Anda di halaman 1dari 8

Inorganica Chimica Acta 359 (2006) 27982805 www.elsevier.

com/locate/ica

The synthesis and structural properties of [M(dippe)(g2-C4H4S)] complexes of Pd and Pt and comparison with their Ni analog
Tu in, Stephen S. Oster, Karlyn Skugrud, William D. Jones lay A. Ates
Department of Chemistry, University of Rochester, Rochester, NY 14627, USA Received 5 October 2005; accepted 17 October 2005 Available online 20 December 2005 Dedicated to Professor Brian James.

Abstract X-ray structural and NMR spectroscopic data for the ring-opened thiophene complexes [Pd(dippe)(T)] (2), and [Pt(dippe)(T)] (3) are now presented. The complex [Ni(dippe)(T)] (1), where T = (g2-C,S-C4H4S), was reported by our group, previously.The structural and bonding properties of complexes 2 and 3 were compared with those of complex 1. DFT calculations were carried out to rationalize their relative stabilities and structural properties. Compound 1 loses thiophene at ambient temperature in solution, while compound 2 decomposes rapidly in both acetone-d6 and THF-d8 with kobs = 7.15(9) 105 and 7.7(3) 105 s1, respectively, to give products that varied by solvent. Complex 3 does not lose thiophene at temperatures below 100 C. The DG0 values determined from DFT calculations are consistent with the observed stabilities of the complexes. The single crystal X-ray structures of all three complexes contain a disordered thienyl fragment in the asymmetric unit due to the interchange of the position of sulfur in the metal-inserted thiophenic ring. The thiophenic moiety is relatively at in 1, 2 and 3, which is attributed to the open ligand environment at the M(dippe) fragment. All three complexes possess square-planar geometry around the metal center and have bond-length alternation among the thiophenic carbons, which indicates double bond localization. The calculated bond lengths are in good agreement with experimental data. Molecular orbital (MO) and natural bonding orbital (NBO) analyses were carried out to rationalize the results. 2005 Elsevier B.V. All rights reserved.
Keywords: Hydrodesulfurization; X-ray crystal structure; DFT calculations; Nickel compounds; Palladium compounds; Platinum compounds; CS cleavage

1. Introduction Hydrodesulfurization (HDS) is the process used to remove sulfur from petroleum feedstocks. The most commonly used catalyst is CoMo, supported on alumina [1]. Other common combinations include NiMo and NiW. A detailed understanding of the reactivity of the catalyst is crucial to the optimization of the HDS process. Recently, the room temperature CS insertion of thiophenes by [Ni(dippe)(l-H)]2, which is a metal-hydride cluster of great utility used to model the nickel site in nickel-doped MoS2,

has been reported by our group [2,3]. The metallacycle complex formed from the reaction with thiophene is [Ni(dippe)(g2-C,S-C4H4S)] (1), or [Ni(dippe)(T)], which is stable indenitely in the solid state and under an inert atmosphere. Yet, in solution it exists in equilibrium with thiophenic-bridged dinuclear complexes and free thiophene, as illustrated in Eq. (1).
P 2 P Ni S

THF Keq = 0.36

P P Ni + thiophene S Ni

P P

1
P
*

Corresponding author. E-mail address: jones@chem.rochester.edu (W.D. Jones).

= dippe P

0020-1693/$ - see front matter 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.ica.2005.10.052

T.A. Ates in et al. / Inorganica Chimica Acta 359 (2006) 27982805

2799

Second and third row metals are known to be considerably more active than their rst row counterparts [4]. Platinum and palladium complexes have been shown to promote the activation and desulfurization of substituted thiophenes [58]. To study the eect of the d8 transition metals on the activation of thiophenic CS bonds and the electronic structure of the thiophene metallacycles formed, the structural analogs of 1 containing Pd and Pt were synthesized, fully characterized and the structural characteristics and bonding properties were compared with 1. 2. Results and discussion 2.1. Synthesis of [Pd(dippe)(T)] (2) The bridging hydrido complex, [Pd(dippe)(l-H)]2, was used as a precursor to 2 and was synthesized via the addition of LiHBEt3 to a THF slurry of [Pd(dippe)Cl2]. However, it is extremely dicult to isolate [Pd(dippe)(l-H)]2 because the bridging hydride complex decomposes when the solvent is removed to dryness under vacuum. The major decomposition product obtained in our attempts to isolate the product was [Pd2(l-dippe)2], which was rst reported by Fink [9]. Consequently, 2 was synthesized by generating [Pd(dippe)(l-H)]2 in situ in neat thiophene, as illustrated by Eq. (2). Complex 2 is stable indenitely in the solid state at 30 C. However, like 1, which loses thiophene at ambient temperature, it is unstable when it is in solution. It rapidly decomposes to form products that vary according to the particular solvent in use.
P Pd P Cl LiHBEt3 P Cl neat C4H4S P Pd 2 S

