• Keine Ergebnisse gefunden

Bonding Situation in Dimeric Group 15 Complexes [(NHC)2

N/A
N/A
Protected

Academic year: 2022

Aktie "Bonding Situation in Dimeric Group 15 Complexes [(NHC)2"

Copied!
11
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Bonding Situation in Dimeric Group 15 Complexes [(NHC)

2

(E

2

)] (E = N–Bi)

Nicole Holzmann and Gernot Frenking

Fachbereich Chemie, Philipps-Universität Marburg, Hans-Meerwein-Strasse, D-35032 Marburg (Germany)

Reprint requests to N. H.; E-mail:holzmann@staff.uni-marburg.deand G. F.; E-mail:frenking@chemie.uni-marburg.de

Z. Naturforsch.69a, 385 – 395 (2014) / DOI: 10.5560/ZNA.2014-0033 Received May 8, 2014 / revised May 19, 2014 / published online July 15, 2014 Dedicated to Professor Dr. Jörg Fleischhauer on the occasion of his 75th birthday

Quantum chemical calculations using density functional theory at the BP86 level in conjunction with triple-zeta polarized basis sets have been carried out for the title compounds. The nature of the bonding between the diatomic fragment and the NHC ligands is investigated with an energy decom- position analysis. The chemical bonds in the [(NHCMe)2(E2)] complexes can be discussed in terms of donor–acceptor interactions which consist of two NHCMe→E2←NHCMedonor components and two weaker components of the NHCMe←E2→NHCMeπ backdonation. The out-of-phase(+)/(−) contribution of theσdonation is always stronger than the in-phase(+)/(+)contribution. The elec- tronic reference state of N2in the dinitrogen complex [(NHCMe)2(N2)] is the highly excited 11Γg

state which explains the anti-periplanar arrangement of the ligands. The gauche arrangement of the ligands in the heavier homologues [(NHCMe)2(E2)] (E=P–Bi) may be discussed using either the excited 11Γgstate or the X1Σg+ground state of E2as reference states for the donor–acceptor bonds.

The EDA-NOCV calculations suggest that the latter bonding model is better suited for the complexes where E=As–Bi while the phosphorus complex is a borderline case.

Key words:Donor–Acceptor Complex; Bonding Analysis; Quantum Chemical Calculations.

1. Introduction

The stabilization of electronically unsaturated main group compounds by coordination of σ donor lig- ands L such as N-heterocyclic carbenes (NHCs) or phosphanes PR3 has been an area of intensive ex- perimental and theoretical research in recent years.

In particular, complexes of monoatomic species such as carbones CL2 [1–18], silylones SiL2 [19–25], germylones GeL2[24,26,27], and adducts of diatomic molecules of groups 13 – 15 such as B2L2 [28–30], Si2L2[31], Ge2L2[32,33], Sn2L2[34], P2L2[35–39], and As2L2 [40] have been isolated, and they were investigated with quantum chemical methods. A par- ticularly interesting species is the dinitrogen com- plex N2L2. Quantum chemical calculations of group 14 and 15 complexes L→E2←L by Wilson et al. [41]

showed that all complexes are stable toward dissocia- tion of the ligands L except for E=N. However, stable adducts [(NHC)2N2] [42] and even [(PPh3)2N2] [43],

which according to the calculations [41] is ther- modynamically unstable toward nitrogen loss by

∼90 kcal/mol, could become isolated and were found to be very stable. The surprisingly high kinetic stabil- ity of [(PPh3)2N2] which melts at 184 – 186C and de- composes only above 215C has been explained with the very strong bidentate Lewis acidity of N2 in the electronically highly excited 11Γgstate [44].

The ligands L in [(NHC)2N2] and [(PPh3)2N2] exhibit an anti-periplanar arrangement in the com- plexes. In contrast, the conformation of the phos- phorus analogues [(NHCR)2(P2)] depends on the size of the substituents R at the nitrogen atoms. The X-ray structure of [(NHCDipp)2(P2)] (Dipp = 2,6- diisopropylphenyl) gives an anti-periplanar orienta- tion of the ligands but the complex [(NHCMes)2(P2)]

(Mes = mesityl) possesses a gauche conformation with a dihedral angle CNHC–P–P–CNHCof 134.1[35].

A gauche conformation is also found for the com- plex [(CAACDipp)2(P2)] (CAAC=cyclic alkyl amino

© 2014 Verlag der Zeitschrift für Naturforschung, Tübingen·http://znaturforsch.com

(2)

2 2

Table 1. EDA-NOCV results of the donor–acceptor interac- tions of [(NHCMe)2(E2)] at BP86/TZ2P+. Fragments are (NHCMe)2and (E2). All energies in kcal/mol.

