPh)3 when using more vigorous conditions.8 Reaction of thiols with the heavier group 8 trimetallic clusters M3(CO)12 (M = Os, Ru) and their diphosphine-stabilized derivatives M3(CO)10(μ-dppm) (dppm = Ph2PCH2PPh2) under mild conditions generally result in the oxidative addition of a single thiol to produce M3(CO)10(μSR)(μ-H)9 and M3(CO)8(μ-SR)(μ-H)(μ-dppm),10 respectively. Under more forcing conditions and with excess thiol, poorly characterized polymeric materials of the form [Ru(SR)2(CO)2]n result.11 We have been interested in the reactivity of low-valent triosmium and triruthenium clusters toward various chalcogen Special Issue: F. Gordon A. Stone Commemorative Issue Received: October 31, 2011 Published: February 22, 2012 2546

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

Scheme 1

isostructural. Each 31P{1H} NMR spectrum displays two doublets and a doublet of doublets, which confirms that all three phosphorus atoms remain in the products. In order to unambiguously determine their molecular structures, singlecrystal X-ray diffraction studies were carried out on 3 and 4, the results of which are summarized in Figures 1 and 2, respectively. Both consist of a triangle of ruthenium atoms with two long (Ru−Ru (range) = 2.958(1)−3.018(1) Å) and one comparatively short (Ru−Ru = 2.7504(9) and 2.787(9) Å) ruthenium−ruthenium interactions, the latter in each case being the metal−metal vector spanned by phosphido and hydride ligands. The disposition of the latter along this edge is confirmed by their 1H NMR spectra, which show a doublet of doublets in the hydride region (δ −16.39 for 3 and δ −16.38 for 4). One thiolato ligand asymmetrically bridges the metal− metal edge, which is also spanned by the diphosphine, while the second spans the third ruthenium−ruthenium edge. A feature of both structures is the tilting of the diphosphine out of the trimetallic plane, presumably so as to minimize the steric crowding created by phenyl groups with those of the thiolato ligands. A second noteworthy feature is the relative anti arrangement of the two thiolato substituents. Thus, the phenyl group on the thiolato ligand which spans the edge also occupied by the diphosphine is constrained to orient away from the latter in order to minimize adverse steric interactions, and this in turn results in the phenyl group of the second thiolato ligand to do likewise. The relative anti orientation of the two thiolato substituents is similar to the situation found in the diiron complexes Fe2(CO)4(μ-SAr)2(μ-dppm).15 The relatively facile addition of 2 equiv of thiol is quite distinct from the reactivity observed for M3(CO)12 or M3(CO)10(μ-dppm).9,10 It has been shown that the first stage in activation of the relatively strong hydrogen−sulfur bond is precoordination of the sulfur to the metal center, which then results in a large reduction in the hydrogen−sulfur bond strength.16 We presume that this initially occurs for 1 and 2 upon loss of a carbonyl. Oxidative addition can then proceed, but as seen in the related addition of Ph2Te2 to 1 (Scheme 1),14 this can be associated with a concomitant transfer of a hydride to the coordinated alkyne, generating an alkenyl ligand. In the absence of labeling studies we cannot ascertain whether this hydrogen atom comes from the thiol or the cluster hydride, but some type of alkenyl-thiolato intermediate seems likely. We have not been able to detect this in the reaction mixture and thus assume that a second carbonyl loss and thiol binding are facile. A second oxidative addition and hydride transfer is then proposed, the latter leading to formation of an alkene which is not expected to bind strongly to the cluster, the loss of this twoelectron-donor ligand being compensated for by the formation of a new metal−metal interaction. Reactions of 1 and 2 with Dithiols. Diiron complexes containing bridging dithiolato ligands have attracted enormous attention over the past decade, as they are biomimetics of the

donor ligands such as thiols, dithiols, and RE−ER (R = alkyl, aryl; E = S, Se, Te), primarily with the aim of preparing novel chalcogen-containing clusters but also of understanding reactivity patterns.12 We recently reported the synthesis of triruthenium furyne and thiophyne clusters Ru3(CO)7(μH)(μ3-η2-C4H2E){μ-P(C4H3E)2}(μ-dppm) (1, 2; E = O, S), formed in high yields upon gentle heating of Ru3(CO)9{P(C4H3E)3}(μ-dppm) in the presence of Me3NO and have found that they show intriguing reactivity patterns toward various substrates.13,14 For instance, diphenyl ditelluride undergoes ready oxidative addition to the furyne cluster 1 to form an alkenyl complex, which in turn rearranges to an unusual carbene cluster upon heating in toluene (Scheme 1).14 Inspired by the facile nature of the addition of the tellurium− tellurium bond, we sought to investigate the incorporation of related thiolato and selenolato ligands into the triruthenium framework via reactions with thiols, dithiols, and benzeneselenol. Herein we report that this strategy is indeed successful and leads to the rapid oxidative addition of up to four E−H bonds to produce a range of otherwise inaccessible low-valent clusters.



RESULTS AND DISCUSSION Reactions of Ru3(CO)7(μ-H)(μ3-η2-C4H2E){μ-P(C4H3E)2}(μ-dppm) (1, 2; E = O, S) with Thiols. The furyne and thiophyne clusters Ru 3 (CO) 7 (μ-H)(μ 3 -η 2 -C 4 H 2 E){μ-P(C4H3E)2}(μ-dppm) (1, 2; E = O, S) react with an excess of thiophenol and 2-methyl-2-ethanethiol to afford Ru3(CO)5(μSR)2{μ-P(C4H3E)2}(μ-dppm)(μ-H) (3−6; R = Ph, iPr) in yields of between 33 and 66% (Scheme 2). The transformation Scheme 2

involves the net addition of 2 equiv of thiol and concomitant loss of two carbonyls accompanied by the elimination of the cluster-bound alkyne, most probably as furan or thiophene after addition of hydrogen, although our experimental conditions did not allow us to confirm this. Considering that the bridging thiolato ligands act as three-electron-donor ligands, then formation of a new metal−metal bond balances the total loss of eight electrons, converting the 50-electron starting clusters into electron-precise 48-electron clusters. All exhibit five bands in the carbonyl region of their IR spectrum, the similar patterns suggesting that they are 2547

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

Figure 1. ORTEP diagrams of the two independent molecules of Ru3(CO)5{μ-P(C4H3O)2}(μ-SPh)2(μ-dppm)(μ-H) (3), showing 50% probability thermal ellipsoids. Selected interatomic distances (Å) and angles (deg) for one molecule: Ru(1)−Ru(2) = 2.9580(10), Ru(1)−Ru(3) = 2.7587(9), Ru(2)−Ru(3) = 3.0011(9), Ru(1)−P(3) = 2.279(2), Ru(3)−P(3) = 2.278(2), Ru(2)−P(2) = 2.335(2), Ru(1)−P(1) = 2.277(2), Ru(2)−S(1) = 2.419(2), Ru(1)−S(2) = 2.357(2), Ru(2)−S(2) = 2.417(2), Ru(3)−S(1) = 2.392(2); Ru(3)−Ru(1)−Ru(2) = 63.22(2), Ru(1)−Ru(2)−Ru(3) = 55.15(2), Ru(1)−Ru(3)−Ru(2) = 61.63(2), P(1)−Ru(1)−P(3) = 99.32(8), P(3)−Ru(1)−Ru(3) = 52.72(6), P(3)−Ru(1)−Ru(2) = 99.96(6), Ru(3)−P(3)−Ru(1) = 74.52(7), S(2)−Ru(2)−S(1) = 90.99(7), S(1)−Ru(2)−Ru(3) = 51.01(5), P(3)−Ru(1)−S(2) = 150.68(8), S(1)−Ru(2)− Ru(1) = 90.90(5), Ru(3)−S(1)−Ru(2) = 77.17(6), P(3)−Ru(3)−S(1) = 147.57(8), Ru(1)−S(2)−Ru(2) = 76.56(6).

Figure 2. Solid-state molecular structure of Ru3(CO)5{μ-P(C4H3S)2}(μ-SPh)2(μ-dppm)(μ-H) (4), showing (a) 50% probability thermal ellipsoids and (b) the cluster core with phenyl groups on the diphosphine removed. Selected interatomic distances (Å) and angles (deg): Ru(1)−Ru(2) = 2.7504(9), Ru(1)−Ru(3) = 3.0175(10), Ru(2)−Ru(3) = 2.9875(9), Ru(1)−P(1) = 2.292(2), Ru(2)−P(1) = 2.301(2), Ru(2)−P(2) = 2.288(2), Ru(3)−P(3) = 2.368(2), Ru(3)−S(1) = 2.420(2), Ru(1)−S(1) = 2.387(2), Ru(2)−S(2) = 2.364(2), Ru(3)−S(2) = 2.439(2); Ru(3)−Ru(1)− Ru(2) = 62.19(2), Ru(1)−Ru(2)−Ru(3) = 63.30(2), Ru(1)−Ru(3)−Ru(2) = 54.52(2), P(2)−Ru(2)−P(1) = 99.90(7), P(3)−Ru(3)−Ru(1) = 122.04(6), P(1)−Ru(1)−Ru(2) = 53.35(5), Ru(1)−P(1)−Ru(2) = 73.57(6), S(1)−Ru(3)−S(2) = 87.26(7), S(2)−Ru(3)−Ru(1) = 90.06(5), P(1)−Ru(1)−S(1) = 147.63(8), S(1)−Ru(3)−Ru(2) = 89.52(5), Ru(2)−S(2)−Ru(3) = 76.91(6), P(1)−Ru(2)−S(2) = 149.65(8), Ru(1)−S(1)− Ru(3) = 77.77(6).

active site of the iron-only hydrogenase enzyme,17 and recently developing this theme, we have shown that related triiron complexes can also act as proton reduction catalysts.18 Given the relatively facile addition of 2 equiv of thiol to 1 and 2, we

aimed to synthesize similar products, viz. Ru3(CO)5{μS(CH2)nS}{μ-P(C4H3E)2}(μ-dppm)(μ-H) (n = 2, 3), from the respective reactions with 1,2-ethanedithiol (n = 2, edt) and 1,3-propane dithiol (n = 3, pdt). Indeed, both 1 and 2 react 2548