mer to the latter was approximately 1:2.5. Free thiophene signals were growing in intensity in the 1H NMR spectrum in the same manner as in the acetone-d6 experiment. 2.2. Synthesis of [Pt(dippe)(T)] (3) The platinum-hydride dimer [Pt(dippe)H]2 was synthesized and isolated according to a previously reported procedure [12]. Conversion of the hydride to 3 in neat thiophene, which is illustrated by Eq. (3), occurs over the course of 10 d at 70 C or over 1 h at 160 C. The dinuclear complex [[Pt(dippe)]2(l-g2,g2-COD)] was also used as a precursor to 3, as illustrated by Eq. (4), and takes place under similar conditions and over the same amount of time. Unlike 1 and 2, complex 3 is thermally stable at temperatures below 100 C and does not lose thiophene at room temperature in solution.
P Pt P H H Pt P P neat C4H4S 70 C P Pt P 3
Pt P P neat C4H4S 2 60 C P 3 P Pt S

P P

Pt

4 2.3. Comparison of X-ray structures Single crystals of 2 and 3 were obtained, and the X-ray structures of these compounds were determined for comparison with the structure of 1. The crystals of all three metallacycles, whose ORTEP drawings are shown in Figs. 13, contained a disordered molecule in the

The kinetics of the degradation of 2 in solution was determined for both acetone and THF using 31P NMR spectroscopy. The observed rst order rate constants are 7.15 (9) 105 and 7.7 (3) 105 s1, respectively. As stated above, the degradation products were dierent in each solvent. 31P NMR spectra of 2 in acetone-d6 that were taken over the course of 4 h showed the metallacycle disappearing as two new unidentied products appeared. These new complexes gave rise to a doublet of doublets at d 60 and a singlet at d 82. Concurrently, free thiophene signals increased in intensity in the 1H NMR spectrum throughout the experiment. In THF-d8, 2 converted to four major phosphorus-containing products that were seen in the 31P NMR spectrum. Two of these are unidentied complexes, corresponding to a singlet at d 79.5 and two doublets at d 70.1 and d 80.1 that increased in concentration during the rst 5 h and then gradually disappeared. The other two products were identied as [Pd(dippe)2] and [Pd2(dippe)2(l-dippe)], which produced a singlet at d 45.5 and two multiplets at d 49.5 and d 54.8, respectively [10,11]. The concentrations of both complexes continued to increase over a period of 17 h, and the nal ratio of the for-

Fig. 1. ORTEP drawing of [Ni(dippe)-T] (1). Ellipsoids are shown at the 30% probability level. All hydrogen atoms are omitted for clarity. Selected ): NiC(1), 1.967; C(1)C(2), 1.352; C(2)C(3), 1.404; bond lengths (A C(3)C(4),1.340; C(4)S, 1.709; SNi, 2.160; NiP(1), 2.189; NiP(2), 2.215. Selected angles (): PNiP, 87.72; P(1)NiC(1), 91.50; P(1)NiS, 170.23; P(2)NiC(1),175.45; P(2)NiS, 86.25; SNiC(1), 95.10.

2800

T.A. Ates in et al. / Inorganica Chimica Acta 359 (2006) 27982805 Table 1 ) Dihedral bond angles () and mean deviation from the thiophene plane (A for [Ni(dippe)-T], [Pd(dippe)-T], and [Pt(dippe)-T] [Ni(dippe)-T] PMP/C1MS C(1)MS/C(1)C(2) C(3)C(4) Mean deviation from plane 10.3 10.0 0.0451 [Pd(dippe)-T] 7.0 3.4 0.0234 [Pt(dippe)-T] 8.2 3.0 0.0199