[(NHCMe)2(E2)] (1Γg) [(NHCMe)2(E2)] (X1Σg+)

E=N C2

∆Eint −374.3

∆EPauli 1237.7

∆Eelstat −530.5 (32.9%)

∆Eorb −1081.6 (67.1%)

∆E1 −424.4 (39.2%)

∆E2 −450.0 (41.6%)

∆E3 −92.0 (8.5%)

∆E4 −34.4 (3.2%)

∆Erest −80.8 (7.5%)

E=P C2 C2

Eint −218.0 −78.0

EPauli 677.3 853.5

Eelstat −409.4 (45.7%) −441.9 (47.4%)

Eorb −485.9 (54.3%) −489.6 (52.6%)

E1 −169.1 (34.8%) −163.0 (33.3%)

E2 −204.6 (42.1%) −243.3 (49.7%)

E3 −47.8 (9.8%) −28.4 (5.8%)

E4 −26.5 (5.5%) −25.1 (5.1%)

Erest −37.9 (7.8%) −29.8 (6.1%)

E=As C2 C2

Eint −181.9 −65.5

EPauli 608.8 681.6

Eelstat −384.5 (48.6%) −385.9 (51.7%)

Eorb −406.1 (51.4%) −361.1 (48.3%)

E1 −142.0 (35.0%) −135.7 (37.6%)

E2 −190.4 (46.9%) −164.0 (45.4%)

E3 −28.2 (6.9%) −20.4 (5.6%)

E4 −20.3 (5.0%) −19.9 (5.5%)

Erest −25.2 (6.2%) −21.1 (5.8%)

E=Sb C2 C2

Eint −149.5 −53.6

EPauli 477.1 514.6

Eelstat −318.6 (50.9%) −318.5 (56.1%)

Eorb −308.0 (49.2%) −249.7 (44.0%)

E1 −103.4 (33.6%) −97.0 (38.8%)

E2 −154.9 (50.3%) −108.7 (43.5%)

E3 −17.5 (5.7%) −14.2 (5.7%)

E4 −13.6 (4.4%) −13.6 (5.4%)

Erest −18.6 (6.0%) −16.2 (6.5%)

E=Bi C2 C2

Eint −133.0 −46.9

EPauli 476.0 423.3

Eelstat −295.8 (48.6%) −272.2 (57.9%)

Eorb −313.2 (51.4%) −198.1 (42.1%)

E1 −105.5 (33.6%) −67.3 (34.0%)

E2 −173.5 (55.4%) −97.4 (49.2%)

E3 −10.9 (3.5%) −9.8 (4.9%)

E4 −10.6 (3.4%) −10.8 (5.5%)

Erest −12.7 (4.1%) −12.8 (6.5%)

carbene) which exhibits a dihedral angle CCAAC–P–

P–CCAAC of 149.2 [36,37]. The arsenic complex [(NHCDipp)2(As2)] has an anti-periplanar arrangement

Table 2. Wiberg bond indices (WBI) of the E–E, E–CNHC, and CNHC–N bonds at BP86/def2-TZVPP.

E–E E–CNHC CNHC–N

NHCMe 1.27

[(NHCMe)2(N2)] 1.08 1.54 1.08/1.09

[(NHCMe)2(P2)] 1.04 1.27 1.13/1.14

[(NHCMe)2(As2)] 1.02 1.19 1.16

[(NHCMe)2(Sb2)] 1.09 1.02 1.18/1.19

[(NHCMe)2(Bi2)] 1.15 0.92 1.20

of the ligands [35]. The bonding analysis of the group 13 complexes L→E2←L, where E=B–In, and the group 14 homologues with E=Si–Pb which pos- sess either a linear (E= B) or an anti-periplanar ar- rangement of the ligands L can straightforwardly be done because the choice of the electronic reference state of E2was easy. This appears not the case for the heavier group 15 complexes [(NHC)2(E2)] where E= P–Bi. Therefore, we investigated the electronic struc- ture of all group 15 adducts [(NHC)2(E2)] (E=N–Bi) using an energy decomposition analysis. The results are reported below.

2. Methods

Geometry optimizations have been carried out us- ing TurboMole 6.1 optimizer [45] and gradients at the BP86 [46,47] /def2-TZVPP [48] level of theory. Sta- tionary points were characterized as minima by calcu- lating the Hessian matrix analytically. For all calcu- lations the resolution-of-identity method has been ap- plied [49].

For the bonding analyses we optimized the molecules at BP86 using uncontracted Slater-type orbitals (STOs) as basis functions [50] with the program package Amsterdam density functional ADF2009.01 [51,52]. The latter basis sets for all el- ements have triple-ζ quality augmented by two sets of polarization functions (ADF-basis set TZ2P). This

Table 3. NBO partial charges of the (E2) moiety and the coor- dinating carbon atom at BP86/def2-TZVPP. All charges in e.

