Jochen
Autschbach
a,
Bernd A.
Hess
a,
Mikael P.
Johansson
b,
Johannes
Neugebauer
a,
Michael
Patzschke
b,
Pekka
Pyykkö
ab,
Markus
Reiher
a and
Dage
Sundholm
b
aLehrstuhl für Theoretische Chemie, Universität Erlangen-Nürnberg, Egerlandstraße 3, D-91058 Erlangen, Germany. E-mail: Jochen.Autschbach@chemie.uni-erlangen.de; Johannes.Neugebauer@chemie.uni-erlangen.de; Markus.Reiher@chemie.uni-erlangen.de; Hess@chemie.uni-erlangen.de
bDepartment of Chemistry, University of Helsinki, P.O. Box 55 (A. I. Virtasen aukio 1), FIN-00014 Helsinki, Finland. E-mail: mpjohans@chem.helsinki.fi; michaelp@chem.helsinki.fi; pyykko@chem.helsinki.fi; sundholm@chem.helsinki.fi
First published on 5th November 2003
The icosahedral cluster-compound WAu12 was recently predicted by Pyykkö and Runeberg and experimentally prepared in the gas phase by the group of Lai-Sheng Wang. The photoelectron spectra and electron affinity were reported; the other physical properties remain unknown. Anticipating further experimental studies on it, we report here predicted vibrational spectra, NMR chemical shifts, spin–spin coupling constants and quadrupole coupling constants as well as optical spectra at the level of single and double excitations. The population analysis is non-trivial. By direct numerical integration, a charge of roughly +1 is obtained for the central tungsten atom. The charge distribution is strongly delocalised but bonding regions are clearly seen. A considerable electric field gradient exists at the gold nuclei. Although the radial bonds are strong, the system is quite elastic. The DFT activation energy for rotating one hemisphere against the other one, at a D5h transition state, is only about 20 kJ mol−1. The corresponding hu vibrational frequency is predicted to be slightly below 30 cm−1.
Apart from these data, and the first, fairly rough calculated vibrational frequencies,1 little is yet known about the properties of this system. Anticipating the preparation of macroscopic amounts of WAu12, we here provide spectroscopical data for interpretation of IR, Raman, UV/VIS, Mössbauer and NMR measurements. A faithful characterization of the electronic charge distribution turned out to be non-trivial. The simplest population analysis methods failed abysmally. Finally we emphasize the fluxionality of WAu12; the lower vibrational modes and hindered rotations can be easily thermally activated.
Two different types of density functionals were employed in the study. The Becke–Perdew functional, BP86,9,10 is a pure GGA functional, whereas Becke's three-parameter functional for exchange in combination with the LYP correlation functional,11–14 B3LYP, is a hybrid functional, containing a portion of exact exchange. In connection with the BP86 functional we used the resolution-of-the-identity (RI) technique15,16 to speed up the calculations where possible.
Apart from the core electrons substituted by the effective core potential, no additional electrons were frozen in the second-order Møller-Plesset (MP2) calculations. The coupled-cluster calculations and coupled-cluster response calculations at the CC2 level17,18 were performed with the TURBOMOLE program package; CC2 is a second-order approximate, coupled-cluster singles and doubles (CCSD) model. In these calculations, the 12 lowest lying electron pairs were kept uncorrelated. The density fitting RI approach was also used in connection with the Møller–Plesset (MP2) and coupled cluster (CC) calculations.19–22
Ahlrichs' Gaussian TZVP and TZVPP basis sets,23 featuring a valence triple-zeta basis set with polarisation functions on all atoms, were used throughout in the Turbomole calculations. The replacement of the single Au f function with exponent 1.056 by two primitive f functions with exponents 1.19 and 0.2 follows Pyykkö et al.24 In the case of the RIMP2/TZVPP(2f) calculations, the def-TZVPP auxiliary basis set25,4 was used for the resolution of the identity. A VQZPP basis set was obtained by augmenting the TZVPP basis set with the additional functions given in Table 1.