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

Scheme 3

Figure 3. Solid-state molecular structure of Ru3(CO)3{μ-pdt}2{μ-P(C4H3O)2}(μ-dppm)(μ-H) (7), showing (a) 50% probability thermal ellipsoids and (b) the cluster core with phenyl and furyl groups removed for clarity. Selected interatomic distances (Å) and angles (deg): Ru(1)−Ru(2) = 3.1193(8), Ru(1)−Ru(3) = 2.9056(8), Ru(2)−Ru(3) = 2.8537(8), Ru(1)−P(1) = 2.3143(18), Ru(1)−P(3) = 2.3560(19), Ru(2)−P(2) = 2.3217(19), Ru(3)−P(3) = 2.3257(19), Ru(1)−S(1) = 2.4340(18), Ru(2)−S(1) = 2.4183(18), Ru(1)−S(2) = 2.4128(18), Ru(3)−S(2) = 2.3912(18), Ru(2)−S(3) = 2.4225(18), Ru(3)−S(3) = 2.4119(18), Ru(3)−S(4) = 2.4061(19), Ru(2)−S(4) = 2.4122(19); Ru(3)−Ru(1)−Ru(2) = 56.409(19), Ru(1)−Ru(2)−Ru(3) = 58.011(18), Ru(1)−Ru(3)−Ru(2) = 65.58(2), P(1)−Ru(1)−P(3) = 100.18(7), P(3)−Ru(1)−Ru(3) = 51.17(5), P(1)−Ru(1)−Ru(2) = 89.13(5), Ru(1)−P(3)−Ru(3) = 76.72(6), S(1)−Ru(2)−S(3) = 92.55(6), S(4)−Ru(2)−S(3) = 80.99(6), S(4)− Ru(2)−S(1) = 147.34(6), Ru(3)−Ru(1)−S(1) = 95.30(5), S(2)−Ru(1)−Ru(2) = 93.01(5), S(1)−Ru(1)−Ru(2) = 49.77(4), Ru(3)−Ru(2)−S(1) = 97.00(5), S(4)−Ru(2)−Ru(1) = 97.80(5), S(3)−Ru(2)−Ru(1) = 92.29(5), S(4)−Ru(3)−Ru(1) = 103.90(5), S(3)−Ru(3)−Ru(1) = 97.97(5), S(1)−Ru(2)−S(3) = 92.55(6), Ru(3)−S(2)−Ru(1) = 74.43(5), Ru(3)−S(4)−Ru(2) = 72.64(5), Ru(2)−S(1)−Ru(1) = 80.01(5), Ru(3)−S(3)− Ru(2) = 72.36(5), P(3)−Ru(3)−S(3) = 150.05(7), P(2)−Ru(2)−S(3) = 171.34(7).

Figures 3 and 4, respectively. The main feature of the molecules is the presence of two propanedithiolato ligands which take up different coordination modes: one capping a face of the ruthenium triangle, the sulfur atoms (S(1) and S(2)) bridging two different ruthenium−ruthenium vectors, while the second (S(3) and S(4)) bridges the third ruthenium−ruthenium bond only. The two phosphorus-containing ligands and a hydride also span metal−metal edges, and each ruthenium atom also supports a terminal carbonyl. Thus, the loss of four carbonyls and addition of the same number of sulfur−hydrogen bonds is in keeping with the supposition that carbonyl loss and sulfur coordination are prerequisites for the observed oxidative addition reactions. Considering each dithiolato ligand to be a six-electron donor, then the cluster has a TEC of 50 and as such would be expected to have only two ruthenium−ruthenium bonds. The ruthenium−ruthenium bond lengths in 7 and 8 fall

smoothly with these dithiols, but the isolated products contain two dithiolato ligands coordinated to the metal core using all four sulfur atoms. Thus, heating 1 and 2 with an excess of dithiol gave Ru3(CO)3(μ-pdt)2{μ-P(C4H3E)2}(μ-dppm)(μ-H) (7, 8; E = O, S; yields 39 and 20%) and Ru3(CO)3(μ-edt)2{μP(C4H3E)2}(μ-dppm)(μ-H) (9, 10; E = O, S; yields 50 and 30%), respectively (Scheme 3). The IR spectra of clusters 7−10 are again very alike, suggesting they have similar distributions of carbonyl ligands. Each shows only two absorption bands below 1950 cm−1 due to increased back-donation from metals, as each metal atom is now bonded to a single CO ligand. While it was obvious that more carbonyls had been lost than might be expected, the precise attachment of ligands to the triruthenium center was unclear, and consequently we have carried out X-ray structure determinations of 7 and 8, the results of which are shown in 2549

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

Figure 4. Solid-state molecular structure of Ru3(CO)3(μ-pdt)2{μ-P(C4H3S)2}(μ-dppm)(μ-H) (8), showing (a) 50% probability thermal ellipsoids and (b) the cluster core with phenyl and thienyl groups removed for clarity. Selected interatomic distances (Å) and angles (deg): Ru(1)−Ru(2) = 3.1095(4), Ru(1)−Ru(3) = 2.9002(4), Ru(2)−Ru(3) = 2.8539(4), Ru(1)−P(1) = 2.3111(7), Ru(1)−P(2) = 2.3648(7), Ru(2)−P(3) = 2.3218(7), Ru(3)−P(2) = 2.3185(7), Ru(1)−S(1) = 2.4274(7), Ru(2)−S(1) = 2.4180(7), Ru(1)−S(2) = 2.4190(7), Ru(3)−S(2) = 2.3825(7), Ru(3)−S(3) = 2.4018(7), Ru(2)−S(3) = 2.3967(7), Ru(3)−S(4) = 2.4099(7), Ru(2)−S(4) = 2.4203(7); Ru(3)−Ru(1)−Ru(2) = 56.576(10), Ru(1)−Ru(2)− Ru(3) = 58.010(7), Ru(1)−Ru(3)−Ru(2) = 65.414(9), P(1)−Ru(1)−P(2) = 100.57(2), P(2)−Ru(1)−Ru(3) = 51.023(17), P(1)−Ru(1)−Ru(2) = 89.133(18), Ru(1)−P(2)−Ru(3) = 76.52(2), S(1)−Ru(2)−S(3) = 147.65(2), S(4)−Ru(2)−S(1) = 92.42(2), Ru(3)−Ru(1)−S(1) = 95.599(17), S(2)−Ru(1)−Ru(2) = 93.241(17), S(1)−Ru(1)−Ru(2) = 49.946(15), S(2)−Ru(1)−Ru(3) = 52.270(16), Ru(2)−Ru(1)−S(4) = 92.145(17), S(1)−Ru(2)−Ru(3) = 97.014(17), S(3)−Ru(3)−Ru(1) = 103.958(19), S(4)−Ru(3)−Ru(2) = 53.950(16), S(4)−Ru(3)−Ru(1) = 97.692(18), S(2)−Ru(1)−S(1) = 96.33(2), Ru(3)−S(2)−Ru(1) = 74.311(19), Ru(3)−S(4)−Ru(2) = 72.43(2), Ru(2)−S(1)−Ru(1) = 79.84(2), Ru(3)−S(3)− Ru(2) = 72.99(2), P(3)−Ru(2)−S(4) = 170.95(2), P(2)−Ru(3)−S(2) = 74.51(2).

L−S) (L = alkyl, aryl) in which the dithiolato ligand spans two separate trirthenium cores. Introduction of a dppm ligand generally results in a stabilization of the triruthenium core,22 but Ru3(CO)10(μ-dppm) gives only the dinuclear complexes Ru2(CO)4(μ-S−L−S)(μ-dppm)21b upon reaction with dithiols. To our knowledge there are only two other examples of triruthenium carbonyl clusters containing two dithiolato ligands: Ru3(CO)7(μ-edt)2,23 reported by Adams from the reaction between Ru3(CO)12 and 1,2,5,6-tetrathiacylooctane, and a related diphosphine derivative, Ru3(CO)5(μ-pdt)2(κ2dppe)24 (dppe = Ph2PCH2CH2PPh2), prepared from the reaction of Ru3(CO)10(μ-dppe) and 1,3-propanedithiol. Both of these, like 7−10, are characterized by a TEC of 50, but unlike the former they contain a linear trinuclear core with two short and one long (nonbonding) ruthenium−ruthenium interaction. The mode of formation of 7−10 is unclear. In no instance did we find evidence of a product resulting from the addition of a single dithiol, which suggests that binding of the first ligand activates the resultant product toward further carbonyl loss and thus dithiol addition. In light of the isolated structures of 3−6 it is tempting to suggest that a related dithiolato-capped triruthenium cluster is the initial reaction product. Unlike 3− 6, the anti arrangement of substituents cannot be achieved with the bridging ligands and it may be this constraint which further activates the initial products toward later reactivity. Reactions of 1 and 2 with Benzeneselenol. The reactivity of selenols toward low-valent metal centers has not

into two distinct categories; those bridged by the phosphido and μ2-dithiolato ligands are in the standard range (Ru(1)− Ru(3) = 2.9056(8) Å and Ru(2)−Ru(3) = 2.8537(8) Å in 7; Ru(1)−Ru(3) = 2.9002(4) Å and Ru(2)−Ru(3) = 2.8539(4) Å in 8), while the vector spanned by the diphosphine is significantly longer (Ru(1)−Ru(2) = 3.1193(8) and 3.1095(4) Å). Similar patterns of ruthenium−ruthenium bond lengths have been noted in 64-electron butterfly clusters, such an arrangement of metal atoms generally being associated with a 62-electron count.19 EHMO calculations on model complexes suggest that in this case the additional electron pair occupies an orbital which has ruthenium−ruthenium antibonding character.19,20 In order to gain more insight into the bonding in 7−10, we have carried out density functional theory (DFT) calculations on 7. The computed geometry of 7 is in good agreement with the X-ray structure and the trend in ruthenium−ruthenium distances is confirmed, although they are overestimated by ca. 0.06−0.12 Å. The calculated Ru(1)− Ru(2) distance is 3.244 Å, while Ru(1)−Ru(3) and Ru(2)− Ru(3) are 2.970 and 2.936 Å, respectively. The longer Ru(1)− Ru(2) distance is consistent with a nonbonding interaction between the two ruthenium atoms, which are held close together by the bridging thiolate and diphosphine ligands. Clusters 7 and 8 are rare examples of triruthenium clusters containing two dithiolato ligands. The direct reaction between Ru3(CO)12 and dithiols affords two product types: binuclear complexes Ru2(CO)6(μ-S−L−S)21 resulting from loss of one metal atom and hexanuclear clusters {(μ-H)Ru3(CO)10}2(μ-S− 2550