(E2) CNHC

NHCMe 0.04

[(NHCMe)2(N2)] −0.89 0.47 [(NHCMe)2(P2)] −0.19 0.04 [(NHCMe)2(As2)] −0.26 0.08 [(NHCMe)2(Sb2)] −0.29 0.09 [(NHCMe)2(Bi2)] −0.33 0.12

(3)

2 2

Table 4. NBO results of the E–E and E–CNHCbonds at BP86/def2-TZVPP.

orbital occupation occupation % (E) % s(E)a % p(E)a % (C) % s(C)a % p(C)a bonding anti-bonding

[(NHCMe)2(N2)] N–N 1.97 0.02 50.0 25.0 74.8

N 1.89 39.8 60.1

N–C 1.99 0.02 56.0 35.4 64.1 44.0 39.6 60.4

N–C 1.94 0.49 64.7 0.0 99.8 35.3 0.0 100.0

[(NHCMe)2(P2)] P–P 1.92 0.04 50.0 13.6 85.9

P 1.94 71.2 28.8

P–C 1.97 0.04 33.5 15.4 84.1 66.5 41.8 57.9

P–C 1.90 0.62 66.4 0.2 99.6 33.6 0.2 99.5

[(NHCMe)2(As2)] As–As 1.93 0.04 50.0 9.2 90.5

As 1.96 80.2 19.7

As–C 1.97 0.05 31.4 10.6 89.0 68.6 41.2 58.8

As–C 1.91 0.64 70.3 0.2 99.6 29.7 0.1 99.7

[(NHCMe)2(Sb2)] Sb–Sb 1.94 0.05 50.0 7.2 92.5

Sb 1.97 85.8 14.2

Sb 1.59 0.4 99.5

Sb–C 1.96 0.07 26.9 6.9 92.8 73.1 39.9 60.1

[(NHCMe)2(Bi2)] Bi–Bi 1.95 0.03 50.0 4.7 95.1

Bi 1.98 91.4 8.6

Bi 1.64 0.4 99.6

Bi–C 1.96 0.09 25.0 3.9 95.9 75.0 39.2 60.8

aThe remaining small contributions which add to 100% come from polarization functions.

level of theory is denoted BP86/TZ2P. An auxiliary set of s, p, d, f, and g STOs was used to fit the molecular densities and to represent the Coulomb and exchange potentials accurately in each self-consistend field cy- cle [53]. Scalar relativistic effects have been incorpo- rated by applying the zero-order regular approximation (ZORA) in all ADF calculations [54].

The interatomic interactions were investigated by means of an energy decomposition analysis (EDA, also termed extended transition state method – ETS) devel- oped independently by Morokuma [55] and by Ziegler and Rauk [56]. The bonding analysis focuses on the instantaneous interaction energy∆Eintof a bond A–B between two fragments A and B in the particular elec- tronic reference state and in the frozen geometry of AB. The interaction energy is divided into three main components:

∆Eint=∆Eelstat+∆EPauli+∆Eorb. (1) The term ∆Eelstat corresponds to the quasiclassi- cal electrostatic interaction between the unperturbed charge distributions of the prepared atoms and is usu- ally attractive. The Pauli repulsion ∆EPauli is the en- ergy change associated with the transformation from the superposition of the unperturbed wavefunctions of the isolated fragments which corresponds to a Hartree product to the wavefunctionΨ0=NA[Ψˆ AΨB], which

properly obeys the Pauli principle through explicit antisymmetrization ( ˆA operator) and renormalization (N= constant) of the product wavefunction.∆EPauli comprises the destabilizing interactions between elec- trons of the same spin on either fragment. The orbital interaction∆Eorb which can be associated with cova- lent bonding accounts for mixing of the fragment or- bitals and polarization effects. The∆Eorbterm can be decomposed into contributions from each irreducible representation of the point group of the interacting system. Further details on the EDA method [51,52]

and its application to the analysis of the chemical bond [57–60] can be found in the literature.

The EDA-NOCV [61] method combines charge (NOCV) and energy (EDA) decomposition schemes to decompose the deformation density which is associ- ated with the bond formation,∆ρ, into different com- ponents of the chemical bond. The EDA-NOCV calcu- lations provide pair wise energy contributions for each pair of interacting orbitals to the total bond energy.

NOCV (natural orbital for chemical valence) [62–64]

is defined as the eigenvector of the valence operator ˆV given by

ˆ iiΨi. (2) In the EDA-NOCV scheme the orbital interaction term

∆Eorbis given by

(4)

2 2

Fig. 1. Optimized geometries at RI-BP86/def2-TZVPP and most important bond lengths [Å] and angles [°] of [(NHCMe)2(E2)]. Experimental values of substituted analogues are given in parentheses.

∆Eorb=

k

∆Ekorb=

N 2

k=1

vk

h−F−k,−kTS +Fk,kTSi

, (3)

in whichF−k,−kTS andFk,kTSare diagonal Kohn–Sham ma- trix elements corresponding to NOCVs with the eigen- values−vk andvk, respectively. The∆Ekorbterms are assigned to a particular type of bond by visual in-

(5)

2 2

Fig. 2. Schematic view of the electronic reference states of E2in the complexes [(NHCMe)2(E2)] (E=N–Bi). (a) X1Σg+ ground state; (b)1Γgexcited state; (c) calculated excitation energies at RI-BP86/def2-TZVPP.

spection of the shape of the deformation density∆ ρk. The EDA-NOCV scheme thus provides both qualita- tive (∆ρorb) and quantitative (∆Eorb) information about the strength of orbital interactions in chemical bonds, even in molecules withC1symmetry. For more details we refer to [63,64].