Fig. 2. ORTEP drawing of [Pd(dippe)-T] (2). Ellipsoids are shown at the 30% probability level. All hydrogen atoms are omitted for clarity. Disorder isomer ratio (fraction of major): 0.772(5). Selected bond lengths ): PdC(1), 2.047; C(1)C(2), 1.317; C(2)C(3), 1.434; C(3)C(4), 1.345; (A C(4)S, 1.723; SPd, 2.277; PdP(1), 2.2898; PdP(2), 2.3204. Selected angles (): PPdP, 86.19; P(1)PdC(1), 90.90; P(1)PdS, 174.79; P(2) PdC(1), 171.5; P(2)PdS, 90.65; SPdC(1), 92.80.

asymmetric unit as a result of the interchange of the position of the sulfur atom in the metal-inserted thiophenic ring. This type of disorder also appeared in the crystals of related rhodiumthiophene and rhodiumselenophene complexes [13,14]. For these complexes, each of their orientations was successfully rened in SHELX by constraining the corresponding bond distances, using the SAME instruction, which rendered the rings to be similar to each other. Likewise, we used this technique in rening the crystal structures of 1, 2, and 3.

Complex 2 diered from 1 and 3 in that it crystallized in the P21/n space group while the other two complexes crystallized in the R3c space group. The thiophenic moiety is relatively at in all three complexes, as is indicated in Table 1. For instance, the dihedral angle between the C(1)MS and C(1)C(2)C(3)C(4) planes and the mean deviation from the plane are at a maximum for 1 and at a minimum of for 3. The planarity of the ring in metallathiophenes, in the past, has been invoked as evidence of electron delocalization to form a metallathiabenzene. However, a joint study of [Cp*Rh(PMe3)(g2-C,S-2,5-Me2C4H2S)] and other closely related thiophene metallacycles by our group and that of Harris suggests that ring planarity is a steric rather than an electronic eect [13]. In this case, the orientation of the ring was directed into a pocket that minimized interaction of an a-methyl group with the other ligands, but also created a dihedral angle of 26 between the SRhC1 plane and the plane of the thiophenic carbons. This compares to a bend angle of 4.1 in the parent molecule [Cp*Rh(PMe3)(g2-C,S-C4H4S)] [15]. The spatial demands of the S and C(1) atoms not withstanding, the planarity of the metallacycle ring in 1, 2, and 3 is to be expected, as the ligand environment above and below the thiophenic

Fig. 3. ORTEP drawing of [Pt(dippe)-T] (3). Ellipsoids are shown at the 30% probability level. All hydrogen atoms are omitted for clarity. Disorder ): PtC(1), 1.930; C(1)C(2), 1.405; C(2)C(3), 1.425; C(3)C(4), 1.345; C(4)S, 1.712; isomer ratio (fraction of major): 0.642(12). Selected bond lengths (A SPt, 2.322; PtP(1), 2.277; PtP(2), 2.29. Selected angles (): PPtP, 86.61; P(1)PtC(1), 91.80; P(1)PtS, 172.51; P(2)PtC(1), 174.45; P(2)PtS, 87.52; SPtC(1), 94.35.

T.A. Ates in et al. / Inorganica Chimica Acta 359 (2006) 27982805

2801

moiety is the same and there are no steric interactions within the square plane. In another relevant case, Angelici found that the ring-opened thiophene ligand in Cp*Ir(g2C,S-2,5-Me2C4H2S) is planar whereas the ring is bent to 32 in Cp*Ir(g2-C,S-2,5-Me2C4H2S)(PMe2Ph) [16]. While structure determination for all three metallacycles was hampered by the disorder that resulted from the orientation of the opened thiophene, the data for 1 and 2 that is shown in Table 2 demonstrate that in both cases there is clear bond-length alternation within the ring of the type that is associated with double bond localization. The C(1)C(2) and C(3)C(4) bond lengths approximate the average distance between sp2 carbons that are separated by a double bond while the C(2) C(3) lengths corresponds to sp2sp2 carbon single bonds [17]. The structural data for 3, shown in Table 2, are different. First, while there is clear alternation in the C(2) C(3) and C(3)C(4) intercarbon distances in 3 just as in its nickel and palladium analogs, there is only a slight alternation in the C(1)C(2) and C(2)C(3) intercarbon distances; the two bonds are very similar in length, and with values that are intermediate between sp2sp2 carbon single and double bonds. Second, the PtC(1) bond is much shorter, and the SC(4)bond is slightly shorter, than the same bonds in the other two complexes. The observed disordered structure for 3 persuaded us to perform density functional theory calculations before reaching any conclusions. 2.4. Computational data To simplify the calculations, the i-Pr groups of 1, 2, and 3 were substituted by methyl groups. This simplication is assumed to have no steric outcome on the calculations, as the X-ray single crystal structures showed no interaction between the methyl groups and the thiophene ring. The