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

substituent on Se(2) is oriented at approximately right angles to the others. This is the group which spans the same ruthenium−ruthenium edge as the diphosphine, and thus we have a situation akin to that seen in 7 and 8, whereby the geometry is controlled by the need to minimize steric interactions with the phenyl groups of the diphosphine. The latter thus spans across the long ruthenium−ruthenium edge, while the phosphido ligand symmetrically bridges the Ru(1)− Ru(2) edge (Ru(2)−P(3) = 2.313(3) Å and Ru(1)−P(3) = 2.318(3) Å). The position of the hydride was not determined by crystallographic analysis but is believed to bridge across the Ru(2)−Ru(3) edge, as this is spanned by only a single benzeneselenolato ligand. The coordination sphere is completed by four carbonyl ligands. Considering the bridging selenolato ligands to be 3-electron donors but the terminally bound only a 1-electron donor, the cluster is characterized by a TEC of 50 electrons. The situation is then akin to that found in 7 and 8 and highlights the flexibility of the ruthenium− ruthenium bonds in these multibridged systems. It also highlights the danger of interpreting bonding and nonbonding contacts in terms of lines drawn between atoms, as each program has a semiarbitrary cutoff for all types of element− element interactions. Spectroscopic data for 11 are consistent with the solid-state structure. Specifically, the hydride was observed as a doublet at δ −20.04 (J = 29.0 Hz) and three inequivalent phosphorus environments were observed in the 31 P{ 1 H} NMR spectrum. The thienyl cluster 13 was spectroscopically very similar to 11; both showed three clear carbonyl absorptions in the IR spectrum and again a single hydride resonance was seen in the 1H NMR spectrum (δ −19.63 (d, J = 27.0 Hz)). On this basis it is clear that 13 is an analogue of 11. The characterization of 12, the second product of the reaction of PhSeH with 1, proved more problematic, as repeated attempts to grow single crystals were unsuccessful. The IR spectrum contains only two carbonyl bands at 1941 and 1922 cm−1, being very similar to the situation found in 7 and 8, indicating they have a similar distribution of carbonyl ligands. The FAB mass spectrum displays an ion at m/z 1561 and further ions due to sequential loss of three carbonyls, which supports the proposed formulation. Hence, we propose that 12 is related to 11 by carbonyl loss. There are a number of possible structures for 12, among the most obvious being (i) a 48electron cluster with retention of the terminal selenolato ligand and formation of a new metal−metal bond and (ii) a 50electron cluster with four bridging selenolato ligands. We favor this second scenario, and one plausible structure is shown in Scheme 4, whereby loss of a carbonyl is compensated for by the conversion of the terminal selenolato ligand to a bridging mode. This requires a concomitant shift of the hydride as shown. Thus, the structure is envisaged to be very similar to those of 7−10. Both the 1H and 31P{1H} NMR spectra of 12 show the existence of two isomeric forms in solution. The 31 1 P{ H} NMR spectrum displays six resonances in two sets: a singlet and two doublets at δ 42.2, 35.3, and −5.5 having similar intensities, with three equally intense multiplets at δ 38.1, 35.6, and −7.6. The intensity ratio between the two sets is approximately 5:1. In accord with this, the hydride region of the 1H NMR spectrum shows two doublets at δ −14.15 and −15.28 with an intensity ratio of 5:1, but the JP−H values for these are similar (JP−H = 34.0 Hz). The migration of the hydride ligand along the ruthenium−ruthenium edges is unlikely for 12, as each edge of the metal triangle is bridged

been extensively studied. Addition of PhSeH to Os3(CO)10(MeCN)2 affords the expected Se−H oxidative addition product Os3(CO)10(μ-H)(μ-SePh),25 and an analogous transformation is noted with Os3(CO)10(μ-dppm).26 Reactions of Ru3(CO)12 or its derivatives with PhSeH appear not to have been reported, although triruthenium selenolato complexes are available upon addition of Ph2Se2 to both Ru3(CO)1227 and Ru3(CO)10(μ-dppm).28 Addition of PhSeH to Fe3(CO)12 is reported to proceed smoothly in the presence of triethylamine, affording [Fe2(CO)6(μ-CO)(μ-SePh)][NEt3H] possibly via a hydride intermediate.29 Reactions of 1 and 2 with PhSeH do not parallel those with thiols. Thus, under similar conditions the furyne cluster 1 reacts with PhSeH to give Ru 3 (CO) 4 (κ 1 -SePh)(μ-SePh)3 {μ-P(C 4 H 3 O) 2 }(μdppm)(μ-H) (11; 20% yield) and Ru3(CO)3(μ-SePh)4{μP(C4H3O)2}(μ-dppm)(μ-H) (12; 27% yield), while the thiophyne cluster 2 afforded Ru3(CO)4(κ1-SePh)(μ-SePh)3{μP(C4H3S)2}(μ-dppm)(μ-H) (13; 14% yield), an analogue of 11, together with Ru 3 (CO) 5 {PPh(C 4 H 3 S) 2 }(μ 3 -Se) 2 (μSePh)2(μ-dppm) (14) (16% yield), which results from selenium−carbon bond cleavage (Scheme 4). Scheme 4

Characterization of 11 and 13 was based predominantly on the single-crystal X-ray diffraction analysis of 11, the results of which are summarized in Figure 5. The molecule consists of a triangular array of ruthenium atoms characterized by three distinct ruthenium−ruthenium interactions: Ru(1)−Ru(2) = 2.958(1) Å, Ru(2)−Ru(3) = 3.131(1) Å, and Ru(1)−Ru(3) = 3.242(1) Å. The first two of these are within the expected range for ruthenium−ruthenium bonds, while the latter falls outside of this (see later). There are four benzeneselenolato ligands: three span ruthenium−ruthenium vectors and lie on the same face of the triruthenium center. The fourth is an extremely rare example of a monodentate selenolato bound to a cluster center.16c The ruthenium−selenium distances for the bridging selenolato ligands range from 2.535(2) to 2.576(2) Å, and while that to the terminal ligand (Ru(3)−Se(4) = 2.514(2) Å) is shorter, the difference is only slight. The four selenium atoms lie approximately in a plane which sits parallel to the triruthenium plane, and there are a series of relatively short selenium−selenium contacts (Se···Se = 3.226(3)−3.619(3) Å). The phenyl groups of three of them (Se(1), Se(3), and Se(4)) also lie approximately in this plane, with the lone pairs thus being oriented away from the cluster core, while in contrast the 2551

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

Figure 5. Solid-state molecular structure of one molecule of Ru3(CO)4(SePh)(μ-SePh)3{μ-P(C4H3S)2}(μ-dppm)(μ-H) (11), showing (a) the full molecule and (b) the cluster core with phenyl and thienyl groups removed from the phosphines for clarity. Selected interatomic distances (Å) and angles (deg): Ru(2)−Ru(3) = 3.1311(14), Ru(1)−Ru(2) = 2.9575(13), Ru(1)−P(1) = 2.320(3), Ru(2)−P(3) = 2.313(3), Ru(1)−P(3) = 2.318(3), Ru(3)−P(2) = 2.326(3), Ru(1)−Se(2) = 2.5511(16), Ru(2)−Se(1) = 2.5381(16), Ru(3)−Se(2) = 2.5760(15), Ru(3)−Se(4) = 2.5143(15), Ru(1)−Se(1) = 2.5415(15), Ru(2)−Se(3) = 2.5351(16), Ru(3)−Se(3) = 2.5567(15); Ru(1)−Ru(2)−Ru(3) = 64.27(3), Ru(1)−Se(2)−Ru(3) = 78.44(5), Ru(2)−Se(1)−Ru(1) = 71.22(4), Ru(2)−Se(3)−Ru(3) = 75.89(4), Se(1)−Ru(1)−Ru(2) = 54.34(4), Se(1)−Ru(2)−Ru(1) = 54.45(4), Se(3)−Ru(3)−Ru(2) = 51.74(4), Se(3)−Ru(2)−Ru(3) = 52.36(3), Se(2)−Ru(1)−Ru(2) = 101.18(4), Se(1)−Ru(2)−Ru(3) = 88.28(4), Se(4)− Ru(3)−Ru(2) = 142.79(5), Se(2)−Ru(3)−Ru(2) = 96.20(4), Se(3)−Ru(2)−Ru(1) = 100.15(5), Se(3)−Ru(2)−Se(1) = 78.97(5), Se(4)−Ru(3)− Se(2) = 88.88(5), Se(1)−Ru(1)−Se(2) = 85.11(5), Se(4)−Ru(3)−Se(3) = 91.05(5), Se(3)−Ru(3)−Se(2) = 99.29(5), Ru(2)−P(3)−Ru(1) = 79.38(11), P(3)−Ru(1)−P(1) = 98.92(11), P(1)−Ru(1)−Ru(2) = 122.65(8), P(3)−Ru(1)−Ru(2) = 50.24(8), P(3)−Ru(2)−Ru(1) = 50.38(8), P(2)−Ru(3)−Ru(2) = 121.20(8).