Wiberg bond indices (WBI), partial charges, and Lewis structures were obtained using natural bond or- bitals NBO 3.1 [65] as implemented in Gaussian 09 Rev. C.01 [66]. NBO electron densities were generated from a single-point calculation of the molecules with BP86/def2-TZVPP basis sets in Gaussian09 from the TURBOMOLE 6.1 optimized geometries.

3. Results

The optimized geometries of the dimeric group 15 complexes [(NHCMe)2(E2)] (E = N–Bi), including

most important bond lengths and angles, are displayed in Figure1. Experimental values of substituted ana- logues are given in parentheses [35–40,42].

The agreement between the experimental and the calculated structures for the nitrogen compound is good, both show a slight deviation from a perfect anti- periplanar arrangement of the NHCMe ligands and a comparably long N–N bond of about 1.4 Å. As we described before for [(PPh3)2(N2)] [44], the electronic reference state for the N2moiety is the excited(1)1Γg state (Fig.2b). The NHCMe→N2←NHCMecharge do- nation occurs from theσdonor orbitals of NHCMeinto the vacant in-plane bonding 1πu0 and anti-bonding 1πg0 orbitals of N2. The NHCMe←N2→NHCMeπbackdo- nation takes place from the occupied out-of-plane π orbitals into the vacantπaccepting orbitals of NHCMe. The charge flow which is associated with the NHCMe→N2←NHCMe σ donation and the NHCMe←N2→NHCMe π backdonation is nicely visualized by the calculated deformation densities∆ ρ which are shown in Figure3, along with the relevant molecular orbitals (MOs). The third and fourth column show the frontier orbitals of N2 and NHCMe which yield the occupied MOs of [(NHCMe)2(N2)] that are shown in the second column. The first column displays the deformation densities which are coupled to the four major orbital interactions, i. e. (+)/(+) and (+)/(−) σ donation (Figs. 3a and b) and (+)/(−) and(+)/(+)πbackdonation (Figs.3c and d). Defor- mation densities visualize the electron density flow of the interactions: Areas of depleting electron density are shown in red (dark grey) and areas of accumulating electron density are shown in light blue (light grey).

The calculated [(NHCMe)2(E2)] complexes of the heavier group 15 atoms all have gauche oriented lig- ands with decreasing dihedral angle CNHC–E–E–CNHC when E becomes heavier (Fig.1). For [(NHCMe)2(P2)]

this orientation correlates with the experimentally known N-mesityl substituted analogue, where the CNHC–P–P–CNHCangles agree quite well with the cal- culated data (132.8for [(NHCMe)2(P2)] calc., 134.1 for [(NHCMes)2(P2)] exp. [35]). However, the experi- mental structures of the diisopropyl substituted com- plexes of P2and As2show an anti-periplanar arrange- ment of the NHCDipp ligands [35,40]. Wilson et al.

could show that the distortion in different group 15 complexes [(NHCX)2(E2)] (X=H, Me, Ph) decreases with the steric demand of the substituent H>Me>

Ph [41]. But the large deviation between the NHCDipp

(6)

2 2

Fig. 3 (colour online). Plot of deformation densities∆ρ of the pairwise orbital interactions between N2in its1Γgstate and (NHCMe)2 in [(NHCMe)2(N2)], associated energies∆E in kcal/mol and eigenvaluesε. Shape of the most important inter- acting occupied and vacant orbitals of N2and (NHCMe)2and resulting molecular orbitals. Left:∆ρ<0 in red (dark grey),

∆ρ>0 in light blue (light grey).

and the NHCMesphosphorus complexes indicates that electronic reasons may also play a role.

While the electronic reference state of N2 in [(NHCMe)2(N2)] could easily be determined, the same is more complicated for the heavier homologues. Due to the gauche orientation of the ligands more than two frontier orbitals of E2 can be considered to ac- cept electron density from theσ donor orbitals of the NHCMeligands. Thus, it is also possible to donate elec- tronic charge to the E2 fragment in its ground state X1Σg+where both anti-bonding 1πgorbitals are vacant (Fig. 2a). The (+)/(−) σ donation could then take place into one 1πgorbital, where the overlap with the σ donor orbitals of the ligands is also significant. Fur- thermore, this could be an explanation of the strong deviation from the planar ligand arrangement as in [(NHCMe)2(N2)] and it would not require the excita- tion energy X1Σg+→(1)1Γg(Fig.2c).

In order to determine the correct reference state, we carried out two EDA-NOCV analyses of the heavier group 15 complexes [(NHCMe)2(E2)] where E=P–Bi.

Besides the ligand fragment (NHCMe)2with its high- est occupied molecular orbital (HOMO) and HOMO–

1 corresponding to the(+)/(+)and(+)/(−)combi- nation of theσ donating orbitals, the E2 fragment is calculated in its excited (1)1Γg state and the ground state X1Σg+, respectively. The results are given in Ta- ble1.