optimized structures of [Ni(dmpe)(T)] (1 0 ), [Pd(dmpe)(T)] (2 0 ) and [Pt(dmpe)(T)] (3 0 ), also show square-planar coordination at the metal centers. Table 2 shows selected calculated bond lengths and angles for 1 0 , 2 0 , and 3 0 , which are in good agreement with the solid state structures of 1, 2, and 3. The data demonstrate that all three have several common structural characteristics such as square-planar geometry around the metal center, coplanar thienyl ligand and clear bond-length alternation (shortlongshort) within the thiophenic ring. The C(1)C(2) and C(3)C(4) bond lengths are within the range of a CC double bond and the C(2)C(3) bond lengths are within the range of a C C single bond. Although the thiophenic carboncarbon bond lengths of 3 obtained from the solid state structure have a longer C(1)C(2) bond implying electron delocalization, the optimized structure shows localized double bonds for 1 0 , 2 0 , and 3 0 . This result is consistent with the bond lengths reported for the substituted thiaplatinacycles in the literature [57]. The dierences in energies (DE and DG0) for the formation of [M(dmpe)(T)] from the M(dmpe) and thiophene fragments shown in Table 3 are in good agreement with the relative stability observed for the complexes 1, 2, and 3. The energy required to remove the thiophene fragment

Table 3 The dierences of energies, enthalpies and entropies for the formation of [M(dmpe)(T)] from the M(dmpe) and thiophene fragments at 298 C, 1 atm Complex 10 20 30 DE (kcal/mol) 35.95 20.89 43.57 DH (kcal/mol) 34.50 19.46 41.95 DG (kcal/mol) 22.49 6.96 29.28 DS (cal/mol/K) 40.3 41.9 42.5

Table 2 ) and angles () for [M(dippe)(T)] (experimental) and [M(dmpe)(T)] (calculated) Selected bond lengths (A Ni Expt. Bond lengths MC(1) C(1)C(2) C(2)C(3) C(3)C(4) C(4)S SM MP(1) MP(2) Bond angles PMP P(1)MC(1) P(1)MS P(2)MC(1) P(2)MS SMC(1) 1.967 1.352 1.404 1.340 1.709 2.160 2.189 2.215 87.72 91.50 170.23 175.45 86.25 95.10 (2) (12) (2) (3) (9) (8) (2) (2) Calc. 1.892 1.363 1.438 1.356 1.736 2.181 2.207 2.278 86.556 89.730 173.895 175.943 87.704 96.066 Pd Expt. 2.047 (2) 1.317 (2) 1.434 (9) 1.345 (9) 1.723 (2) 2.277 (2) 2.2898 (10) 2.3204 (10) 86.19 90.90 174.79 171.50 90.65 92.80 (4) (6) (14) (11) (7) (5) Calc. 2.017 1.358 1.442 1.357 1.736 2.326 2.320 2.398 85.150 91.072 174.827 176.126 89.734 94.046 Pt Expt. 1.93 (3) 1.405 (3) 1.425 (3) 1.345 (3) 1.712 (2) 2.322 (3) 2.277 (2) 2.290 (3) 86.61 91.80 172.51 174.45 87.52 94.35 (9) (10) (15) (10) (14) (9) Calc. 2.025 1.363 1.440 1.358 1.736 2.338 2.296 2.376 85.626 90.988 174.956 176.610 89.355 94.030

(5) (3) (6) (5) (2) (4)

2802

T.A. Ates in et al. / Inorganica Chimica Acta 359 (2006) 27982805

from the metal complexes is greatest for 3 and smallest for 2. These dierences in reactivities among three similar complexes prompted us to examine the electronic structures of these complexes and attempt to relate the electronic structures to the known stabilities of the complexes. Bonding in these complexes was considered in terms of the interactions between a metal fragment and an opened thiophene fragment [18]. These are shown on the left and right hand sides of Fig. 4, respectively. While only the MO interaction diagram for 1 0 is discussed in the text, those for 2 0 and 3 0 are essentially the same as for 1 0 . The major dierences are emphasized in Fig. 5. The electron density of the ve highest occupied molecular orbitals and the two lowest unoccupied molecular orbitals of the Ni(dmpe) fragment is localized mainly on the metal atom. The HOMO (highest occupied molecular orbital) is largely the dx2 y 2 orbital of the nickel atom and the SHOMO (second highest occupied molecular orbital) is largely the dz2 orbital. The third, fourth and the fth highest energy occupied molecular orbitals are all some combination of the remaining d orbitals of the metal atom (dxy, dxz and dyz). The LUMO (lowest unoccupied molecular orbital) is mainly Ni sp hybrid orbital directed toward

-2

Energy (eV)

-4

-6

-8

-10

Ni(dmpe)

Pd(dmpe)

Pt(dmpe)

Fig. 5. The energies of the molecular orbitals for the metal fragments. The highest occupied molecular orbitals are marked with an electron pair.

