(DFT) Calculations of Structure and 119Sn ... - ACS Publications

Oct 23, 2015 - School of Chemical Sciences, Devi Ahilya University Indore, Khandwa Road Campus, Indore 452001, India. •S Supporting Information...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/IC

Dispersion-Corrected Relativistic Density Functional Theory (DFT) Calculations of Structure and 119Sn Mö ssbauer Parameters for MSnR Bonding in Filippou’s Stannylidyne Complexes of Molybdenum and Tungsten Krishna K. Pandey* School of Chemical Sciences, Devi Ahilya University Indore, Khandwa Road Campus, Indore 452001, India S Supporting Information *

ABSTRACT: 119Sn Mössbauer isomer shift (IS) and quadrupole splitting (ΔEQ) for MSnR bonding in metal−stannylidyne complexes trans-[Cl(PMe3)4MoSn−R] (1), trans-[Cl(PMe3)4WSn−R] (2), trans-[Cl(dppe)2MoSn−R] (3), trans[Cl(dppe)2WSn−R] (4), [(dppe)2MoSn−R]+ (5), [(dppe)2WSn−R]+ (6) (R = C6H3−2,6-Mes2) have been investigated for the first time. Calculations of optimized structures and 119Sn Mössbauer parameters were carried out at the DFT/ TPSS-D3(BJ)/TZVPP/ZORA level of theory. The calculated geometry parameters of stannylidyne complexes of molybdenum and tungsten (1−6) are in good agreement with experimental values of W−Sn and Sn−C bond distances. The calculated values of the isomer shift for the complexes (1−6) are almost same to the experimental values (within ±0.1 mm/s). Experimental values (ISexptl, 2.38−2.50 mm/s) and calculated values (IScalcd, 2.37−2.49 mm/s) of isomer shifts indicate that the oxidation state of tin in the studied complexes with MSn−R bonding is Sn(II). The variations of ISexptl, as a function of Sn s electrons (Ns(Sn)), also exhibit a linear trend. (IS = 0.477Ns(Sn) − 1.888, R2 = 0.9973). Calculated values of isomer shift (IScalcd) using the linear regression with the Ns(Sn) electron density are in excellent concord with the ISexptl.The calculated values of nuclear quadrupole splitting parameters (ΔEQ(calcd)) of 119Sn using the relation ΔEQ(calcd) = (0.540 + 0.28) V are in agreement with the experimental values.



six more stannylidyne complexes.15−17 Structurally characterized metal−stannylidyne complexes are presented in Chart 1. The 119Sn Mössbauer spectral studies for stannylidyne complexes of molybdenum and tungsten have been investigated.17c Theoretical calculations of 119Sn Mössbauer isomer shift (IS) and quadrupole splitting (ΔEQ) for MSn−R bonding in metal stannylidyne complexes have not been investigated so far. Therefore, it is interesting to assess whether DFT is a valuable aid in quantitative predictions of 119Sn Mössbauer isomer shift (IS) and quadrupole splitting (ΔEQ) for MSn−R bonding. The 119Sn isomer shift (IS) and quadrupole splitting (ΔEQ) can be anticipated using DFT methods from the calculated electron density at the nucleus ρ(0) and the electric field gradient (EFG), respectively.41−54 The isomer shift can be calculated using the formula

INTRODUCTION The chemistry of tin compounds has been a provocative subject in a variety of contexts, spanning basic research and industrial applications.1−3 There are three magnetically active isotopes of tin with natural abundance: 119Sn (8.58%), 117Sn (7.61%), and 115 Sn (0.35%). Because of its reasonably high natural abundance and magnetogyric ratio, the 119Sn isotope is generally preferred for Mössbauer and NMR studies.1 The chemistry of transition-metal carbyne complexes have attracted much as an important class of compounds involving triple bonds between a transition metal and carbyne ligands.4−7 By contrast, the transition-metal complexes of heavier ER (E = Si, Ge, Sn, Pb) ligands has been less flourished. Power and Simons have succeeded in isolating the first stable compound with a MGeR bond, [(η5-C5H5)(CO)2MoGeR] (where R = 2,6-bis(2,4,6-trimethylphenyl)phenyl) in 1996.8,9 Since then, a several metal ylidyne complexes are synthesized and spectroscopically as well as structurally characterized.10−30 The theoretical calculations of the MER bonds, to evaluate the nature of transition metal and heavier group 14 ligands interactions have been reported.31−40 Only seven examples of transition metal-stannylidyne complexes have been reported so far. First structurally characterized stannylidyne complex trans-[Cl(PMe3)4WSn− C6H3−2,6-Mes2)], 1Sn has been reported by Filippou’s group in 2003.14 In addition, Filippou and co-workers have isolated © 2015 American Chemical Society