Type | W | Au |
---|---|---|
s | 0.01508098701 | 0.01721719738 |
p | 0.00673701583 | 0.00900421289 |
d | 0.03752146174 | 0.05643075118 |
f | 2.06784897500 | 1.19000000000 |
0.33085583600 | 0.43652508514 | |
0.16012953778 |
Further calculations were done with the Amsterdam Density Functional (ADF) package. Mainly the BP86 functional was used, but for control calculations, also other functionals were employed. Relativistic effects were included by means of the zeroth-order regular approximation (ZORA).26 The Pauli approximation, which is also included in ADF, was found to give very unreliable results. The core electrons were treated as frozen in the population analyses and electron-localisation-function (ELF) calculations. A large core, 60 electrons for gold and tungsten, was used, except for the oxidation-state calculations (see section 3.5), where a smaller core with 46 electrons was used. The 60-electron core includes the 4f electrons of gold and tungsten. The Slater basis set used was the ‘ZORA-TZP’ one.27 In the ADF calculations, the importance of f-type polarisation functions was found to be small. The bond lengths obtained using the standard TZV2P basis set27 which includes f-type polarisation functions is less than 1 pm shorter than the bond distances obtained with the TZP basis set.
Sk![]() ![]() ![]() ![]() | (2.1) |
![]() | (2.2) |
![]() | (2.3) |
![]() | (2.4) |
The charge-density distribution was also studied by explicit numerical integration. The spherical integration of the electron density was done as described in refs. 33 and 34. For the non-spherical integration the program basin35 was used.
Three different methods for partitioning of the electrons between the atoms were employed: generalized atomic polar tensor (GAPT), shared electron number (SEN) and Voronoi charges. The GAPT charges are based on atomic polar tensors36 and are calculated from the gradients of the dipole moment components. In the SEN approach,37,38 partial charges are calculated from modified atomic orbitals in such a way that the unassigned charge in the molecule is minimized. Finally, the Voronoi partitioning works by integrating the electron density over the Voronoi cell fragments of the molecule. The Voronoi cell of an atom is defined as the region of space that is closer to that atom than to any other atom.39
Method | R(W–Au) | R(Au–Au) |
---|---|---|
a Agrees with the RIMP2/TZVPP![]() ![]() |
||
RIMP2/TZVP | 279.8 | 294.2 |
RIMP2/TZVPP | 270.6 | 284.5 |
RIMP2/TZVPP(2f) | 268.0a | 281.8 |
RICC2/TZVPP | 269.9 | 283.8 |
BP86/RI/TZVP | 279.8 | 294.2 |
BP86/TZVPP | 278.6 | 292.9 |
BP86/TZVPP(2f) | 276.1 | 290.3 |
BP86/VQZPP | 273.9 | 288.0 |
B3LYP/TZVP | 283.0 | 297.6 |
B3LYP/TZVPP | 282.0 | 296.5 |
B3LYP/TZVPP(2f) | 279.8 | 294.2 |
B3LYP/VQZPP | 278.0 | 292.3 |
The distances for WAg12 are shown in Table 3. They actually are closely similar to the corresponding values for WAu12. Further test calculations also show similar bond distances for diatomic Ag and Au systems, which are presented in Table 4. The MP2 results are seen to be closer to experiment for these diatomics.