The data in Table1show that the EDA-NOCV cal- culations using the excited1Γgstate of E2yield much larger interaction energies∆Eintthan the calculations which employ the X1Σg+ ground state. However, the crucial term which indicates the most suitable choice of the electronic states of the interaction fragments is the orbital energy term ∆Eorb. The size of the latter is a measure for the change of the electronic struc-

(7)

2 2

Fig. 4 (colour online). Plot of deformation densities∆ρof the pairwise orbital interactions between P2in its1Γg state and (NHCMe)2in [(NHCMe)2(P2)], associated energies∆Ein kcal/mol and eigenvaluesεfor all [(NHCMe)2(E2)] (E=P–Bi).

Shape of the most important interacting occupied and vacant orbitals of P2and (NHCMe)2and resulting molecular orbitals.

Left:∆ρ<0 in red (dark grey),∆ρ>0 in light blue (light grey).

tures of the fragments. The smaller the absolute value of ∆Eorb, the closer is the electronic structure of the fragments to the final molecule. Table1 shows that the EDA-NOCV calculations of [(NHCMe)2(P2)] give a slightly smaller value of∆Eorb=−485.9 kcal/mol when the excited 1Γg state of P2 is used, while the calculations using the electronic ground state give

∆Eorb= −489.6 kcal/mol. For the heavier systems [(NHCMe)2(E2)] where E=As–Bi, the EDA calcula- tions using the ground state of E2give always a clearly

smaller value for∆Eorbthan the calculations with the excited1Γgstate. The EDA-NOCV results thus suggest that the NHCMe–E2–NHCMe bonds with E=As–Bi should be discussed in terms of donor–acceptor inter- actions using the ground state rather than the excited

1Γgstate of E2while NHCMe–P2–NHCMeis a border- line case.

The relevant deformation densities of the EDA- NOCV with their corresponding molecular and frag- ment orbitals are given in Figure4(1Γg) and Figure5

(8)

2 2

Fig. 5 (colour online). Plot of deformation densities∆ρof the pairwise orbital interactions between P2in its X1Σg+state and (NHCMe)2in [(NHCMe)2(P2)], associated energies∆Ein kcal/mol and eigenvaluesεfor all [(NHCMe)2(E2)] (E=P–Bi).

Shape of the most important interacting occupied and vacant orbitals of P2and (NHCMe)2and resulting molecular orbitals.

Left:∆ρ<0 in red (dark grey),∆ρ>0 in light blue (light grey).

(X1Σg+). Two of the orbital interactions for the dif- ferent occupations are qualitatively and quantitatively alike: The NHCMe→E2←NHCMe (+)/(+) σ dona- tion from the (NHCMe)2HOMO–1 into the vacant in- plane anti-bonding 1πg orbital on E2. It contributes with 30 – 40% to the orbital energy and can be assigned to the HOMO–8 of the complex (Figs. 4a and 5a).

The NHCMe←E2→NHCMeπ backdonation from the occupied out-of-plane bonding 1πu orbital of E2into

the lowest unoccupied molecular orbital (LUMO) of (NHC)2 accounts for about 3 – 6% of the orbital en- ergy (Figs.4d and5d). The resulting molecular orbital is the HOMO–1.

The(+)/(−)σdonation in the complex with E2in the excited1Γgstate (Fig.4b) occurs into the vacant in- plane bonding 1πuorbital. The shape of the resulting molecular HOMO–7 features both fragment orbitals which can easily be identified. Additionally, the defor-

(9)

2 2

mation density∆ρ1shows areas of electron density de- pletion on the E2moiety perpendicular to the accepting 1πuorbital. These are in shape of the occupied out-of- plane anti-bonding 1πgorbital, which according to the EDA-NOCV participates in this interaction. As a re- sult, the amount of shifted electron density gets quite large and also the energetic contribution dominates the orbital term by 42.1% for the phosphorus complex and up to 55.4% for [(NHCMe)2(Bi2)]. A smaller part of less than 10% of the orbital term is the π back- donation from the occupied out-of-plane anti-bonding 1πg orbital of E2 into the LUMO+1 on (NHCMe)2 (Fig.4c). The [(NHCMe)2(P2)] HOMO matches the oc- cupied anti-bonding 1πgorbital of P2(HOMO), which is delocalized towards the carbon atoms of NHC inπ fashion.

For the complexes with E2 in its ground state, the(+)/(−)σdonation NHCMe→E2←NHCMetakes place into the in-plane anti-bonding 1πgorbital, which is occupied in the excited state (Fig. 5b). The result- ing molecular HOMO lies energetically higher than its corresponding fragment orbitals. Still, the associated energy contribution of this donation dominates the or- bital term by 40 – 50%. The amount of the electron density shift shown in deformation density∆ ρ1is even larger than in the(+)/(−)σdonation into E2of its ex- cited state as described above. In the ground state, the electron density shift into the E2goccurs not only from the ligands, the occupied in-plane bonding 1πu (HOMO) participates as well. This is the reversed sit- uation as for the major contribution∆E1to the orbital term of the excited case. But there, the shape of the re- sulting molecular HOMO is better represented by the chosen fragment orbitals than for the complexes with E2in its ground state. This is also the case for the inter- action shown in Figure5c, theπbackdonation into the accepting orbital on (NHCMe)2from the occupied in- plane bonding 1πuorbital on E2. Although the corre- sponding molecular orbital HOMO–7 is low in energy, the contribution of this interaction to∆Eorbis only mi- nor (5 – 6%). Also, its shape shows predominantly fea- tures of the participating E2HOMO, but the accepting LUMO+1 of (NHC)2is not represented well.