Fig. 4. The molecular orbital interaction diagram for 1. Bonding is considered in terms of the interactions between a metal fragment and an opened thiophene fragment, which are shown on the left and right hand sides, respectively. The highest occupied molecular orbitals are marked with an electron pair.

the incoming thiophene ligand and it is in the plane of the dmpe ring system. The SLUMO (second lowest unoccupied molecular orbital) has a p orbital symmetry and it is perpendicular to the ligand ring system. The molecular orbitals calculated for the opened thiophene are very similar to the ones reported by Harris for the opened 2,5-dimethylthiophene [18]. The HOMO in the opened thiophene fragment is a p orbital perpendicular to the plane of the ring system. The electron density of this orbital is mainly localized on the sulfur atom with signicant p electron density on C(1). The SHOMO and the LUMO in the ring-opened thiophene are both r orbitals containing electron density on both the sulfur and C(1). The SHOMO is bonding with respect to the CS interaction while the LUMO is antibonding. The third highest occupied and the second lowest unoccupied molecular orbitals are both p orbitals perpendicular to the ring system. The bonding interactions between the metal and the thiophene fragments result from the donation of electron density from the HOMO of the Ni(dmpe) fragment into the LUMO of the opened thiophene fragment, back donation from the SHOMO and HOMO of the thiophene fragment into the LUMO and SLUMO of the metal fragment orbitals, respectively. The relative stabilities of 1, 2 and 3 can be explained in terms of these two important bonding interactions. The energies of the molecular orbitals for the three metal fragments of the model compounds 1 0 , 2 0 and 3 0 , are shown in Fig. 5. Among these three fragments, the

T.A. Ates in et al. / Inorganica Chimica Acta 359 (2006) 27982805

2803

HOMO of the Ni(dmpe) fragment has the highest energy. When the metal fragment donates electron density from this orbital to the thiophene fragment, 1 is expected to be stabilized the most based on this bonding interaction. However, the back donation from the thiophene fragment to the metal fragment is also as important as the donation from the metal to the thiophene ligand. When the metal fragment accepts electron density from the SHOMO of the thiophene fragment, compound 3 is more stabilized than 1 and 2, as the Pt(dmpe) fragment has the lowest energy LUMO among all three fragments. The combination of these two bonding interactions explains the lowest stability of 2, since it is less stabilized than 1 in terms of the rst bonding interaction mentioned above and more destabilized than 3 in terms of the second one. Based on the fact that 3 is more stable than 1, back donation from the thiophene ligand to the metal atom is the predominant bonding interaction in the formation of the thiametallacycles. The donation and back donation interactions between the metal fragment and the thiophene fragment can easily be seen from the natural bonding orbital (NBO) analysis, which provides the natural atomic orbital occupancies and natural charges of the atoms. The occupancy of the metal valence s orbital, which is the LUMO of the metal fragment, indicates the amount of back donation of electron density to the metal, while the occupancy of the metal valence dx2 y 2 orbital, which is the HOMO of the fragment, indicates the amount of donation of electron density from the metal fragment to the thiophene fragment. The natural atomic orbital occupancies for the valence dx2 y 2 orbital of Ni (1.31), Pd (1.46) and Pt (1.44) are consistent with the above interpretation that Ni is a better donor than both Pd and Pt. Similarly, the natural atomic orbital occupancies for the valence s orbitals of Ni (0.46), Pd (0.5) and Pt (0.66) show that the best acceptor is Pt. The natural population analyses indicate more positive charge on Ni (0.367) than Pd (0.214) and Pt (0.149), which is also consistent with the above discussion. 3. Conclusion In summary, complex 2 rapidly decomposes when it is in both THF and acetone to give products that vary by solvent, while 3 is stable in solution to elevated temperatures. The X-ray crystal structures of 1, 2, and 3 show disorder as a result of the interchange of the position of the sulfur atom in the metal-inserted rings. Complexes 1, 2 and 3 have localized bonding in the thiophenic moiety in their solid state structure. DFT calculations in the gas phase show bond-length alternation among the thiophenic carbons for all three complexes, which indicates double bond localization. All three have square-planar geometry around the metal center and coplanar thienyl ligand. The calculated DG0 values are consistent with the observed stabilities of the complexes.