Figure 6. Two views of the molecular structure of Ru3(CO)5{PPh(C4H3S)2}(μ3-Se)2(μ-SePh)2(μ-dppm) (14).

selenido ligand to be a 4-electron donor, then the cluster has a TEC of 54 electrons and thus metal−metal interactions are not expected. The structure of 14 is very similar to that of Ru3(CO)6(μ3-Se)2(μ-SePh)2(μ-dppm),12a except that a carbonyl group is replaced by monodentate phosphine. The 31P{1H} NMR spectrum of 14 shows two multiplets at δ 27.0 and 23.4 with a relative intensity of 2:1. In addition to aromatic proton resonances, the 1H NMR spectrum displays two multiplets at δ 3.74 and 3.38, each integrating to one proton, attributable to the methylene protons of the dppm ligand. Clusters 11−14 all result from the oxidative addition of 4 equiv of PhSeH to the triruthenium center. In light of the addition of PhSH to 1 and 2, it is envisaged that bis(selenolato)

by at least two bridging ligands. Therefore, we speculate that the different orientation of the phenyl groups on the selenium atoms is responsible for this isomerization. Characterization of 14 was made on the basis of a partially resolved single-crystal X-ray diffraction analysis (Figure 6). While the refinement of this structure is poor (ca. 10%), it gives sufficient information about the geometry of the cluster and the orientation of the ligands on the cluster surface. It consists of an Ru3Se2 core, the two selenido ligands capping the open triruthenium moiety. Selenolate ligands bridge two of the edges, while the third is spanned by the diphosphine. Five terminal carbonyls and a phenylbis(2-thienyl)phosphine ligand complete the coordination sphere. Thus considering each 2552

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

Scheme 5

Thus as already discussed, initial CO loss and coordination of the thiol leads to activation of the S−H bond. This can then undergo oxidative addition to the metal center, and the later rearrangement of the terminal thiolato ligand into the bridging site activates proton transfer to the alkyne, generating an alkenyl ligand. Such species are not isolated from reactions of 1 and 2 but are seen upon addition of Ph2Te2 to 1 (Scheme 1).14 In the related dimolybdenum chemistry they are accessible via a different route, namely the insertion of an alkyne into Mo2(μH)(μ-SR)(CO)4Cp2, but are difficult to isolate, as they are readily oxidized.35 A repeat of these transformations upon addition of another 1 equiv of thiol then generates a metalbound alkene which is likely to be readily lost, producing a bis(thiolato) complex, examples of which are Ru3(CO)5(μSR)2{μ-P(C4H3E)2}(μ-dppm)(μ-H) (3−6). Morris does not generally isolate related species, Mo2(μ-SR)2(CO)2Cp2, although again they are known,36 but in one instance an alkene complex, viz. [Mo2(μ-S)2(μ-MeO2CCHCHCO2Me)Cp2], is isolated33 although carbon−sulfur bond scission has also occurred. In the case of the reactions of 1 and 2 with dithiols and PhSeH, then these putative intermediates must undergo further facile reactions with E−H bonds. The likely course of these reactions is less clear. Oxidative addition of the E−H bond adds a further two electrons to the cluster center, and this may well be compensated for by the temporary cleavage of a metal−metal bond. What is clear from the chemistry desribed herein is that the ruthenium−ruthenium interactions are extremely soft and easily deformed, with lengths ranging over ca. 0.5 Å. Molybdenum−molybdenum interactions are similarly known to vary over a similar range, and thus it is tempting to suggest that this is a necessary prerequisite for the facile oxidative addition reactions. A further requirement is the ready loss of CO. Generally as carbonyls are lost from a low-valent metal center, the degree of π back-bonding to those remaining increases and thus the metal−carbon bond becomes stronger. Thiolato and selenolato groups are relatively poor donor ligands and hence may not affect the remaining carbonyls to a great extent. It is useful to contrast this with Morris’s observation that reaction of Mo2(μ-C2R2)(CO)4Cp2 (R = CO2Me) with Ph2PH affords only a product from the addition of a single P−H bond, presumably since the remaining carbonyls are now too strongly held to the dimolybdenum core.34 A further observation by Morris is that addition of i PrSH to Mo2(μ-C2Ph2)(CO)4Cp2 affords as the only product the parallel alkyne complex Mo2(μ-C2Ph2)(μ-SiPr)2(μ-S)Cp2,32 being possibly rationalized by the known reversibility of the

complexes of the type Ru3(CO)5(μ-SeR)2{μ-P(C4H3E)2}(μdppm)(μ-H) are intemediates. Formation of 14 results from further carbon−selenium bond cleavage and carbon−phosphorus bond formation. The cluster-mediated scission of carbon− selenium bonds is precedented. Adams has reported that heating Os 3 (CO) 10 (μ-H)(μ-SePh) at 160 °C affords Os3(CO)9(μ3-Se)2 along with benzene and Os(CO)5.25 Even more pertinent to this work is the report by Deeming and coworkers30 that Os3(CO)10(μ-SePh)2 rearranges in refluxing cyclohexane to afford Os3(CO)8(μ3-Se2)(μ-Ph){μ-PhC(O)}, in which the two cleaved phenyl groups are captured by the cluster. In the case of 14 only one of the cleaved phenyl groups is maintained via formation of a new carbon−phosphorus bond, the fate of the second remaining unknown. Interestingly, Deeming’s acyl cluster is also characterized by a TEC of 54 electrons.30



SUMMARY AND COMMENTS

We have demonstrated in this work that the triruthenium furyne and thiophyne clusters Ru 3 (CO) 7 (μ-H)(μ 3 -η 2 C4H2E){μ-P(C4H3E)2}(μ-dppm) (1, 2; E = O, S) display high reactivity toward reagents with group 16 E−H bonds at moderate temperatures. Thus, while Ru3(CO)12 and its dppm derivative Ru3(CO)10(μ-dppm) react with 1 equiv of thiol under moderate conditions,10,11 these clusters add at least two thiolate ligands, while with dithiols and PhSeH the oxidative addition of four E−H bonds is noted. The reason for this enhanced reactivity is almost certainly the presence of the coordinated alkyne. Morris and co-workers have extensively studied reactions of dimolybdenum alkyne complexes Mo2(μC2R2)(CO)4Cp2 (Cp = η-C5H5) with thiols,31−34 the products of which are strongly dependent upon the nature of the alkyne substituents. Thus, when one of these is −CO2Me, alkenyl complexes are isolated resulting from oxidative addition of the sulfur−hydrogen bond and transfer of the proton to the alkyne.33,34 Generally such complexes are not isolated with other alkyne substituents, but notably reaction of tBuSH with Mo2(μ-C2Me2)(CO)4Cp2 does afford Mo2(μ-MeCCHMe)(μ-StBu)(CO)3Cp2.31 With less sterically demanding thiols (R = Et, iPr) the same alkyne complex reacts with an excess at 110 °C to afford Mo2(μ-SR)2(μ-S)2Cp2, resulting from the complete displacement of all four carbonyls.31 This behavior is akin to that observed in our work. On the basis of our work and that of Morris, we propose a general scheme for the reactivity of such alkyne complexes with thiols (Scheme 5). 2553

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

plates prepared from silica gel GF254 (type 60, E. Merck). All new complexes are air-stable in the solid state and in solution over moderate periods. Reaction of 1 with PhSH. Excess thiophenol (ca. 10-fold) was added to a benzene solution (10 mL) of 1 (50 mg, 0.045 mmol), and the mixture was heated to reflux for 24 h. The solvent was removed under reduced pressure and the residue separated by TLC on silica gel. Elution with hexane/CH2Cl2 (3/2, v/v) developed four bands. The first band was unreacted 1 (trace), while the second band yielded Ru3(CO)5(μ-SPh)2{μ-P(C4H3O)2}(μ-dppm)(μ-H) (3; 18 mg, 33%) as orange crystals after recrystallization from hexane/CH2Cl2 at 4 °C. The contents of the other two bands were too small for characterization. Data for 3 are as follows. Anal. Calcd for C50H39O7P3Ru3S2: C, 49.55; H, 3.09. Found: C, 49.88; H, 3.15. IR (ν(CO), CH2Cl2): 2024 s, 2000 s, 1964 s, 1946 m, 1924 m cm−1. 1H NMR (CDCl3): δ 7.75 (m, 2H), 7.62 (m, 5H), 7.46 (m, 2H), 7.31− 7.14 (m, 14H), 7.09 (m, 3H), 7.01 (m, 4H), 6.90 (m, 2H), 6.58 (m, 2H), 6.42 (m, 1H), 6.15 (m, 1H), 4.00 (m, 1H), 3.15 (m, 1H), −16.39 (dd, J = 40.0, 22.0 Hz, 1H). 31P{1H} NMR (CDCl3): δ 115.5 (d, J = 22.4 Hz, 1P), 36.5 (dd, J = 51.5, 22.4 Hz, 1P), 19.2 (d, J = 51.5 Hz, 1P). Reaction of 2 with PhSH. Excess PhSH (ca. 10-fold) was added to a benzene solution (10 mL) of 2 (50 mg, 0.043 mmol), and the mixture was heated to reflux for 8 h. The solvent was removed under vacuum and the residue chromatographed by TLC on silica gel. Elution with hexane/CH2Cl2 (7/3, v/v) developed only one band, which gave Ru3(CO)5(μ-SPh)2{μ-P(C4H3S)2}(μ-dppm)(μ-H) (4; 30 mg, 56%) as orange crystals after recrystallization from hexane/ CH2Cl2 at 4 °C. Data for 4 are as follows. Anal. Calcd for C50H39O5P3Ru3S4: C, 48.27; H, 3.16. Found: C, 48.53; H, 3.23. IR (ν(CO), CH2Cl2): 2021 s, 1998 s, 1963 m, 1944 m, 1922 m cm−1. 1H NMR (CDCl3): δ 7.88 (m, 2H), 7.84 (m, 5H), 7.77−7.50 (m, 26H), 7.42 (m, 2H), 7.31 (m, 1H), 3.94 (m, 1H), 3.23 (m, 1H), −16.38 (dd, J = 39.6, 22.4 Hz, 1H). 31P{1H} NMR (CDCl3): δ 128.2 (d, J = 16.0 Hz, 1P), 35.3 (dd, J = 51.8, 16.0 Hz, 1P), 16.6 (d, J = 51.8 Hz, 1P). Reaction of 1 with iPrSH. Excess iPrSH (ca. 10-fold) was added to a benzene solution of 1 (50 mg, 0.045 mmol) and heated to reflux for 15 h. The solvent was removed under reduced pressure and the residue chromatographed by TLC on silica gel. Elution with hexane/ CH2Cl2 (1/1, v/v) developed two bands. The second band was unreacted 1 (trace), while the first band afforded Ru3(CO)5{μSCH(CH3)2}2{μ-P(C4H3O)2}(μ-dppm)(μ-H) (5; 30 mg, 59%) as red crystals after recrystallization from hexane/CH2Cl2 at 4 °C. Data for 5 are as follows. Anal. Calcd for C44H43O7Ru3P3S2: C, 46.20; H, 3.80. Found: C, 46.82; H, 3.98. IR (νCO, CH2Cl2): 2017 s, 1994 s, 1957 s, 1937 m, 1918 m cm−1. 1H NMR (CDCl3): δ 7.60 (m, 4H), 7.50 (m, 2H), 7.25 (m, 11H), 6.99 (m, 3H), 6.88 (m, 2H), 6.55 (m, 2H), 6.37 (m, 1H), 6.16 (m, 1H), 4.0 (m, 1H), 3.02 (m, 1H), 2.71 (m, 1H), 2.50 (m, 1H), 1.55 (m, 12H), −16.70 (dd, J = 40.4, 22.0 Hz, 1H). 31P{1H} NMR: δ 113.6 (d, J = 18.0 Hz, 1P), 38.6 (dd, J = 52.0, 18.0 Hz, 1P), 17.7 (d, J = 52.0 Hz, 1P). MS (FAB): m/z 1143 (M+). Reaction of 2 with iPrSH. Excess iPrSH (ca. 10-fold) was added to a benzene solution of 2 (30 mg, 0.027 mmol) and heated to reflux for 15 h. The solvent was removed under reduced pressure and the residue chromatographed by TLC on silica gel. Elution with hexane/ CH2Cl2 (1/1, v/v) developed two bands. The second band was unreacted 2 (trace), while the first band afforded Ru3(CO)5{μSCH(CH3)2}2{μ-P(C4H3S)2}(μ-dppm)(μ-H) (6; 20 mg, 66%) as red crystals after recrystallization from hexane/CH2Cl2 at 4 °C. Data for 6 are as follows. Anal. Calcd for C44H43O5Ru3P3S4: C, 44.92; H, 3.69. Found: C, 44.80; H, 3.45. IR (νCO, CH2Cl2): 2015 s, 1994 s, 1956 s, 1936 m, 1917 m cm−1. 1H NMR (CDCl3): δ 7.66−6.61 (m, 26H), 3.89 (m, 1H), 3.09 (m, 1H), 2.79 (m, 1H), 2.85 (m, 1H), 1.64 (m, 12H), −16.64 (dd, J = 41.2, 22.8 Hz, 1H). 31P{1H} NMR: δ 125.5 (d, J = 28.0 Hz, 1P), 36.3 (dd, J = 86.0, 28.0 Hz, 1P), 13.6 (d, J = 86.0 Hz, 1P). MS (FAB): m/z 1175 (M+). Reaction of 1 with 1,3-Propanedithiol. 1 (75 mg, 0.067 mmol) was dissolved in benzene (15 mL), and to the resultant solution was added excess 1,3-propanedithiol (ca. 10-fold). The mixture was boiled under reflux for 15 h. The solvent was removed by rotary evaporation