Method | R(W–Ag) | R(Ag–Ag) |
---|---|---|
RIMP2/TZVPP | 271.1 | 285.0 |
BP86/RI/TZVP | 277.6 | 291.9 |
BP86/TZVPP | 277.2 | 291.5 |
B3LYP/TZVPP | 281.4 | 295.9 |
Basis set | No. | Irrep |
![]() |
![]() |
![]() |
A RIMP2 | A BP86 | A B3LYP | ⊥ σ u,BP86 | ⊥ σ u,B3LYP |
---|---|---|---|---|---|---|---|---|---|---|
TZVP | 1 | hu | 36.6 | 28.9 | 28.0 | — | — | — | — | — |
2 | fg | 78.8 | 45.9 | 43.7 | — | — | — | — | — | |
3 | hg | 73.9 | 68.1 | 65.4 | — | — | — | 7674 | 8054 | |
4 | t1u | 77.5 | 72.2 | 72.5 | 0.116 | 0.480 | 1.719 | — | — | |
5 | t2u | 83.5 | 81.4 | 77.1 | — | — | — | — | — | |
6 | fu | 89.0 | 88.3 | 84.7 | — | — | — | — | — | |
7 | hg | 131.8 | 110.6 | 104.0 | — | — | — | 2307 | 2706 | |
8 | ag | 125.8 | 131.1 | 124.5 | — | — | — | 2784 | 3165 | |
9 | t1u | 186.3 | 177.0 | 170.0 | 1.070 | 2.788 | 0.197 | — | — | |
TZVPP | 1 | hu | 51.4 | 29.5 | 26.4 | — | — | — | — | — |
2 | fg | 81.6 | 46.8 | 42.0 | — | — | — | — | — | |
3 | hg | 86.1 | 68.7 | 64.5 | — | — | — | 7505 | 7909 | |
4 | t1u | 90.3 | 72.9 | 71.5 | 0.005 | 0.431 | 1.695 | — | — | |
5 | t2u | 92.6 | 81.0 | 75.8 | — | — | — | — | — | |
6 | fu | 118.9 | 89.7 | 83.2 | — | — | — | — | — | |
7 | hg | 141.8 | 111.4 | 102.7 | — | — | — | 2109 | 2655 | |
8 | ag | 150.7 | 131.7 | 123.5 | — | — | — | 2552 | 2943 | |
9 | t1u | 201.7 | 175.3 | 168.0 | 0.118 | 2.900 | 0.219 | — | — |
Basis set | No. | Irrep |
![]() |
![]() |
A BP86 | A B3LYP | ⊥ σ u,BP86 | ⊥ σ u,B3LYP |
---|---|---|---|---|---|---|---|---|
TZVPP(2f) | 1 | hu | 31.9 | 28.0 | — | — | — | — |
2 | fg | 48.8 | 43.2 | — | — | — | — | |
3 | hg | 71.0 | 65.9 | — | — | 6728 | 7218 | |
4 | t1u | 73.7 | 71.6 | 0.541 | 1.803 | — | — | |
5 | t2u | 83.9 | 77.5 | — | — | — | — | |
6 | fu | 91.3 | 84.3 | — | — | — | — | |
7 | hg | 113.1 | 104.0 | — | — | 2374 | 2897 | |
8 | ag | 133.7 | 125.3 | — | — | 2397 | 2702 | |
9 | t1u | 176.7 | 167.9 | 4.693 | 0.959 | — | — | |
VQZPP(2f) | 1 | hu | 28.2 | 29.3 | — | — | — | — |
2 | fg | 45.6 | 45.9 | — | — | — | — | |
3 | hg | 70.9 | 67.7 | — | — | 5202 | 6058 | |
4 | t1u | 72.2 | 73.1 | 0.677 | 1.809 | — | — | |
5 | t2u | 87.1 | 79.5 | — | — | — | — | |
6 | fu | 88.2 | 87.3 | — | — | — | — | |
7 | hg | 113.8 | 107.5 | — | — | 2971 | 2997 | |
8 | ag | 133.7 | 128.9 | — | — | 2290 | 2493 | |
9 | t1u | 181.8 | 170.9 | 4.623 | 0.967 | — | — |
The character of certain vibrations of WAu12 is shown in Fig. 1. Only the two t1u modes are IR active. Of them the highest, 9t1u, looks like a particle inside a box, vibrating against it. Both the single, breathing ag, and the two hg vibrations are Raman active.
![]() | ||
Fig. 1 Normal modes of the IR and Raman active vibrations and of the (silent) lowest frequency mode. |
The main character of the lowest, hu vibration is having the two hemispheres rotating against each other. Note that there are six five-fold axes while hu is five-dimensional. The D5h transition state from Ih back to Ih will be discussed below.