4. Experimental 4.1. General considerations All operations were performed under a nitrogen atmosphere unless stated otherwise. THF and hexanes were distilled from dark purple solutions of benzophenone ketyl. Thiophene (99+%) was purchased from Aldrich Chemical Co. and puried according to a literature procedure [19]. A Siemens-SMART 3-Circle CCD diractometer was used for X-ray crystal structure determination. Elemental analyses were obtained from Desert Analytics. All 1H, 13C{1H}, and 31P{1H} NMR spectra were recorded on a Bruker Avance 400 NMR spectrometer. All 1H and 13C chemical shifts are reported relative to the solvent. All 31P chemical shifts are reported relative to the signal of external 85% H3PO4. The synthesis, analytical, and structural data for compound 1 have been published previously [3]. 4.2. Theoretical details The known experimental structures for the complexes were used as the starting point for the calculations. To simplify the calculations, the i-Pr groups were substituted by methyl groups. This simplication is assumed to have no steric outcome on the calculations. 1 0 , 2 0 and 3 0 were fully optimized in redundant internal coordinates [20], with density-functional theory (DFT) and a wave function incorporating Beckes three-parameter hybrid functional (B3) [21], along with the LeeYangParr correlation functional (LYP) [22]. All calculations were performed using the GAUSSIAN98 [23] and GAUSSIAN03 [24] packages. The metal, P and S atoms were represented with the eective core pseudopotentials of the Stuttgart group and the associated basis sets improved with a set of f-polarization functions for the transition metals (a = 3.130, Ni; a = 1.472, Pd; a = 0.993, Pt) [25], and a set of d-polarization functions for the main group elements (a = 0.387, P; a = 0.503, S) [26]. The remaining atoms (C and H) were represented with 6-31G(d, p) [27] basis sets. The geometry optimizations were performed without any symmetry constraints and the local minima were checked by frequency calculations. The energies discussed throughout the text are electronic energies without any ZPE corrections. Gibbs free energies have been calculated at 298.15 K and 1 atm. The description of bonds between atoms as a linear combination of hybrids located on each partner of the bond was obtained from natural bonding orbital (NBO) calculations [28]. The Molden package was used to display the molecular orbitals and the electron densities [29]. 4.3. Synthesis of [Pd(dippe)-T] (2) LiHBEt3 (1.3 mL of 1 M solution in THF, 1.31 mmol) was added dropwise to a stirred suspension of [Pd(dippe)Cl2] (275 mg, 0.62 mmol) in approximately 25 mL of