proton addition to an alkyne, as previously noted at a diiron center.37 The latter also results from a sulfur−carbon bond scission process which mirrors the double selenium−carbon bond cleavage noted in the formation of Ru3(CO)5{PPh(C4H3S)2}(μ3-Se)2(μ-SePh)2(μ-dppm) (14). The variations in reactivity observed by Morris and coworkers as a function of the substituents on the coordinated alkyne31−34 led us to consider if the furyne and thiophyne ligands in 1 and 2, respectively, were a prerequisite for the reactivity observed or whether any alkyne complex of the type Ru3(CO)7(μ-H)(μ3-η2-C2R2){μ-P(C4H3E)2}(μ-dppm) would suffice. At this time we cannot answer this question, since analogues of 1 and 2 are currently inaccessible, although some of our earlier work with 1 and 2 suggests that they are highly flexible ligands. More generally, complexes with a Ru3(μ3alkyne) core might be expected to show this behavior: viz., the multiple oxidative addition of E−H bonds. One candidate worthy of exploration is Ru3(CO)7(μ3-η2-C2Ph2)(μ-dppm),38 which like 1 and 2 contains a 6-electron-donor alkyne and has been shown to add H2 to afford Ru3(CO)7(μ-H)2(μ3-η2C2Ph2)(μ-dppm). This demonstrates that oxidative addition is facile but perhaps in the absence of the extra lone pair of the thiolato/selenolato ligand proton transfer to the alkyne is not activated. We are currently exploring reactions of this type. As far as we aware, clusters 1 and 2 are the first examples of triruthenium systems that facilitate multiple additions of thiols, ditiols, and selenols through S−H and Se−H bond activation. Hitherto such multiple additions of thiolato or selenolato ligands into triruthenium carbonyl clusters were only achieved by the oxidative addition of RE−ER (E = S, Se).12g,h,28 In somewhat related work over the past decade, Adams and coworkers have introduced multiple tin and germanium atoms into transition-metal cluster cores using Ph3SnH and Ph3GeH as the main-group-element source.39 For example, heating H4Ru4(CO)12 with excess Ph3GeH at 175 °C produces Ru4(CO)8(μ-GePh2)4(μ4-GePh)2 in high yields.39g Such clusters are potentially useful as precursors to heterometallic nanoparticles, as the ratio of different metal types is tunable in order to optimize catalytic performance. Likewise, related sulfur- and/or selenium-containing clusters may be valuable precursors to ruthenium sulfides or selenides for applications in microelectronics or heterogeneous catalysis.11c



EXPERIMENTAL SECTION

All reactions were carried out under a nitrogen atmosphere using standard Schlenk techniques unless otherwise stated. Reagent-grade solvents were dried using appropriate drying agents and distilled by standard methods prior to use. Infrared spectra were recorded on a Shimadzu FTIR 8101 spectrophotometer. NMR spectra were recorded on Bruker DPX 400 and Varian Inova 500 instruments. Elemental analyses were performed by Microanalytical Laboratories, University College London. Fast atom bombardment mass spectra were obtained on a JEOLSX-102 spectrometer using 3-nitrobenzyl alcohol as matrix and CsI as calibrant. Ru3(CO)12 was purchased from Strem Chemicals Inc. and used without further purification. Me3NO·2H2O was purchased from Lancaster; the water was removed using a Dean− Stark apparatus by azeotropic distillation from benzene, and the anhydrous Me3NO was stored under nitrogen. Tris(2-furyl)phosphine and tris(2-thienyl)phosphine were purchased from E. Merck and bis(diphenylphosphino)methane, PhSH, 1,3-propanedithiol, 1,2-ethanedithiol and PhSeH from Aldrich, and all were used as received. The starting clusters Ru3(CO)7(μ-H)(μ3-η2-C4H2O){μ-P(C4H3O)2}(μdppm) (1) and Ru3(CO)7(μ-H)(μ3-η2-C4H2S){μ-P(C4H3S)2}(μdppm) (2) were prepared according to literature methods.13 Preparative thin-layer chromatography was carried out on 1 mm 2554

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

δ 8.29 (m, 4H), 7.68 (m, 5H), 7.29 (m, 7H) 6.91 (m, 24H), 6.62 (m, 2H), 6.24 (m, 1H), 6.06 (m, 1H), 5.90 (m, 1H), 5.04 (m, 1H), 3.79 (m, 1H), 3.65 (m, 1H), −20.04 (d, J = 29.0 Hz, 1H). 31P{1H} NMR (CDCl3): δ 36.5 (s, 1P), 21.2 (d, J = 49.0 Hz, 1P), 1.6 (d, J = 49.0 Hz, 1P). Data for 12 are as follows. Anal. Calcd for C60H49O5P3Ru3Se4: C, 46.15; H, 3.17. Found: C, 46.49; H, 3.23. IR (ν(CO), CH2Cl2): 1941 s, 1922 m cm−1. 1H NMR (CDCl3): major isomer: δ 7.93 (m, 2H), 7.61 (m, 4H), 7.45 (m, 11H) 7.27 (m, 4H), 7.07 (m, 8H), 6.91 (m, 6H) 6.71 (m, 2H), 6.64 (m, 2H), 6.52 (m, 3H), 6.21 (m, 1H), 5.94 (m, 1H), 5.76 (m, 1H), 5.10 (m, 1H), 3.83 (m, 1H), 3.72 (m, 1H); hydride region, major isomer, −14.15 (d, J = 34.0 Hz, 1H), minor isomer, −15.28 (d, J = 34.0 Hz, 1H). 31P{1H} NMR (CDCl3): major isomer, δ 42.2 (s, 1P), 35.3 (d, J = 53.3 Hz, 1P), −5.5 (d, J = 53.3 Hz, 1P); minor isomer, δ 38.1 (m), 35.6 (m), −7.6 (m). MS (FAB): m/z 1561 (M+). Reaction of 2 with PhSeH. A benzene solution (15 mL) of 2 (50 mg, 0.043 mmol) and excess PhSeH was boiled under reflux for 15 h. The solvent was removed under vacuum and the residue separated by TLC on silica gel. Elution with hexane/CH2Cl2 (3/2, v/v) developed three bands. The second and fourth bands yielded Ru3(CO)5{PPh(C4H3S)2}(μ3-Se)2(μ-SePh)2(μ-dppm) (14; 11 mg, 16%) as orange crystals and Ru3(CO)4(κ1-SePh)(μ-SePh)3{μ-P(C4H3S)2}(μ-dppm)(μ-H) (13; 10 mg, 14%) as green crystals after recrystallization from hexane/CH2Cl2 at 4 °C. The content of the third band was too small for characterization. Data for 13 are as follows. Anal. Calcd for C61H49O4P3Ru3S2Se4: C, 45.17; H, 3.05. Found: C, 45.51; H, 3.12. IR (ν(CO), CH2Cl2): 2032 s, 1977 m, 1947 m cm−1. 1H NMR (CDCl3): δ 8.35 (m, 2H), 8.24 (m, 2H), 7.66 (m, 4H), 7.46 (m, 2H), 7.38 (m, 2H), 7.29 (m, 2H), 7.19 (m, 10H), 7.04 (m, 10H), 6.95 (m, 6H) 6.73 (m, 6H), 3.72 (m, 1H), 3.43 (m, 1H), −19.63 (d, J = 27.0 Hz, 1H). 31 1 P{ H} NMR (CDCl3): δ 36.5 (s, 1P), 21.2 (d, J = 49.0 Hz, 1P), 1.6 (d, J = 49.0 Hz, 1P). Data for 14 are as follows. Anal. Calcd for C56H43O5P3Ru3S2Se4: C, 42.79; H,2.76. Found: C,43.13; H, 2.83. IR (ν(CO), CH2Cl2): 2023 s, 1966 m, 1934 w cm−1. 1H NMR (CDCl3): δ 8.31 (m, 1H), 8.21 (m, 5H), 7.51 (m, 6H), 7.39 (m, 10H), 7.22 (m, 8H), 7.18 (m, 2H), 7.02 (m, 8H), 6.66 (m, 1H), 3.74 (m, 1H), 3.38 (m, 1H). 31P{1H} NMR (CDCl3): δ 27.0 (d, J = 15.2 Hz, 2P), 23.4 (t, J = 15.2 Hz, 1P). X-ray Structure Determinations. Single crystals of 3, 4, 7, 8, and 11 suitable for X-ray diffraction were obtained by recrystallization from hexane/CH2Cl2 at 4 °C and mounted on nylon fibers with a mineral oil, and diffraction data were collected at low temperatures on a Bruker SMART APEX CCD diffractometer using Mo Kα radiation (λ = 0.710 73 Ǻ ). Data reduction and integration were carried out with SAINT+, and absorption corrections were applied using the program SADABS.40 Structures were solved by direct or Patterson methods and developed using alternating cycles of least-squares refinement and difference-Fourier synthesis. All non-hydrogen atoms were refined anisotropically. Hydrogen atoms were placed in calculated positions, and their thermal parameters were linked to those of the atoms to which they were attached (riding model). The SHELXTL PLUS V6.10 program package was used for structure solution and refinement.41 For 3 there are two independent clusters and also two molecules of dichloromethane in the asymmetric unit. The metal-bound hydride was not located. Crystals of 4 rapidly desolvated. Consequently the dichloromethane solvate was refined only isotropically. One carbon atom, C(8), of a thienyl ring was also only refined isotropically. The metal-bound hydride was located in a difference map and refined isotropically. Crystals of 7 contained one molecule of dichloromethane in the asymmetric unit and were only weakly diffracting. The metalbound hydride was not located. Crystals of 8 contained one molecule of dichloromethane in the asymmetric unit, which was refined anisotropically. In one ring there was a disorder of sulfur S(5) and carbon C(11) over two sites. This was modeled with occupancies of 60 and 40%, respectively. Due to this, no hydrogen atoms were generated attached to C(11)−C(13). The metal-bound hydride was not located. Crystals of 11 contained two independent clusters in the asymmetric unit along with four molecules of toluene. Only the heavy atoms Ru, P, and Se were refined anisotropically. Diffraction data were collected on a small orange crystal of 12·3H2O. The following monoclinic cell