IS = α(ρ(0) − C) + β

(1)

where α and β are constants and C is held constant in the fit. The Mössbauer quadrupole splitting ΔEQ is defined by the following equation: Received: August 20, 2015 Published: October 23, 2015 10849

DOI: 10.1021/acs.inorgchem.5b01921 Inorg. Chem. 2015, 54, 10849−10854

Article

Inorganic Chemistry Chart 1. Structurally Characterized Metal−Stannylidyne Complexes

Table 1. Selected Optimized Structural Parameters and Mayer Bond Orders for Stannylidyne Complexes of Molybdenum and Tungsten (1−6) at the TPSS-D3(BJ)/TZVP/ZORA Level of Theory Bond Distance (Å)

a

M−Sn

Sn−C

2.480 2.4808(4)

2.185 2.177(3)

2.486 2.4901(7)

2.179 2.179(5)

2.509 2.5034(4)

2.209 2.192(2)

2.492 2.504(1)

2.177 2.183(3)

2.464 2.4599(6)

2.151 2.161(4)

2.457 2.4641(7)

2.144 2.146 (3)

Bond Angle (deg)

Mayer Bond Orders

M−Sn−C

M−Sn

Sn−C

trans-[Cl(PMe3)4MoSn−C6H3−2,6-Mes2] (1) 180.0 2.12 0.77 178.02(9)a trans-[Cl(PMe3)4WSn−C6H3−2,6-Mes2] (2) 180.0 2.86 0.96 178.2(1)b trans-[Cl(dppe)2MoSn−C6H3−2,6-Mes2] (3) 172.3 1.83 0.82 175.88(5)c trans-[Cl(dppe)2WSn−C6H3−2,6-Mes2] (4) 169.2 2.37 0.99 175.88(8)b [(dppe)2MoSn−C6H3−2,6-Mes2]+ (5) 165.1 1.91 0.85 178.77(9)a [(dppe)2WSn−C6H3−2,6-Mes2]+ (6) 165.8 2.29 0.99 178.77(9)b

Sn−P

Sn−H

0.11, 0.11

0.18, 0.18

0.11, 0.12

0.20

0.14, 0.16

0.12

0.19

0.15

0.11

X-ray structural data for complexes 1 and 5.17b bX-ray structural data for complexes 2, 4, and 6.14,15 cX-ray structural data for complex 3.17a 1/2 ⎡ ⎛ Vxx − Vyy ⎞2 ⎤ 1 1 ΔEQ = eQVzz⎢1 + ⎜ ⎟⎥ ⎢⎣ 2 3 ⎝ Vzz ⎠ ⎥⎦

[Cl(PMe3)4MoSn−R] (1), trans-[Cl(PMe3)4WSn−R] (2), trans-[Cl(dppe)2MoSn−R] (3), trans-[Cl(dppe)2W Sn−R] (4), [(dppe)2MoSn−R]+ (5), and [(dppe)2WSn− R]+ (6) (R = C6H3−2,6-Mes2) at the DFT/TPSS-D3(BJ)/ def2-TZVP/ZORA level of theory. Theoretical estimate of the 119 Sn Mössbauer isomer shift (IS) and quadrupole splitting (ΔEQ) for MSn−R bonding in complexes 1−6 have been performed for the first time. The main goals of the present

(2)

where Vxx, Vyy, and Vzz are the electric field gradient (EFG) tensors, e is the proton charge, and Q is the quadrupole moment of the tin nucleus. In the present article, we report the optimized geometries of the real (nonmodeled) six stannylidyne complexes trans10850

DOI: 10.1021/acs.inorgchem.5b01921 Inorg. Chem. 2015, 54, 10849−10854

Article

Inorganic Chemistry

Figure 1. Optimized structures of the trans-[Cl(PMe3)4MoSn−R] (1), trans-[Cl(dppe)2MoSn−R] (3), and [(dppe)2MoSn−R]+ (5); R = C6H3−2,6-Mes2.