2804

T.A. Ates in et al. / Inorganica Chimica Acta 359 (2006) 27982805

thiophene to immediately give a brown-red solution and a beige precipitate. After stirring for approximately 1.5 h, the solution was vacuum ltered through 80200 mesh neutral alumina on a ne glass frit. The solvent was evaporated under vacuum to yield a light red solid. The solid was washed three times with cold hexanes to yield 183 mg (64.6%) of a light yellow solid, which was then dried overnight under vacuum. 1H NMR (400 MHz, THF-d8, 25 C): d 7.56 (m, J = 10.6, 10.0, 11.8 Hz, 1H), 6.87 (dm, J = 6.2, 5.0, 3.7 Hz, 1H), 6.74 (dd, J = 9.4, 8.1 Hz, 1H), 6.44 (dd, J = 7.5, 1.6 Hz, 1H), 2.47 (m, 2H), 2.28 (m, 2H), 1.55 (d, J(HP) = 10.4 Hz, 4H), 1.301.11 (m, 24H). 13C{1H} NMR (100 MHz, THF-d8, 35 C): d 141.01 (s), 122.93 (s), 121.07 (s), 117.84 (s), 23.85 (m), 20.99 (m), 19.51 (d, J(CP) = 3.6 Hz), 19.26 (d, J(CP) = 5.5 Hz), 18.54 (s), 18.29 (s). The signals of two isopropyl primary carbons are obscured. 31P{1H} NMR (162 MHz, THF-d8, 25 C): d 72.43 (d, J = 26.0 Hz). 65.21 (d, J = 26.0 Hz). Anal. Calc. for C18H36PdP2S: C, 47.70; H, 8.01. Found: C, 47.44; H, 8.19%. 4.4. Synthesis of [Pt(dippe)-T] (3) [Pt(dippe)]2(l-g2,g2-COD)] (2.6 mg, 0.0025 mmol) was placed in a resealable NMR tube that contained neat thiophene. The contents of the tube were heated at 55 C for two weeks and the thiophene and COD were evaporated under vacuum to yield 2.2 mg of a yellow solid (80.4%). 1 H NMR (400 MHz, THF-d8, 25 C): d 8.19 (m, J = 9.4, 8.7, 2.3 Hz, 1H), 7.15 (m, 1H), 7.07 (m, 1H), 6.67 (m, 1H), 2.53 (m, 2H), 2.38 (m, 2H), 1.98 (m, 2H), 1.80 (m, 2H), 1.271.09 (m, 24H). 13C{1H} NMR (100 MHz, THF-d8, 25 C): d 131.07 (d, J = 8.7 Hz), 121.47 (J(CPt) = 67 Hz), 118.19 (s), 115.82 (s, J(CPt) = 17.6 Hz), 19.27 (d, J(CPt) = 24.0 Hz), 19.01 (d, J(CPt) = 24.8 Hz), 16.39 (d, J(CP) = 2.8 Hz, J(CPt) = 8.2 Hz), 16.13 (d, J(CP) = 3.7 Hz), 15.58 (J(CPt) = 6.6 Hz), 15.24 (J(CPt) = 11.0 Hz). The signals of two isopropyl primary carbons are obscured. 31P{1H} NMR (162 MHz, THF-d8, 25 C): d 66.39 (d, J(PP) = 4.9 Hz, J(PPt) = 842.0 Hz), 65.65 (d, J(PP) = 4.9 Hz, J(PPt) = 1493.9 Hz). Anal. Calc. for C18H36PtP2S: C, 39.51; H, 6.57. Found: C, 39.92; H, 6.70%. 4.5. Alternative synthesis of [Pt(dippe)-T] (3) LiHBEt3 (1.0 mL of 1 M in THF, 1.0 mmol) was added dropwise to a stirred suspension of [Pt(dippe)Cl2] (235 mg, 0.5 mmol) in approximately 25 mL of neat thiophene to immediately give a yellow solution and a white precipitate. After stirring for approximately 1.0 h at 160 C, the solution was vacuum ltered through 80200 mesh neutral alumina on a ne glass frit. The solvent was evaporated under vacuum to yield a yellow solid. The solid was washed three times with cold hexanes to yield 193 mg (85.7%) of a light yellow solid, which was then dried overnight under vacuum.