For further information on the bonding situation in [(NHCMe)2(E2)] (E = P–Bi) and the electronic ref- erence state of E2, we carried out NBO calculations which are independent of the reference state. The cal- culated Wiberg bond indices (WBIs) of the E–E, E–

CNHC, and CNHC–N bonds are shown in Table2, while

the partial charges of the bonding atoms are given in Table3.

The charge donation from the ligands to the E2moi- ety is large only for the [(NHCMe)2(N2)] complex. This corresponds to the high electronegativity of nitrogen that exceeds those of the higher group 15 elements by far. The high WBI for the E–CNHC donor–acceptor bond of the nitrogen complex of 1.54 indicates some double bond character which correlates with the large amount of shifted electron density in theπ backdona- tion seen in the EDA-NOCV. Additionally, this back- donation lowers the bond index of the CNHC–N bond by raising the electron density in the CNHC–N anti- bonding orbital. The bond index of the donor–acceptor bond is still larger than one for the complexes with E

=P and As but decreases continuously for the heav- ier analogues. This is represented well in the EDA- NOCV of the phosphorus and arsenic complexes with E2in its excited1Γg state, where theπ backdonation in∆ρ3(Fig.4c) shows a larger energetic contribution and more shifted electron density than in their heav- ier analogues. The E–E bond index which varies only slightly from nitrogen to bismuth indicates essentially a single bond for all complexes. This can be interpreted in different ways. There is an E–E single bond present in E2in the excited1Γgstate. In the EDA-NOCV, the (+)/(−) σ donation contributes most to the orbital energy and occurs into a vacant E–E bonding orbital (Fig.4b), which should increase the WBI of this single bond. The σ donation (Fig.4a) into the correspond- ing anti-bonding orbital is weaker. Additionally, the π backdonation from the occupied anti-bonding E–

E π orbital (Fig. 4c) exceeds the backdonation from bonding the E–E orbital (Fig.4d). On the other hand, there is a triple bond in E2 in its ground state 1Σg+. In the EDA-NOCV it could be seen that bothσ do- nations occur into vacant anti-bonding orbitals of E2 (Figs.5a and b) while bothπbackdonations take place from occupied bonding orbitals on E2(Figs.5c and d).

This would decrease the bond order of the triple bond massively, which would also correspond to the given WBIs in Table2. Thus, the NBO data in Table2do not discriminate between the models using different elec- tronic reference states.

Orbital information on the E–E and E–C bonds and E lone pair orbitals from the Lewis structure determi- nation of the NBO calculation are given in Table4.

The results suggest that the [(NHCMe)2(E2)] com- plexes have a E–E single bond where the hybridiza-

(10)

2 2

tion changes toward higher %p character from E = N (74.8%) to E =Bi (95.1%). The calculations give Lewis structures with E=CNHCdouble bonds and one lone-pair orbital at atom E for E =N, P, As. For the heavier system where E=Sb, Bi the calculations give E–CNHCsingle bonds and two lone-pair orbitals at An- timony and Bismuth. Note, however, that the E–CNHC π bonds in the lighter systems have a large occupation from the antibonding orbital. The NBO data should rather be interpreted as evidence for strongerπ back- donation in the lighter complexes which agrees with the EDA-NOCV results.

4. Summary and Conclusion

The results of this work can be summarized as follows. The chemical bonds of the diatomic fragments E2 and the NHCMe ligands in the

[(NHCMe)2(E2)] complexes can be discussed in terms of donor–acceptor interactions which consist of two NHCMe→E2←NHCMe donor components and two weaker components of the NHCMe←E2→NHCMe π backdonation. The out-of-phase(+)/(−)contribution of theσdonation is always stronger than the in-phase (+)/(+) contribution. The electronic reference state of N2 in the dinitrogen complex [(NHCMe)2(N2)] is the highly excited 11Γgstate which explains the anti- periplanar arrangement of the ligands. The gauche ar- rangement of the ligands in the heavier homologues [(NHCMe)2(E2)] (E =P–Bi) may be discussed using either the excited 11Γgstate or the X1Σg+ground state of E2as reference states for the donor–acceptor bonds.

The EDA-NOCV calculations suggest that the latter bonding model is better suited for the complexes where E=As–Bi while the phosphorus complex is a border- line case.

[1] C. A. Dyker, V. Lavallo, B. Donnadieu, and G.

Bertrand, Angew. Chem.120, 3250 (2008).

[2] C. A. Dyker, V. Lavallo, B. Donnadieu, and G.

Bertrand, Angew. Chem. Int. Ed.47, 3206 (2008).

[3] A. Fürstner, M. Alcarazo, R. Goddard, and C. W.

Lehmann, Angew. Chem.129, 3254 (2008).

[4] A. Fürstner, M. Alcarazo, R. Goddard, and C. W.

Lehmann, Angew. Chem. Int. Ed.47, 3210 (2008).