Table 2. Experimental and Calculated Values of 119Sn Mössbauer Isomer Shifts, Calculated Electron Density ρ(0) [a.u.−3] at the Sn Nucleus, and Loewdin Sn s Electrons (Ns(Sn)) for Metal−Stannylidyne Complexes of Molybdenum and Tungsten (Complexes 1−6) Isomer Shifts (mm/s) complex

a

trans-[Cl(PMe3)4MoSn−R] (1) trans-[Cl(PMe3)4WSn−R] (2) trans-[Cl(dppe)2MoSn−R] (3) trans-[Cl(dppe)2WSn−R] (4) [(dppe)2MoSn−R]+ (5) [(dppe)2WSn−R]+ (6)

−3 b

ρ(0) [a.u. ] 165298.53 165297.85 165299.99 165298.46 165299.47 165298.66

ISexp 2.436 2.380 2.503 2.396 2.467 2.422

± ± ± ± ± ±

c

0.002 0.002 0.002 0.002 0.002 0.003

IScalcdd

IScalcde

Ns(Sn)f

Np(Sn)g

2.410 2.371 2.493 2.406 2.464 2.418

2.433 2.376 2.500 2.395 2.462 2.415

9.06 8.94 9.20 8.98 9.12 9.02

19.98 19.95 19.92 19.76 19.75 19.66

a R = C6H3−2,6-Mes2. bCalculated electron density at the nucleus. cExperimental values of the isomer shift (mm/s). dCalculated values of the isomer shift (mm/s) using equation (IScalcd = 0.057[ρ(0) − 165200] − 3.206). eCalculated values of the isomer shift (mm/s) using equation (IScalcd = 0.477Ns(Sn) −1.888. fCalculated values of the Loewdin Sn s electrons Ns(Sn). gCalculated values of the Loewdin Sn p electrons Ns(Sn).



RESULTS AND DISCUSSION Geometries. The important geometrical parameters of the stannylidyne complexes of molybdenum and tungsten trans[Cl(PMe3)4MoSn−R] (1), trans-[Cl(PMe3)4WSn−R] (2), trans-[Cl(dppe)2MoSn−R] (3), trans-[Cl(dppe)2W Sn−R] (4), [(dppe)2MoSn−R]+ (5), [(dppe)2WSn−R]+ (6) calculated at the TPSS-D3(BJ)/TZVP/ZORA level of theory are presented in Table 1. Their geometry structures are shown in Figure 1 (only for the case of M = Mo, i.e., for complexes 1, 3, and 5; the structures of the tungsten complexes are very similar to molybdenum complexes). The optimized Cartesian coordinates of all studied complexes are given in the Supporting Information. The calculated geometry parameters of stannylidyne complexes of molybdenum and tungsten (1−6) (Table 1) are in excellent agreement with experimental values of M−Sn and Sn−C bond distances.14,15,17a,b The M−Sn bond distances in complexes 1−6 are markedly shorter than those expected from the sum of single bond radii (Mo−Sn = 2.78 Å, W−Sn = 2.77 Å) and much closer to sum of respective triple bond covalent radii (Mo−Sn = 2.45 Å, W−Sn = 2.47 Å.63 The calculated Pauling bond orders corresponding to the calculated MSn distances are 2.65 (1), 2.51 (2), 2.41 (3), 2.46 (4), 2.79 (5), and 2.76 (6).64 It can be inferred from these observations that the M−Sn bonds in the metal-stannylidyne complexes (1−6) have MSn triple-bond characters. The MSn bond distances are shortest for cationic complexes. As shown in Table 1, the Mayer bond orders for the M−Sn bonds are 2.12

study are (i) to investigate the structure of real stannylidyne, (ii) to explain the observed main trends in the variations of the 119 Sn Mössbauer isomer shift (IS), quadrupole splitting (ΔEQ), and (iii) to correlate them to the Sn local environments and with the experimental values.