and the residue separated by TLC on silica gel. Elution with hexane/ CH2Cl2 (3/2, v/v) developed two bands. The second band afforded Ru3(CO)3(μ-pdt)2{μ-P(C4H3O)2}(μ-dppm)(μ-H) (7; 30 mg, 39%) as orange crystals after recrystallization from hexane/CH2Cl2 at 4 °C. Data for 7 are as follows. Anal. Calcd for C42H41O5P3Ru3S4: C, 43.86; H, 3.60. Found: C, 44.13; H, 3.65. IR (ν(CO), CH2Cl2): 1942 s, 1917 m cm−1. 1H NMR (CDCl3): δ 7.67 (m, 3H), 7.51 (m, 2H), 7.39−7.06 (m, 14H), 6.96 (m, 2H), 6.82 (m, 2H), 6.41 (m, 1H), 6.09 (m, 1H), 6.02 (m, 1H), 3.65 (m, 1H), 3.10 (m, 1H), 2.69 (m, 1H), 2.53 (m, 1H), 2.47 (m, 1H), 2.19 (m, 3H), 2.06 (m, 1H), 1.99 (m, 1H), 1.88 (m, 1H), 1.42 (m, 1H), 1.28 (m, 2H), −14.92 (ddt, J = 42.5, 13.5, 3.4 Hz, 1H). 31P{1H} NMR (CDCl3): δ 37.1 (dd, J = 71.7, 21.8 Hz, 1P), 33.8 (d, J = 71.7 Hz, 1P), −0.4 (d, J = 21.8 Hz, 1P). Reaction of 2 with 1,3-Propanedithiol. 2 (50 mg, 0.043 mmol) was dissolved in benzene (15 mL), and to the resultant solution was added excess 1,3-propanedithiol (ca. 10-fold). The mixture was boiled under reflux for 15 h. Chromatographic separation and workup similar to those for the above reaction developed two bands. The second band afforded Ru3(CO)3(μ-pdt)2{μ-P(C4H3S)2}(μ-dppm)(μ-H) (8; 10 mg, 20%) as orange crystals after recrystallization from hexane/CH2Cl2 at 4 °C. Data for 8 are as follows. Anal. Calcd for C42H41O3P3Ru3S6: C, 42.67; H, 3.50. Found: C, 42.93; H, 3.54. IR (ν(CO), CH2Cl2): 1940 s, 1913 m cm−1. 1H NMR (CDCl3): δ 7.56−6.81 (m, 26H), 3.60 (m, 1H), 2.94 (m, 1H), 2.71 (m, 1H), 2.62 (m, 1H), 2.53 (m, 1H), 2.17 (m, 4H), 2.02 (m, 2H), 1.46 (m, 3H), −14.99 (ddt, J = 40.5, 10.0, 1.1 Hz, 1H). 31P{1H} NMR (CDCl3): δ 33.6 (m, 2P), 13.7 (m, 1P). Reaction of 1 with 1,2-Ethanedithiol. To a benzene solution (20 mL) of 1 (100 mg, 0.089 mmol) was added excess 1,2ethanedithiol (ca. 10-fold), and the mixture was heated to reflux for 15 h. The solvent was removed under reduced pressure and the residue separated by TLC on silica gel. Elution with hexane/CH2Cl2 (1/1, v/ v) developed two bands. The second band afforded Ru3(CO)3(μedt)2{μ-P(C4H3O)2}(μ-dppm)(μ-H) (9; 50 mg, 50%) as orange crystals after recrystallization from hexane/CH2Cl2 at 4 °C. Data for 9 are as follows. Anal. Calcd for C40H37O5P3Ru3S4: C, 42.82; H, 3.30. Found: C, 43.17; H, 3.36. IR (ν(CO), CH2Cl2): 1943 s, 1920 m cm−1. 1 H NMR (CDCl3): δ 7.63 (m, 1H), 7.49 (m, 2H), 7.38 (m, 8H), 7.33 (m, 2H), 7.19 (m, 3H), 7.12 (m, 1H), 7.00 (m, 4H), 6.86 (m, 2H), 6.40 (m, 1H), 6.15 (m, 1H), 5.90 (m, 1H), 3.74 (m, 1H), 3.26 (m, 1H), 3.22 (m, 1H), 2.99 (m, 1H), 2.79 (m, 2H), 2.56 (m, 2H), 2.30 (m, 1H), 2.20 (m, 1H), −15.97 (ddt, J = 38.5, 11.0, 1.0 Hz, 1H). 31 1 P{ H} NMR (CDCl3): δ 37.5 (m, 2P), −8.9 (m, 1P). Reaction of 2 with 1,2-Ethanedithiol. To a benzene solution benzene (20 mL) of 2 (50 mg, 0.043 mmol) was added excess 1,2ethanedithiol (ca. 10-fold), and the mixture was heated to reflux for 15 h. The solvent was removed under reduced pressure and the residue chromatographed by TLC on silica gel. Elution with hexane/CH2Cl2 (3/2, v/v) developed two bands. The second band afforded Ru3(CO)3(μ-edt)2{μ-P(C4H3S)2}(μ-dppm)(μ-H) (10; 15 mg, 30%) as orange crystals after recrystallization from hexane/CH2Cl2 at 4 °C. Data for 10 are as follows. Anal. Calcd for C40H37O3P3Ru3S6: C, 41.63; H, 3.20. Found: C, 41.92; H, 3.26. IR (ν(CO), CH2Cl2): 1943 s, 1922 m cm−1. 1H NMR (CDCl3): δ 7.38 (m, 7H), 7.28 (m, 7H), 7.16 (m, 3H), 7.02 (m, 2H), 6.94 (m, 3H), 6.83 (m, 3H), 6.64 (m, 1H), 3.66 (m, 1H), 3.23 (m, 1H), 3.08 (m, 2H), 2.79 (m, 2H), 2.55 (m, 2H), 2.39 (m, 1H), 2.27 (m, 1H), −16.05 (m, 1H). 31P{1H} NMR (CDCl3): δ 36.0 (m, 2P), 4.9 (m, 1P). Reaction of 1 with PhSeH. A benzene solution (15 mL) of 1 (50 mg, 0.045 mmol) and PhSeH (ca. 5 fold) was boiled under reflux for 15 h. The solvent was removed under vacuum and the residue chromatographed by TLC on silica gel. Elution with hexane/CH2Cl2 (7/3, v/v) developed five bands. The first band was unreacted 1 (12 mg). The third and second bands afforded Ru3(CO)4(κ1-SePh)(μSePh)3{μ-P(C4H3O)2}(μ-dppm)(μ-H) (11; 14 mg, 20%) as green crystals and Ru3(CO)3(μ-SePh)4(μ-P(C4H3O)2}(μ-dppm)(μ-H) (12; 19 mg, 27%) as orange crystals after recrystallization from hexane/ CH2Cl2 at 4 °C. The contents of the other two bands were too small for characterization. Data for 11 are as follows. Anal. Calcd for C61H49O6P3Ru3Se4: C, 46.09; H, 3.11. Found: C, 46.43; H, 3.16. IR (ν(CO), CH2Cl2): 2036 s, 1979 m, 1947 m cm−1. 1H NMR (CDCl3): 2555

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

Table 1 empirical formula fw temp (K) wavelength (Å) cryst syst space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z Dcalcd (Mg m−3) μ(Mo Kα) (mm−1) F(000) cryst size (mm) θ range (deg) limiting indices

no. of rflns collected no. of indep rflns (Rint) max, min transmissn no. of data/restraints/params goodness of fit on F2 final R indices (I > 2σ(I)) R1 wR2 R indices (all data) R1 wR2 largest diff in peak and hole (e Å−3)

3

4

7

8

11

C102H82Cl4O14P6Ru6S4 2593.96 150(2) 0.71073 monoclinic P21/c 21.585(3) 21.581(3) 21.924(3) 90 93.659(2) 90 10 192(2) 4 1.690 1.210 5168 0.14 × 0.04 × 0.03 1.34−28.44 −28 ≤ h ≥ 28 −27 ≤ k ≥ 28 −29 ≤ l ≥ 28 87 844 24 236 (0.1287) 0.9646, 0.8498 24 236/0/1225 0.956

C51H41Cl2O5P3Ru3S4 1329.10 150(2) 0.71073 monoclinic, P21/c 20.046(3) 12.570(2) 21.113(4) 90 112.080(2) 90 4930(1) 4 1.791 1.331 2648 0.26 × 0.24 × 0.14 1.93−28.33 −26 ≤ h ≥ 25 −16 ≤ k ≥ 16 −28 ≤ l ≥ 27 38 641 11 455 (0.0473) 0.8355, 0.7234 11 455/0/585 1.114