[5] R. Tonner, F. Öxler, B. Neumüller, W. Petz, and G.

Frenking, Angew. Chem.118, 8206 (2006).

[6] R. Tonner, F. Öxler, B. Neumüller, W. Petz, and G.

Frenking, Angew. Chem. Int. Ed.45, 8038 (2006).

[7] R. Tonner and G. Frenking, Angew. Chem.119, 8850 (2007).

[8] R. Tonner and G. Frenking, Angew. Chem. Int. Ed.46, 8695 (2007).

[9] R. Tonner and G. Frenking, Chem. Eur. J. 14, 3260 (2008).

[10] R. Tonner and G. Frenking, Chem. Eur. J. 14, 3273 (2008).

[11] R. Tonner, G. Heydenrych, and G. Frenking, Chem.

Phys. Chem.9, 1474 (2008).

[12] G. Frenking and R. Tonner, Pure Appl. Chem.81, 597 (2009).

[13] G. Frenking and R. Tonner, WIREs Comput. Mol. Sci.

1, 869 (2011).

[14] S. Klein, R. Tonner and G. Frenking, Chem. Eur. J.16, 10160 (2010).

[15] S. Klein and G. Frenking, Angew. Chem. 122, 7260 (2010).

[16] S. Klein and G. Frenking, Angew. Chem. Int. Ed.49, 7106 (2010).

[17] C. Esterhuysen and G. Frenking, Chem. Eur. J. 17, 9944 (2011).

[18] C. Esterhuysen and G. Frenking, J. Chem. Soc., Dalton Trans.42, 13349 (2013).

[19] K. C. Mondal, H. W. Roesky, F. Klinke, M. C. Schw- arzer, G. Frenking, B. Niepötter, H. Wolf, R. Herbst- Irmer, and D. Stalke, Angew. Chem.125, 3036 (2013).

[20] K. C. Mondal, H. W. Roesky, F. Klinke, M. C. Schw- arzer, G. Frenking, B. Niepötter, H. Wolf, R. Herbst- Irmer, and D. Stalke, Angew. Chem. Int. Ed.52, 2963 (2013).

[21] Y. Xiong, S. Yao, S. Inoue, J. D. Epping, and M. Driess, Angew. Chem.125, 7287 (2013).

[22] Y. Xiong, S. Yao, S. Inoue, J. D. Epping, and M. Driess, Angew. Chem. Int. Ed.52, 7147 (2013).

[23] N. Takagi, T. Shimizu, and G. Frenking, Chem. Eur. J.

15, 3448 (2009).

[24] N. Takagi, T. Shimizu, and G. Frenking, Chem. Eur. J.

15, 8593 (2009).

[25] N. Takagi, R. Tonner, and G. Frenking, Chem. Eur. J.

18, 1772 (2012).

[26] Y. Xiong, S. Yao, G. Tan, S. Inoue, and M. Driess, J.

Am. Chem. Soc.135, 5004 (2013).

[27] Y. Li, K. C. Mondal, H. W. Roesky, H. Zhu, P. Stoll- berg, R. Herbst-Irmer, D. Stalke, and D. M. Andrade, J.

Am. Chem. Soc.135, 12422 (2013).

[28] H. Braunschweig, R. D. Dewhurst, K. Hammond, J.

Mies, K. Radacki, and A. Vargas, Science336, 1420 (2012).

[29] N. Holzmann, A. Stasch, C. Jones, and G. Frenking, Chem. Eur. J.17, 13517 (2011).

(11)

2 2

[30] G. Frenking and N. Holzmann, Science 336, 1394 (2012).

[31] Y. Wang, Y. Xie, P. Wei, R. B. King, H. F. Schaefer, III, P. v. R. Schleyer, and G. H. Robinson, Science321, 1069 (2008).

[32] A. Sidiropoulos, C. Jones, A. Stasch, S. Klein, and G.

Frenking, Angew. Chem. Int. Ed.48, 9701 (2009).

[33] A. Sidiropoulos, C. Jones, A. Stasch, S. Klein, and G.

Frenking, Angew. Chem.121, 9881 (2009).

[34] C. Jones, A. Sidiropoulos, N. Holzmann, G. Frenking, and A. Stasch, J. Chem. Soc. Chem. Commun.48, 9855 (2012).

[35] Y. Wang, Y. Xie, P. Wie, R. B. King, H. F. Schaefer, III, P. v. R. Schleyer, and G. Robinson, J. Am. Chem. Soc.

130, 14970 (2008).

[36] O. Back, G. Kuchenbeiser, B. Donnadieu, and G.

Bertrand, Angew. Chem. Int. Ed.48, 5530 (2009).

[37] O. Back, G. Kuchenbeiser, B. Donnadieu, and G.

Bertrand, Angew. Chem.121, 5638 (2009).

[38] C. A. Russell, Angew. Chem. Int. Ed.49, 9711 (2010).

[39] C. A. Russell, Angew. Chem.122, 9766 (2010).

[40] M. Y. Abraham, Y. Wang, Y. Xie, P. Wei, H. F. Schae- fer, III, P. v. R. Schleyer, and G. H. Robinson, Chem.