COMPUTATIONAL METHODS

All calculations were carried out with the ORCA-3.0.3 software package.55 Geometry optimizations of stannylidyne complexes trans[Cl(PMe3)4MoSn−R] (1), trans-[Cl(PMe3)4WSn−R] (2), trans[Cl(dppe) 2MoSn−R] (3), trans-[Cl(dppe) 2WSn−R] (4), [(dppe)2MoSn−R]+ (5), and [(dppe)2WSn−R]+ (6) (R = C6H3-2,6-Mes2) were carried out without any symmetry restrictions with the scalar-relativistic ZORA formalism56 in a DFT framework employing the TPSS density functional57 and all electron TZVP basis set.58 The effects of Grimme’s dispersion interactions59 with Becke− Johnson damping,60 D3(BJ), have been employed. Frequency calculation on the optimized structures of complexes 1−6 has been performed which determines that structures are global minima on the potential energy surface. The electronic structures of the studied complexes are examined by Mayer bond order61 and Loewdin atomic charges.62 The calculation of the electron density at the tin nucleus (ρ(0)) and electric field gradient (EFG) electron components MSn−R bonding in the stannylidyne complexes (1−6) were performed with the scalarrelativistic ZORA formalism in a DFT framework employing the TPSS-D3(BJ) density functional and all-electron TZVP basis set. 10851

DOI: 10.1021/acs.inorgchem.5b01921 Inorg. Chem. 2015, 54, 10849−10854

Article

Inorganic Chemistry

tin in the studied complexes with MSn−R bonding is Sn(II). Isomer shifts of the metal stannylidyne complexes have smaller values than the two-coordinated stannylenes (:SnR2).41−44 The values of electron density ρ(0) at the Sn nucleus and 119 Sn isomer shifts (IS) for MSnR bonding in studied stannylidyne complexes are greater (i) for molybdenum complexes than the tungsten complexes, (ii) for six-coordinated complexes than the five-coordinated complexes (except complex 6), and (iii) for bidentate dppe ligand than the PMe3 ligand. On the basis of theoretical studies on structure and bonding analysis of stannylidyne complexes, it has been concluded that, for the relatively stronger MSnR bonding, the electronic population at the Sn atom will be smaller.31−40 Thus, the trends of electron density ρ(0) at the Sn nucleus, as well as the values of 119Sn isomer shifts (IS) are consistent with the results of structure and bonding energy analysis. The variations of ρ(0) may be correlated with changes in the electronic populations. The Sn s electrons strongly influence electron densities ρ(0), which explains the correlation between the 119Sn Mössbauer isomer shift and the Sn oxidation state. The variations of ISexptl, as a function of Sn s electrons (Ns) are shown in Figure 3, and clearly exhibit a linear trend. Estimated

(1), 2.86 (2), 1.83 (3), 2.37 (4), 1.91 (5), and 2.29 (6), which indicate the presence of M−Sn triple-bond characters. It has been observed previously that Mayer bond orders are usually lower than the classical integer values.65,66 The bond orders for the Sn−C bonds (0.77−0.99) in complexes 1−6 indicate that the Sn−C bond has single-bond character. The noncovalent Sn···P and Sn···H interactions have been observed. The bending of the M−Sn−C angle increases because of the relatively greater mixing of HOMO and LUMO.67 One of the factors for smaller M−Sn−C bond angle in the cationic complexes 5 and 6 is due to smaller values of LUMO − HOMO energies (1.748 eV in 5 and 1.658 eV in 6, compared to 1.801 eV in 3 and 1.946 eV in 4. Detailed theoretical studies on structure and bonding energy analysis of model metal− stannylidyne complexes of molybdenum and tungsten with a Mes substituent at the Sn atom have been reported.35−39 The 119Sn Isomer Shift. The calculated electron densities at the Sn nucleus (ρ(0)) and 119Sn isomer shifts (IScalcd) are reported in Table 2, along with the experimental 119Sn isomer shifts (ISexptl). Variations of the ρ(0) can be related with changes in the local electronic structure, which are dependent on the local environment (nature of nearest neighbors). The results of linear regression analysis of the theoretical electron densities at the tin nucleus versus the experimental isomer shifts are presented in Figure 2.

Figure 3. Experimental values of the 119Sn Mössbauer isomer shift (ISexptl) (mm/s) versus calculated values of the Loewdin Sn s electrons (Ns(Sn)) for the metal−stannylidyne complexes of molybdenum and tungsten (1−6).