C43H42Cl2O5P3Ru3S4 1234.03 123(2) 0.71073 monoclinic P21/c 16.510(1) 15.184(1) 18.735(1) 90 102.050(1) 90 4593.1(5) 4 1.785 1.421 2460 0.12 × 0.10 × 0.06 2.68−28.32 −21 ≤ h ≥ 21 −19 ≤ k ≥ 19 −24 ≤ l ≥ 24 38 721 10 695 (0.1541) 0.9196, 0.8480 10 695/0/541 0.936

C43H43Cl2O3P3Ru3S6 1267.15 150(2) 0.71073 monoclinic P21/c 16.842(2) 15.214(2) 18.792(2) 90 101.997(2) 90 4710(1) 4 1.787 1.471 2528 0.135× 0.15 × 0.05 2.26−28.31 −22 ≤ h ≥ 22 −19 ≤ k ≥ 19 −24 ≤ l ≥ 24 39 430 10 942 (0.0345) 0.9301, 0.6271 10 942/0/540 1.107

C150H130O12P6Ru6Se8 3548.46 150(2) 0.71073 orthorhombic P212121 14.635(3) 18.585(3) 50.286(9) 90 90 90 13 678(4) 4 1.723 2.906 7008 0.40 × 0.22 × 0.04 1.17−28.29 −19 ≤ h ≥ 19 −24 ≤ k ≥ 24 − 67 ≤ l ≥ 66 118 372 32 910 (0.0605) 0.8926, 0.3894 32 910/0/829 1.060

0.0780 0.1596

0.0727 0.2024

0.0632 0.0889

0.0275 0.0711

0.0503 0.1023

0.1661 0.1912 3.008 and −1.384

0.0961 0.2209 3.376 and −1.539

0.1398 0.1113 1.114 and −1.052

0.0344 0.0724 1.229 and −0.978

0.0741 0.1116 0.883 and −0.847

parameters were found; a = 14.648(3) Å, b = 20.880(5) Å, c = 19.271(6) Å, β = 110.137(25)°, V = 5533.60(2.49) Å3. A solution was found in P21/c, and all atoms were refined anisotropically. The best refinement was at ca. 10% and showed five to six large peaks in the difference map that could not be accounted for. Further details of the data collection and structure refinement for 3, 4, 7, 8, and 11 are given in Table 1. Computational Details. Calculations were performed on all derivatives with the Gaussian 03 (G03) program package42 employing the DFT method with the Becke three-parameter hybrid functional43 and Lee−Yang−Parr’s gradient-corrected correlation functional44 (B3LYP). The LanL2DZ basis set45 and effective core potential were used for the Ru, P, and S atoms, and the split-valence 3-21G** basis set46 was applied for all other atoms. The geometry of the clusters and the electronic structures were calculated in the gas phase. Normal mode analysis was employed to confirm the stationary point nature of the geometry calculated.



Road, Cambridge CB2 1FZ, U.K.; fax +44-1223-336033; e-mail [email protected]).



AUTHOR INFORMATION

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This research has been sponsored by the Ministry of Science and Information & Communication Technology, Government of the People’s Republic of Bangladesh. M.I.H. thanks the Bangladesh Academy of Sciences for a fellowship. We thank Dr. Michael Morris (University of Sheffield) for reading a draft of this work and passing on many helpful comments.



ASSOCIATED CONTENT

REFERENCES

(1) (a) Henkel, G.; Weissgraber, S. In Metal Clusters in Chemistry; Braunstein, P., Oro, L. A., Raithby, P. R., Eds.; Wiley-VCH: Weinheim, Germany, 1999; Vol. 1, p 163. (b) Ansari, M. A.; Ibers, J. A. Coord. Chem. Rev. 1990, 100, 223. (c) Roof, L. C.; Kolis, J. W. Chem. Rev. 1993, 93, 1037. (d) Kanatzidis, M. G.; Huang, S. P. Coord. Chem. Rev. 1994, 130, 223. (e) Mathur. Adv. Organomet. Chem. 1997, 41, 243. (f) Degroot, M. W.; Corrigan, J. F. Ed. McCleverty, J. A.; Meyer, T. J. Compr. Coord. Chem. II 2004, 7, 57. (g) Corrigan, J. F.; DeGroot, M. W.; Rao, C. N. R.; Mueller, A.; Cheetham, A. K. Chem. Nanomater.

S Supporting Information *

Text, tables and cif files giving details of the X-ray crystallographic structure determinations of 3, 4, 7, 8, and 11. This material is available free of charge via the Internet at http://pubs.acs.org. Crystal data are also available from http:// www.ccdc.cam.ac.uk/conts/retrieving.html as supplementary publications CCDC Nos. 864698−864702 (CCDC, 12 Union 2556

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

2004, 2, 418. (h) Hogarth, G.; Kabir, S. E.; Nordlander, E. Dalton Trans. 2010, 39, 6153. (2) (a) Hunter, C. A. Chem. Soc. Rev. 1994, 101. (b) Johnson, B. F. G.; Lewis, J.; Lodge, P. G.; Raithby, P. R.; Henrick, K; McPartlin, M. J. Chem. Soc., Chem. Commun. 1979, 719. (3) (a) Hogarth, G.; Taylor, N. J.; Carty, A. J.; Meyer, A. J. Chem. Soc., Chem. Commun. 1988, 834. (b) Humphrey, M. G.; LockhartGillett, B.; Samoc, M.; Skelton, B. W.; Tolhurst, V.-A.; White, A. H.; Wilson, A. J.; Yates, B. F. J. Organomet. Chem. 2005, 690, 1487. (c) Graiff, C.; Predieri, G.; Tiripicchio, A. Eur. J. Inorg. Chem. 2003, 1659. (d) Baistrocchi, P.; Careri, M.; Cauzzi, D.; Graiff, C.; Lanfranchi, M.; Manini, P.; Predieri, G.; Tiripicchio, A. Inorg. Chim. Acta 1996, 252, 367. (4) (a) Broadhurst, P. V.; Johnson, B. F. G.; Lewis, J. J. Chem. Soc., Dalton Trans. 1982, 1881. (b) Arce, A. J.; Arrojo, P.; Sanctis, Y.; De.; Deeming, A. J.; West, D. J. Polyhedron 1992, 11, 1013. (c) Arce, A. J.; Arrojo, P.; De Sanctis, Y.; Deeming, A. J. J. Chem. Soc, Chem. Commun. 1991, 1491. (d) Zhang, J.; Leong, W. K. Dalton Trans. 2000, 1279. (5) (a) Fomperyrine, P.; Lavigne, G.; Bonnet, J. J. J. Chem. Soc., Dalton Trans. 1987, 91. (b) Hodge, S. R.; Johnson, B. F. G.; Lewis., J.; Raithby, P. R. J. Chem. Soc., Dalton Trans. 1987, 931. (c) Bryan, E. G.; Johnson, B. F. G.; Lewis, J. J. Chem. Soc., Dalton Trans. 1977, 1328. (d) Kiel, G. Y.; Takats, J. Organometallics 1989, 8, 839. (e) Adams, R. D. Polyhedron 1985, 4, 2003. (f) Adams, R. D.; Horvath, I. T. Prog. Inorg. Chem. 1985, 33, 127. (g) Adams, R. D.; Horvath, I. T. Prog. Inorg. Synth. 1989, 26, 303. (6) Hieber, W.; Spacu, P Z. Anorg. Chem. 1937, 233, 353. (7) Kettle, S. F. A.; Orgel, L. E. J. Chem. Soc. A 1960, 3890. (8) Osbourne, A. G.; Stone, F. G. A. J. Chem. Soc., Chem. Commun. 1965, 361. (9) (a) Bryan, E. G.; Johnson, B. F. G.; Lewis, J. J. Chem. Soc., Dalton Trans. 1977, 1328. (b) Allen, V. F.; Mason, R.; Hitchcock, J. J. Organomet. Chem. 1977, 140, 297. (c) Leadbeater, N. E. J. Organomet. Chem. 1999, 573, 211. (d) Crooks, G. R.; Johnson, B. F. G.; Lewis, J.; Williams, I. G. J. Chem. Soc. A 1969, 797. (e) Evans, J.; Gracey, B. P. J. Chem. Res. Synop. 1986, 42. (f) Evans, J.; Gracey, B. P. J. Chem. Soc., Dalton Trans. 1982, 1123. (10) (a) Kabir, S. E.; Johns, C. A.; Malik, K. M. A.; Mottalib, M. A.; Rosenberg, E. J. Organomet. Chem. 2001, 625, 112. (b) Azam, K. A.; Kabir, S. E.; Miah, A.; Day, M. W. J. Organomet. Chem. 1992, 435, 157. (c) Fomperyrine, P.; Lavigne, G.; Bonnet, J. J. J. Chem. Soc., Dalton Trans. 1987, 91. (d) Hodge, S. R.; Johanson, B. F. G.; Lewis., J.; Raithby, P. R. J. Chem. Soc., Dalton Trans. 1987, 931. (e) Kiel, G. Y.; Takats, J. Organometallics 1989, 8, 839. (f) Ahmad, J. U.; Kabir, S. E.; Miah, M. A. J. Banglad. Chem. Soc. 1991, 4, 61. (g) Adams, R. D.; Babin, J. E.; Tasi, M. Inorg. Chem. 1986, 25, 4514. (11) (a) Johnson, B. F. G.; Johnston, R. D.; Josty, P. L.; Lewis, J.; Williams, I. G. Nature 1967, 213, 901. (b) Schermer, E. D.; Baddley, W. H. J. Organomet. Chem. 1971, 30, 67. (c) Chan, L.-F. S.; Low, K.H.; So, G. K.-M.; Chui, S. S.-Y.; Che, C.-M. Chem. Commun. 2011, 47, 8808. (12) (a) Begum, N.; Hyder, M. I.; Kabir, S. E.; Kundu, K. Indian J. Chem. 2005, 44A, 521. (b) Kabir, S. E.; Pervin, S.; Sarker, N. C.; Yesmin, A.; Sharmin, A.; Siddiquee, T. A.; Haworth, D. T.; Bennett, D. W.; Malik, K. M. A. J. Organomet. Chem. 2003, 681, 237. (c) Akter, T.; Begum, N.; Yesmin, A.; Haworth, D. T.; Bennett, D. W.; Kabir, S. E.; Miah, Md. A.; Sarker, N. C.; Siddiquee, T. A.; Rosenberg, E. J. Organomet. Chem. 2004, 689, 237. (d) Ahmed, S. J.; Hyder, Md. I.; Kabir, S. E.; Miah, Md. A.; Deeming, A. J.; Nordlander, E. J. Organomet. Chem. 2006, 691, 309. (e) Akter, H.; Deeming, A. J.; Hossain, G. M. G.; Kabir, S. E.; Mondol, D. N.; Nordlander, E.; Sharmin, A.; Tocher, D. A. J. Organomet. Chem. 2005, 690, 4628. (f) Hyder, Md. I.; Hossain, G. M. G.; Kabir, S. E.; Miak, Md. A. Polyhedron 2005, 24, 872. (g) Kabir, S. E.; Begum, N.; Hassan, Md. M.; Hyder, Md. I.; Nur, H.; Bennett, D. W.; Siddiquee, T. A.; Haworth, D. T.; Rosenberg, E. J. Organomet. Chem. 2004, 689, 1569. (h) Kabir, S. E.; Saha, M. S.; Tocher, D. A.; Hossain, G. M. G.; Rosenberg, E. J. Organomet. Chem. 2006, 691, 97. (i) Abedin, S. M. T.;