Eur. J.16, 432 (2010).

[41] D. J. D. Wilson, S. A. Couchman, and J. L. Dutton, In- org. Chem.51, 7657 (2012).

[42] M. Reinmuth, C. Neuhäuser, P. Walter, M. Enders, E.

Kaifer, and H.-J. Himmel, Eur. J. Inorg. Chem.1, 83 (2011).

[43] R. Appel and R. Schöllhorn, Angew. Chem. Int. Ed.3, 805 (1964).

[44] N. Holzmann, D. Dange, C. Jones, and G. Frenking, Angew. Chem. Int. Ed.52, 3004 (2013).

[45] R. Ahlrichs, M. Baer, M. Haeser, H. Horn, and C.

Koelmel, Chem. Phys. Lett.162, 165 (1989).

[46] A. D. Becke, Phys. Rev. A38, 3098 (1988).

[47] J. P. Perdew, Phys. Rev. B33, 8822 (1986).

[48] F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys.

7, 3297 (2005).

[49] K. Eichhorn, O. Treutler, H. Ohm, M. Häser, and R.

Ahlrichs, Chem. Phys. Lett.242, 652 (1995).

[50] J. G. Snijders, E. J. Baerends, and P. Vernoojs, At. Data Nucl. Data Tables26, 483 (1981).

[51] F. M. Bickelhaupt and E. J. Baerends, Rev. Comput.

Chem.15, 1 (2000).

[52] G. Te Velde, F. M. Bickelhaupt, E. J. Baerends, C. F.

Guerra, J. A. Van Gisbergen, J. Snijders, and T. Ziegler, J. Comp. Chem.22, 931 (2001).

[53] J. Krijn and E. J. Baerends, Fit Functions in the HFS- Method, Internal Report (in Dutch), Vrije Universiteit, Amsterdam 1984.

[54] E. Van Lenthe, E. J. Baerends, and J. G. Snijders, J.

Chem. Phys.99, 4597 (1993).

[55] K. Morokuma, J. Chem. Phys.55, 1236 (1971).

[56] T. Ziegler and A. Rauk, Inorg. Chem.18, 1755 (1979).

[57] G. Frenking, K. Wichmann, N. Fröhlich, C. Loschen, M. Lein, J. Frunzke, and M. Rayón, Coord. Chem. Rev.

55, 238 (2003).

[58] M. Lein and G. Frenking, Theory and Applications of Computational Chemistry: The First 40 Years, Eds.

C. E. Dykstra, G. Frenking, K. S. Kim, and G. E. Scuse- ria, Elsevier, Amsterdam 2005, p. 367–414.

[59] A. Krapp, F. M. Bickelhaupt, and G. Frenking, Chem.

Eur. J.12, 9196 (2006).

[60] G. Frenking and F. M. Bickelhaupt, The Chemical Bond: Fundamental Aspects of Chemical Bonding, Eds. G. Frenking and S. Shaik, Wiley-VCH, Weinheim 2014, p. 166–203.

[61] M. P. Mitoraj, A. Michalak, and T. Ziegler, J. Chem.

Theory Comput.5, 962 (200).

[62] M. P. Mitoraj and A. Michalak, Organometallics 26, 6576 (2007).

[63] M. P. Mitoraj and A. Michalak, J. Mol. Model.14, 681 (2008).

[64] M. P. Mitoraj, A. Michalak, and T. Ziegler, J. Phys.

Chem. A112, 1933 (2008).

[65] A. E. Reed, R. B. Weinstock, and F. Weinhold, J. Chem.

Phys.83, 735 (1985).

[66] Gaussian 09, Revision C.01, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R.

Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.

A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P.

Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L.

Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O.

Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E.

Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R.

Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N.

Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.

E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C.

Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannen- berg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Fores- man, J. V. Ortiz, J. Cioslowski, and D. J. Fox, Gaussian, Inc., Wallingford CT, 2010.

Referenzen

ÄHNLICHE DOKUMENTE

One cyano group of the counter anion completes the coordina- tion environment of the caesium atom (Cs-N 3.420(8) ˚ A), while the second one is involved in hydrogen bonding. *

We provide an overview on the relationships between P 1 and the polytope Q 1 associated with the two-layer crossing minimization problem with one xed layer, the linear

To the extent that the competition concentrates on the factors time and price, established manufacturers like IBM and Compaq lose market shares to these new

– If there are mutants found, then extend the multiplication history by an ar- ray of the number of elements of the same length as the new polynomials initialized by ones, multiply

Three earlier publications in the present series: Unequal in Life (1999), Community Adversity and Resilience (2004), and Dropping Off the Edge (2007), have charted the

The red-green government of Chancellor Gerhard Schröder enforced promotion of electricity produced from renewable energy sources and the gradual restriction of

Bottom left (hydrostatic vs. quasi-hydrostatic): difference between the hydrostatic, Boussinesq model and a model where some of the non-hydrostatic terms in the horizontal

We could not estimate the accuracy and precision of our dissolved oxygen data. We did not have enough equipment for an accurate measurement of dissolved oxygen. Since we