Figure 2. Experimental values of the 119Sn isomer shift (IS, mm/s) versus calculated values of the electron density at the nucleus (ρ(0), a.u.−3) for the stannylidyne complexes of molybdenum and tungsten (1−6).

values of isomer shift (IScalcd) using the linear regression with the Ns(Sn) electron density are in excellent agreement with the ISexptl. The 119Sn Quadrupole Splitting. The calculated values of 119 Sn Mössbauer parameters η, Vzz, and V and the experimental ΔEQ(exptl) and calculated ΔEQ(calcd) values of quadrupole splitting are presented in Table 3. The correlation plot of the experimental ΔEQ(exptl) versus the corresponding calculated EFG values (V) for complexes (1−6) is shown in Figure 4. The variations of the experimental values of the quadrupole splitting ΔEQ are linearly correlated with the calculated EFG values. The linear fit for the six data points yielded a slope of 0.540. The calculated ΔEQ(calcd) values are evaluated by using the correlation of the experimental ΔEQ(exptl) with the calculated EFG components (the present expression for stannylidyne complexes and Lindh’s equation (ΔEQ(calcd) = (0.726 V ± 0.27) with −0.27).47

As seen in Table 2 and Figure 2, the values of ISexptl for complexes 1 and 4 deviate from the linear correlation. The lowest value of Sn−C bond order (0.77) and longer Sn−C bond distance (1.854 Å) for molybdenum complex (1) may be responsible for lower value of calculated electron density ρ(0), compared to molybdenum complexes 3 and 5. Moreover, deviation of the calculated electron density ρ(0) of complex 4 from the linear correlation may be due to the relative stronger noncovalent interactions Sn···P and Sn···H. The calculated values of the isomer shift for the complexes 1−6 are in good agreement with the experimental values.17c The results of the previous 119Sn Mössbauer studies on Sn(IV) and Sn(II) compounds divulge that the values of isomer shift are lower than ∼2 mm/s for the Sn(IV) compounds and are greater than ∼2.7 mm/s for the Sn(II) compounds. Experimental values (ISexptl, 2.38−2.50 mm/s) and calculated values (IScalcd, 2.37− 2.49 mm/s) of isomer shifts indicate that the oxidation state of 10852

DOI: 10.1021/acs.inorgchem.5b01921 Inorg. Chem. 2015, 54, 10849−10854

Article

Inorganic Chemistry

Table 3. Calculated Values of Electric Field Gradients (EFG), Experimental and Calculated Values of 119Sn Mössbauer Nuclear Quadrupole Splitting Parameter (mm/s) for Metal-Stannylidyne Complexes of Molybdenum and Tungsten (1−6) complex

ηa

Vzz

Vb

trans-[Cl(PMe3)4MoSn−R] (1) trans-[Cl(PMe3)4WSn−R] (2) trans-[Cl(dppe)2MoSn−R] (3) trans-[Cl(dppe)2WSn−R] (4) [(dppe)2MoSn−R]+ (5) [(dppe)2WSn−R]+ (6)

0.439 0.428 0.601 0.555 0.347 0.375

2.842 2.808 3.174 3.279 4.697 4.544

2.932 2.893 3.360 3.442 4.790 4.649

ΔEexpc

ΔEQ(calcd)d

ΔEQ(calcd)e

± ± ± ± ± ±

1.831 1.810 2.062 2.107 2.835 2.758

1.859 1.830 2.169 2.229 3.207 3.105

1.849 1.814 2.072 2.062 2.817 2.787

0.005 0.004 0.004 0.005 0.003 0.007

a η = (Vxx − Vyy)/Vzz with |Vzz| ≥ |Vyy| ≥ |Vxx|. bV = Vzz(1 + 1/3η2)1/2, in atomic units. cExperimental values of the nuclear quadrupole splitting parameter (in mm/s), obtained from the literature.17 dQuadrupole splitting parameter calculated through the slope obtained by the linear fitting of Figure 2. eQuadrupole splitting parameter calculated using Lindh’s equation (ΔEQ(calcd) = (0.726 V ± 0.27)) with −0.27.47

linear regression with an electron density Ns(Sn) are in excellent concord with the ISexptl.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.5b01921. Cartesian coordinates of the optimized geometries of stannylidyne complexes of molybdenum and tungsten (1−6) (PDF)



Figure 4. Experimental values of the 119Sn quadrupole splitting ΔEQ (mm/s) versus calculated values of the V = Vzz(1 + 1/3η2)1/2, in atomic units for the stannylidyne complexes of molybdenum and tungsten (1−6).