Acknowledgements The US Department of Energy, Grant FG02-86ER5369, and the National Science Foundation, Grant CHE0414325, are acknowledged for support of this work. Appendix A. Supplementary material Tables of crystallographic data, atomic coordinates, bond distances and angles, and anisotropic thermal parameters for 2 (CCDC #285741) and 3 (CCDC #285742), and input and optimized geometry les for the DFT calculations together with the important molecular orbitals and energies are available. Supplementary data associated with this article can be found, in the online version, at doi:10.1016/j.ica.2005.10.052. References
[1] R.A. Sanchez-Delgado, Organometallic Modelling of the Hydrodesulfurization and Hydronitrogenation Reactions, Kluwer Academic Publishers, Dordrecht, 2002, p. 5. [2] D.A. Vicic, W.D. Jones, J. Am. Chem. Soc. 121 (1999) 7606. [3] D.A. Vicic, W.D. Jones, J. Am. Chem. Soc. 119 (1997) 10855. [4] M.J. Ledoux, O. Michaud, G. Agostini, P. Panissod, J. Catal. 102 (1986) 275. [5] J.J. Garcia, A. Arevalo, S. Capella, A. Chehata, M. Hernandez, V. Montiel, G. Picazo, F. Del Rio, R.A. Toscano, H. Adams, P.M. Maitlis, Polyhedron 16 (1997) 3185. [6] J.J. Garcia, A. Arevalo, V. Montiel, F. Del Rio, B. Quiroz, H. Adams, P.M. Maitlis, Organometallics 16 (1997) 3216. [7] J.J. Garcia, B.E. Mann, H. Adams, N.A. Bailey, P.M. Maitlis, J. Am. Chem. Soc. 117 (1995) 2179. [8] M. Hernandez, G. Miralrio, A. Arevalo, S. Bernes, J.J. Garcia, C. Lopez, P.M. Maitlis, F. del Rio, Organometallics 20 (2001) 4061. [9] S.M. Reid, M.J. Fink, Organometallics 20 (2001) 2959. [10] M. Portnoy, D. Milstein, Organometallics 12 (1993) 1655. [11] M.D. Fryzuk, G.K.B. Clentsmith, S.J. Rettig, Organometallics 15 (1996) 2083. [12] D.J. Schwartz, R.A. Andersen, J. Am. Chem. Soc. 117 (1995) 4014. [13] C. Blonski, A.W. Myers, M. Palmer, S. Harris, W.D. Jones, Organometallics 16 (1997) 3819. [14] D.A. Vicic, A.W. Myers, W.D. Jones, Organometallics 16 (1997) 2751. [15] W.D. Jones, D.A. Vicic, M.C. Chin, J.H. Roache, A.W. Myers, Polyhedron 16 (1997) 3115. [16] J. Chen, L.M. Daniels, R.J. Angelici, Polyhedron 9 (1990) 1883. [17] J.S. Dewar, W. Thiel, J. Am. Chem. Soc. 99 (1977) 4907. [18] M. Palmer, K. Carter, S. Harris, Organometallics 16 (1997) 2448. [19] G.H. Spies, R.J. Angelici, Organometallics 6 (1987) 1897. [20] C. Peng, P.Y. Ayala, H.B. Schlegel, M.J. Frisch, J. Comput. Chem. 17 (1996) 49. [21] A.D. Becke, J. Chem. Phys. 98 (1993) 5648. [22] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785. [23] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, V.G. Zakrzewski, J.A. Montgomery, R.E. Stratmann, J.C. Burant, S. Dapprich, J.M. Millam, A.D. Daniels, K.N. Kudin, M.C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Cliord, J. Ochterski, G.A. Petersson, P.Y. Ayala, Q. Cui, K. Morokuma, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman, J. Cioslowski, J.V. Ortiz, A.G. Baboul, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, G. Gomperts, R.L. Martin, D.J. Fox, T. Keith, M.A.

T.A. Ates in et al. / Inorganica Chimica Acta 359 (2006) 27982805 Al-Laham, C.Y. Peng, A. Nanayakkara, C. Gonzalez, M. Challacombe, P.M.W. Gill, B.G. Johnson, W. Chen, M.W. Wong, J.L. Andres, M. Head-Gordon, E.S. Replogle, J.A. Pople, GAUSSIAN 98, Gaussian Inc., Pittsburgh, PA, 1998. [24] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, J.A. Montgomery Jr., T. Vreven, K.N. Kudin, J.C. Burant, J.M. Millam, S.S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J.E. Knox, H.P. Hratchian, J.B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, P.Y. Ayala, K. Morokuma, G.A. Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S. Dapprich, A.D. Daniels, M.C. Strain, O. Farkas, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman, J.V. Ortiz, Q. Cui, A.G. Baboul, S.

2805

[25]

[26]

[27] [28] [29]

Cliord, J. Cioslowski, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R.L. Martin, D.J. Fox, T. Keith, M.A. AlLaham, C.Y. Peng, A. Nanayakkara, M. Challacombe, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, C. Gonzalez, J.A. Pople, GAUSSIAN 03, Gaussian, Inc., Wallingford, CT, 2004. A.W. Ehlers, M. Bohme, S. Dapprich, A. Gobbi, A. Hollwarth, V. Jonas, K.F. Kohler, R. Stegmann, A. Veldkamp, G. Frenking, Chem. Phys. Lett. 208 (1993) 111. A. Hollwarth, M. Bohme, S. Dapprich, A.W. Ehlers, A. Gobbi, V. Jonas, K.F. Kohler, R. Stegmann, A. Veldkamp, G. Frenking, Chem. Phys. Lett. 208 (1993) 237. W.J. Hehre, R. Ditcheld, J.A. Pople, J. Chem. Phys. 56 (1972) 2257. A.E. Reed, L.A. Curtiss, F. Weinhold, Chem. Rev. 88 (1988) 899. G. Schaftenaar, J.H. Noordik, J. Comput. Aided Mol. Des. 14 (2000) 123.

Anda mungkin juga menyukai