Azam, K. A.; Hursthouse, M. B.; Kabir, S. E.; Malik, K. M. A.; Mottalib, M. A.; Rosenberg, E. J. Cluster Sci. 2001, 12, 5. (13) Uddin, M. N.; Begum, N.; Hassan, M. R.; Hogarth, G.; Kabir, S. E.; Miah, M. A.; Nordlander, E.; Tocher, D. A. Dalton Trans. 2008, 6219. (14) Ghosh, S.; Hogarth, G.; Kabir, S. E.; Nordlander, E.; Salassa, L.; Tocher, D. A. J. Organomet. Chem. 2011, 696, 1982. (15) Hogarth, G.; O’Brien, M.; Tocher, D. A. J. Organomet. Chem. 2003, 672, 29. (16) (a) Ju, T. D.; Lang, R. F.; Roper, G. C.; Hoff, C. D. J. Am. Chem. Soc. 1996, 118, 5328. (b) Zhao, X.; Hsiao, Y.-M.; Lai, C.-H.; Reibenspies, J. H.; Darensbourg, M. Y. Inorg. Chem. 2002, 41, 699. (c) Jennings, M. C.; Puddephatt, R. J. Inorg. Chem. 1988, 27, 4280. (17) (a) Georgakaki, I. P.; Thomson, L. M.; Lyon, E. J.; Hall, M. B.; Darensbourg, M. Y. Coord. Chem. Rev. 2003, 238−239, 255. (b) Evans, D. J.; Pickett, C. J. Chem. Soc. Rev. 2003, 32, 268. (c) Rauchfuss, T. B. Inorg. Chem. 2004, 43, 14. (d) Sun, L.; Åkermark, B.; Ott, S. Coord. Chem. Rev. 2005, 249, 1653. (e) Liu, X.; Ibrahim, S. K.; Tard, C.; Pickett, C. J. Coord. Chem. Rev. 2005, 249, 1641. (f) Adam, F. I.; Hogarth, G.; Richards, I. J. Organomet. Chem. 2007, 692, 3957. (g) Ezzaher, S.; Capon, J.-F.; Gloaguen, F.; Petillon, F. Y.; Schollhammer, P.; Talarmin, J. Inorg. Chem. 2007, 46, 9863. (h) Adam, F. I.; Hogarth, G.; Richards, I.; Sanchez, B. E. Dalton Trans. 2007, 2485. (18) Ghosh, S.; Hogarth, G.; Holt, K. B.; Kabir, S. E.; Rahaman, A.; Unwin, D. G. Chem. Commun. 2011, 47, 11222. (19) (a) Hogarth, G.; Phillips, J. A.; Van Gastel, F.; Taylor, N. J.; Marder, T. B.; Carty, A. J. J. Chem. Soc., Chem.Commun. 1988, 1570. (b) Corrigan, J. F.; Dinardo, M.; Doherty, S.; Hogarth, G.; Sun, Y.; Taylor, N. J.; Carty, A. J. Organometallics 1994, 13, 3572. (20) Mealli, C.; Proserpio, D. M. J. Am. Chem. Soc. 1990, 112, 5484. (21) (a) Cabeza, J. A.; Matinez-Garcia, M. A.; Riera, V.; Ardura, D.; Garcia-Granda, S. Organometallics 1998, 17, 1471. (b) Hanif, K. M.; Kabir, S. E.; Mottalib, M. A.; Hursthouse, M. B.; Malik, K. M. A.; Rosenberg, E. Polyhedron 2000, 19, 1073. (c) Hossain, G. M. G.; Hyder, M. I.; Kabir, S. E.; Malik, K. M. A.; Miah, M. A.; Siddiquee, T. A. Polyhedron 2003, 22, 633. (22) Kabir, S. E.; Hogarth, G. Coord. Chem. Rev. 2009, 253, 1285. (23) (a) Adams, R. D.; Chen, L.; Yamamoto, J. H. Inorg. Chim. Acta 1995, 229, 47. (b) Adams, R. D.; Yamamoto, J. H. Organometallics 1995, 14, 3704. (c) Adams, R. D.; Yamamoto, J. H. J. Cluster Sci 1996, 7, 643. (24) Kabir, S. E.; Malik, K. M. A.; Rosenberg, E.; Siddiquee, T. A. Inorg. Chem. Commun. 2000, 3, 140. (25) Adams, R. D.; Horváth, I. T. Inorg. Chem. 1984, 23, 4718. (26) Begum, N.; Hassan, M.; Kabir, S. E. Indian J. Chem., Sect. A 2005, 44A, 490. (27) Schermer, E. D.; Baddley, W. H. J. Organomet. Chem. 1971, 30, 67. (28) Begum, N; Hyder, M. I.; Hassan, M. R.; Kabir, S. E.; Bennett, D. W.; Haworth, D. T.; Siddiquee, T. A.; Rokhsana, D.; Sharmin, A.; Rosenberg, E. Organometallics 2008, 27, 1550. (29) (a) Song, L.-C.; Yan, C.-G.; Hu, Q.-M.; Wang, R.-J.; Wang, H.G. Huaxue Xuebao 1995, 53, 402. (b) Song, L.-C.; Sun, Y.; Hu, Q.-M.; Lin, Y J. Organomet. Chem. 2003, 676, 80. (c) Song, L.-C.; Lu, G.-L.; Hu, Q.-M.; Fan, H.-T.; Chen, J.-B.; Sun, J.; Huang, X.-Y. J. Organomet. Chem. 2001, 627, 255. (30) Arce, A. J.; Arrojo, P.; Deeming, A. J.; De Sanctis, Y. J. Chem. Soc., Chem. Commun. 1991, 1491. (31) Adams, H.; Bailey, N. A.; Gay, S. R.; Gill, L. J.; Hamilton, T.; Morris, M. J. J. Chem. Soc., Dalton Trans. 1996, 2403. (32) Adams, Morris, M. J.; Mountford, P.; Patel, R.; Spey, S. E. J. Chem. Soc., Dalton Trans. 2001, 2601. (33) Adams, H.; Bailey, N. A.; Gay, S. R.; Hamilton, T.; Morris, M. J. J. Organomet. Chem. 1995, 493, C25. (34) Adams, H.; Biebricher, A.; Gay, S. R.; Hamilton, T.; McHugh, P. E.; Morris, M. J.; Mays, M. J. J. Chem. Soc., Dalton Trans. 2000, 2983. (35) (a) Schollhammer, P.; Pétillon, F. Y.; Poder-Guillou, S.; Talarmin, J.; Muir, K. W.; Yufit, D. S. J. Organomet. Chem. 1996, 513, 2557

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558

Organometallics

Article

181. (b) Pétillon, F. Y.; Schollhammer, P.; Talarmin, J.; Muir, K. W. Coord. Chem. Rev. 1998, 178−180, 203. (36) (a) Benson, I. B.; Killops, S. D.; Knox, S. A. R.; Welch, A. J. J. Chem. Soc., Chem. Commun. 1980, 1137. (b) Shaver, A.; Soo Lum, B.; Bird, P.; Livingstone, E.; Schweitzer, M. Inorg. Chem. 1990, 29, 1832. (37) (a) Anwar, M. K.; Hogarth, G.; Senturk, O. S.; Clegg, W.; Elsegood, M. R. J.; Doherty, S. J. Chem. Soc., Dalton Trans. 2001, 341. (b) Doherty, S.; Hogarth, G. Chem. Commun. 1998, 1815. (38) (a) Rivomanana, S.; Lavigne, G.; Lugan, N.; Bonnet, J.-J. Inorg. Chem. 1991, 30, 4110. (b) Rivomanana, S.; Mongin, C.; Lavigne, G. Inorg. Chem. 1996, 35, 1195. (39) (a) Adams, R. D.; Boswell, E. M.; Captain, B.; Patel, M. A. Inorg. Chem. 2007, 46, 533. (b) Adams, R. D.; Captain, B.; Fu, W.; Smith, M. D. Inorg. Chem. 2002, 41, 5593. (c) Adams, R. D.; Captain, B.; Fu, W.; Smith, M. D. Inorg. Chem. 2002, 41, 2302. (d) Adams, R. D.; Captain, B.; Smith, J. L. Jr.; Hall, M. B.; Beddie, C. L.; Webster, C. E. Inorg. Chem. 2004, 43, 7576. (e) Adams, R. D.; Captain, B.; Fu, W. Inorg. Chem. 2003, 42, 1328. (f) Adams, R. D.; Captain, B.; Smith, J. L Jr. Inorg. Chem. 2005, 44, 1413. (g) Adams, R. D.; Smith, J. L. Jr. Inorg. Chem. 2005, 44, 4276. (40) SMART and SAINT+ software for CCDC diffractometers, version 6.1; Bruker AXS, Madison, WI, 2000. (41) Sheldrick, G. M. SHELXTL PLUS, version 6.1; Bruker AXS, Madison, WI, 2000. (42) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A. Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Forestman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzales, C.; Pople, J. A. Gaussian 03. (revision D.01); Gaussian Inc., Wallingford, CT, 2004. (43) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (44) Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B: Condens. Matter 1988, 37, 785. (45) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270. (46) Binkley, J. S.; Pople, J. A.; Hehre, W. J. J. Am. Chem. Soc. 1980, 102, 939.

2558

dx.doi.org/10.1021/om201052u | Organometallics 2012, 31, 2546−2558