AUTHOR INFORMATION

Corresponding Author

*Tel.: 00 91 731 2460208. Fax: 00 91 731 2762342. E-mail: k_ k_pandey3@rediffmail.com. Notes

The calculated values of nuclear quadrupole splitting parameters (ΔEQ(calcd)) of 119Sn using the relation ΔEcalcd = (0.540 + 0.248) V) are closure to the experimental values.17c The quadrupole splitting parameter calculated using Lindh’s equation (ΔEcalcd = (0.726 ± 0.27) V) with −0.27,47 deviates more from the experimental values. The isomer shift (IS) and quadrupole splitting (ΔEQ) parameters are closely related to the electron density distribution. Thus, the values of ΔEQ(exptl) and ΔEQ(calcd) are largest for cationic complexes [(dppe)2M Sn−R]+ (M = Mo, W) and lowest for trans-[Cl(PMe3)4Mo Sn−R] (M = Mo, W; R = C6H3−2,6-Mes2).

The authors declare no competing financial interest.



REFERENCES

(1) Smith, P. J. Chemistry of Tin, 2nd Edition; Blackie Academic & Professional: London, 1998. (2) Davies, A. G. Organotin Chemistry, 2nd Edition; Wiley−VCH: Weinheim, Germany, 2004. (3) Davies, A. G.; Gielen, M.; Pannell, K. H. Tiekink, E. R. T., Eds. Tin Chemistry; Wiley; Chichester, U.K., 2008. (4) Kreissl, F. R. Transition-Metal Carbyne Complexes; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1993. (5) Garrison, J. C.; Youngs, W. J. Chem. Rev. 2005, 105, 3978−4008. (6) Diez-Gonzalez, S.; Marion, N.; Nolan, S. P. Chem. Rev. 2009, 109, 3612−3676. (7) Schrock, R. R. Chem. Rev. 2002, 102, 145−180. (8) Simons, R. S.; Power, P. P. J. Am. Chem. Soc. 1996, 118, 11966− 11967. (9) Pu, L.; Twamley, B.; Haubrich, S. T.; Olmstead, M. M.; Mork, B. V.; Simons, R. S.; Power, P. P. J. Am. Chem. Soc. 2000, 122, 650−656. (10) Filippou, A. C.; Philippopoulos, A. I.; Portius, P.; Neumann, D. U. Angew. Chem., Int. Ed. 2000, 39, 2778−2781. (11) Filippou, A. C.; Portius, P.; Philippopoulos, A. I. Organometallics 2002, 21, 653−661. (12) Mork, B. V.; Tilley, T. D. Angew. Chem., Int. Ed. 2003, 42, 357. (13) Filippou, A. C.; Philippopoulos, A. I.; Portius, P.; Schnakenburg, G. Organometallics 2004, 23, 4503−4512. (14) Filippou, A. C.; Portius, P.; Philippopoulos, A. I.; Rohde, H. Angew. Chem., Int. Ed. 2003, 42, 445−447. (15) Filippou, A. C.; Philippopoulos, A. I.; Schnakenburg, G. Organometallics 2003, 22, 3339−3341. (16) Filippou, A. C.; Ghana, P.; Chakraborty, U.; Schnakenburg, G. J. Am. Chem. Soc. 2013, 135, 11525−11528. (17) (a) Rohde, H. Die ersten Molybdän-Stannylin-Komplexe, Diploma Thesis, Humboldt-Universität zu Berlin: Berlin, 2002. (b) Rohde, H.



CONCLUSIONS Sn Mössbauer isomer shift (IS) and quadrupole splitting (ΔEQ) for MSnR bonding in metal-stannylidyne complexes trans-[Cl(PMe3)4MoSn−R] (1), trans-[Cl(PMe3)4WR] (2), trans-[Cl(dppe)2MoSn−R] (3), trans-[Cl(dppe)2W Sn−R] (4), [(dppe)2MoSn−R]+ (5), and [(dppe)2WSn− R]+ (6) have been interpreted for the first time from the calculated electron density at the nucleus and the electric field gradient (EFG), respectively. Calculations of optimized structures and 119Sn Mössbauer parameters were carried out at the DFT/TPSS-D3(BJ)/TZVPP/ZORA level of theory. Calculated values of the 119Sn isomer shift (IS) and quadrupole splitting (ΔEQ) parameters are in agreement with the experimental values. 17c Using the relations IS calcd = 0.057[ρ(0) − 165200] − 3.206 and ΔEQ(calcd) = (0.540 + 0.248) V, values of the 119Sn isomer shift (IS) and quadrupole splitting (ΔEQ) parameters can be calculated for MSnR bonding in metal−stannylidyne complexes. The variations of ISexptl, as a function of Sn s electrons (Ns(Sn)) also exhibit a linear trend. Calculated values of isomer shift (IScalcd) using the 119

10853

DOI: 10.1021/acs.inorgchem.5b01921 Inorg. Chem. 2015, 54, 10849−10854

Article

Inorganic Chemistry Synthesis und Untersuchungen zur Reactivität von Stannylidinkomplexen, Dissertation, University of Bonn: Bonn, Germany, 2007. (c) Rohde, H.; Menzel, M.; Renz, F.; Filippou, A. C. Hyperfine Interact. 2008, 185, 129−135. (18) Filippou, A. C.; Weidemann, N.; Schnakenburg, G.; Rohde, H.; Philippopoulos, A. I. Angew. Chem., Int. Ed. 2004, 43, 6512−6516. (19) Filippou, A. C.; Rohde, H.; Schnakenburg, G. Angew. Chem., Int. Ed. 2004, 43, 2243−2247. (20) Filippou, A. C.; Schnakenburg, G.; Philippopoulos, A. I.; Weidemann, N. Angew. Chem., Int. Ed. 2005, 44, 5979−5985. (21) Filippou, A. C.; Weidemann, N.; Philippopoulos, A. I.; Schnakenburg, G. Angew. Chem., Int. Ed. 2006, 45, 5987−5991. (22) Filippou, A. C.; Weidemann, N.; Schnakenburg, G. Angew. Chem., Int. Ed. 2008, 47, 5799−5802. (23) Filippou, A. C.; Chernov, O.; Stumpf, K. W.; Schnakenburg, G. Angew. Chem., Int. Ed. 2010, 49, 3296−3300. (24) Filippou, A. C.; Chernov, O.; Schnakenburg, G. Angew. Chem., Int. Ed. 2011, 50, 1122−1126. (25) Filippou, A. C.; Barandov, A.; Schnakenburg, G.; Lewall, B.; van Gastel, M.; Marchanka, A. Angew. Chem., Int. Ed. 2012, 51, 789−793. (26) Hashimoto, H.; Fukuda, T.; Tobita, H.; Ray, M.; Sakaki, S. Angew. Chem., Int. Ed. 2012, 51, 2930−2933. (27) Hicks, J.; Hadlington, T. J.; Schenk, C.; Li, J.; Jones, C. Organometallics 2013, 32, 323−329. (28) Balázs, G.; Gregoriades, L. J.; Scheer, M. Organometallics 2007, 26, 3058−3075. (29) Filippou, A. C.; Chakraborty, U.; Schnakenburg, G. Chem.Eur. J. 2013, 19, 5676−5687. (30) Hayes, P. G.; Xu, Z.; Beddie, C.; Keith, J. M.; Hall, M. B.; Tilley, T. D. J. Am. Chem. Soc. 2013, 135, 11780−11783. (31) Pandey, K. K.; Lein, M.; Frenking, G. J. Am. Chem. Soc. 2003, 125, 1660−1668. (32) Lein, M.; Szabó, A.; Kovacs, A.; Frenking, G. Faraday Discuss. 2003, 124, 365−378. (33) Takagi, N.; Yamazaki, K.; Nagase, S. Bull. Korean Chem. Soc. 2003, 24, 832−836. (34) Pandey, K. K.; Lledós, A. Inorg. Chem. 2009, 48, 2748−2759. (35) Pandey, K. K.; Patidar, P.; Power, P. P. Inorg. Chem. 2011, 50, 7080−7089. (36) Pandey, K. K.; Patidar, P. J. Organomet. Chem. 2012, 702, 59− 67. (37) Pandey, K. K.; Patidar, P. Polyhedron 2012, 37, 85−93. (38) Pandey, K. K.; Patidar, P. RSC Adv. 2014, 4, 13034−13044. (39) Pandey, K. K.; Patidar, P.; Bariya, P. K.; Patidar, S. K.; Vishwakarma, R. Dalton Trans. 2014, 43, 9955−9967. (40) Pandey, K. K.; Patidar, P.; Vishwakarma, R. Eur. J. Inorg. Chem. 2014, 2014, 2916−2923. (41) Long, G. J., Ed. Mössbauer Spectroscopy Applied to Inorganic Chemistry; Plenum: New York, 1984. (42) Haas, H.; Menningen, M.; Andreasen, H.; Damgaard, S.; Grann, H.; Pedersen, F. T.; Petersen, J. W.; Weyer, G. Hyperfine Interact. 1983, 15, 215−218. (43) Svane, A.; Christensen, N. E.; Rodriguez, C. O.; Methfessel, M. Phys. Rev. B: Condens. Matter Mater. Phys. 1997, 55, 12572. (44) Lippens, P. E.; Olivier-Fourcade, J.; Jumas, J. C. Hyperfine Interact. 2000, 126, 137−141. (45) Pyykkö, P. Mol. Phys. 2001, 99, 1617−1629. (46) Barone, G.; Silvestri, A.; Ruisi, G.; La Manna, G. Chem.Eur. J. 2005, 11, 6185−6191. (47) Krogh, J. W.; Barone, G.; Lindh, R. Chem.Eur. J. 2006, 12, 5116−5121. (48) Filatov, M. J. Chem. Phys. 2007, 127, 084101. (49) Filatov, M.; Dyall, K. G. Theor. Chem. Acc. 2007, 117, 333−338. (50) Kurian, R.; Filatov, M. J. Chem. Theory Comput. 2008, 4, 278− 285. (51) Barone, G.; Mastalerz, R.; Reiher, M.; Lindh, R. J. Phys. Chem. A 2008, 112, 1666−1672. (52) Kurian, R.; Filatov, M. J. J. Chem. Phys. 2009, 130, 124121. (53) Zwanziger, J. W. J. Phys.: Condens. Matter 2009, 21, 195501.

(54) Darriba, G. N.; Muñoz, E. L.; Errico, L. A.; Rentería, M. J. Phys. Chem. C 2014, 118, 19929−19939. (55) Neese, F. ORCA program system, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 73−78. (56) (a) Chang, C.; Pelissier, M.; Durand, Ph. Phys. Scr. 1986, 34, 394−404. (b) van Lenthe, E.; van Leeuwen, R.; Baerends, E. J.; Snijders, J. G. Int. J. Quantum Chem. 1996, 57, 281−293. (c) van Lenthe, E.; Baerends, E. J.; Snijders, J. G. J. Chem. Phys. 1996, 105, 6505. (d) van Lenthe, E.; Ehlers, A. E.; Baerends, E. J. J. Chem. Phys. 1999, 110, 8943. (57) (a) Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Phys. Rev. Lett. 2003, 91, 146401. (b) Perdew, J. P.; Tao, J.; Staroverov, V. N.; Scuseria, G. E. J. Chem. Phys. 2004, 120, 6898. (58) Pantazis, D. A.; Chen, X. Y.; Landis, C. R.; Neese, F. J. Chem. Theory Comput. 2008, 4, 908−919. (59) Grimme, S.; Ehrlich, S.; Goerigk, L. J. Comput. Chem. 2011, 32, 1456−1465. (60) (a) Becke, A. D.; Johnson, E. R. J. Chem. Phys. 2005, 123, 154101. (b) Johnson, E. R.; Becke, A. D. J. Chem. Phys. 2006, 124, 174104. (61) Mayer, I. Chem. Phys. Lett. 1983, 97, 270−274. (62) Loewdin, P. O. J. Chem. Phys. 1950, 18, 365. (63) Pyykkö, P.; Atsumi, M. Chem.Eur. J. 2009, 15, 12770−12779. (64) Pauling, L. The Nature of the Chemical Bond, 3rd Edition; Cornell University Press: New York, 1960; p 239. The relationship of bond order to length is dn = d1 − 0.71 log(n), where n is the bond order, d1 and dn are the lengths of bonds with bond order 1 and n, respectively. (65) Mayer, I.; Salvador, P. Chem. Phys. Lett. 2004, 383, 368−373. (66) Jacobsen, H. J. Comput. Chem. 2009, 30, 1093−1102. (67) Albright, T. A.; Burdett, J. K.; Whangbo, M.-H. Orbital Interactions in Chemistry, 2nd Edition; Wiley: New York, 2013; p 136.

10854

DOI: 10.1021/acs.inorgchem.5b01921 Inorg. Chem. 2015, 54, 10849−10854