As can be understood from Tables 5 and 6, the Raman intensities show a small dependence on the size of the basis set, which is, for instance, discussed in ref. 28. This is not the case for IR intensities, which suffer a larger change, in particular, for MP2 calculations. In general, the reason for this is the strong basis set dependence of the small dipole moment generated upon distortion of the equilibrium structure, which has no dipole moment due to inversion symmetry. In particular, the results from ab initio MP2 calculations depend much more strongly on the size of the basis set than DFT results do.
![]() | ||
Fig. 2 The D5h transition state (left) and the Oh (right) stationary point of WAu12. |
Frequency analyses using density functional methods suggest that the Oh structure is a minimum (see Table 7), whereas RIMP2 predicts a fourth-order saddle point. The accuracy of the numerical procedure was tested for the critical mode and found to be much higher than the size of imaginary frequency. The D5h structure is a transition state between two equivalent Ih structures (Table 8) and roughly 20 kJ mol−1 above them at DFT level (Table 9). The RIMP2 value is higher, over 70 kJ mol−1.
No. | Irrep |
![]() |
![]() |
![]() |
![]() |
---|---|---|---|---|---|
1 | a2g | 14.50 | i 51.89 | 13.02 | 12.25 |
2 | t2u | 23.13 | i 32.35 | 22.30 | 22.25 |
3 | eg | 47.28 | 55.45 | 46.88 | 45.49 |
4 | eu | 50.83 | 63.32 | 50.94 | 46.37 |
5 | t2g | 75.64 | 87.85 | 75.52 | 70.99 |
6 | t2u | 77.10 | 78.44 | 76.56 | 71.71 |
7 | t1u | 80.56 | 97.21 | 80.38 | 78.70 |
8 | t1g | 82.54 | 99.53 | 82.93 | 75.84 |
9 | eg | 100.02 | 130.54 | 100.06 | 92.50 |
10 | t1u | 102.99 | 120.81 | 103.09 | 96.26 |
11 | t2g | 127.33 | 145.00 | 127.47 | 119.14 |
12 | a1g | 129.97 | 158.29 | 130.43 | 122.19 |
13 | a2u | 135.02 | 155.16 | 135.30 | 126.85 |
14 | t1u | 162.49 | 192.11 | 161.28 | 153.71 |
No. | Irrep |
![]() |
![]() |
![]() |
![]() |
---|---|---|---|---|---|
1 | a″1 | 8.66i | 23.59i | 8.48i | 8.09i |
2 | e″2 | 25.81 | 35.80 | 25.16 | 25.00 |
3 | e″1 | 28.91 | 38.82 | 29.10 | 28.10 |
4 | e′1 | 46.44 | 67.82 | 46.67 | 43.47 |
5 | e″2 | 60.10 | 65.69 | 60.21 | 56.84 |
6 | e′2 | 66.00 | 77.91 | 66.04 | 63.01 |
7 | e″1 | 66.57 | 75.42 | 66.76 | 62.91 |
8 | a′1 | 74.53 | 84.72 | 74.23 | 69.96 |
9 | a″2 | 75.05 | 76.22 | 74.94 | 70.45 |
10 | a″2 | 83.76 | 92.05 | 82.22 | 80.63 |
11 | e′1 | 83.89 | 94.12 | 83.79 | 83.05 |
12 | e″2 | 84.70 | 109.34 | 85.50 | 78.63 |
13 | e′2 | 89.43 | 101.18 | 90.21 | 84.38 |
14 | e″1 | 112.99 | 129.75 | 113.25 | 106.30 |
15 | e′2 | 115.73 | 138.19 | 114.29 | 108.19 |
16 | e′1 | 119.13 | 137.07 | 118.51 | 112.30 |
17 | a′1 | 124.29 | 137.11 | 123.31 | 116.40 |
18 | a′1 | 129.18 | 149.09 | 128.44 | 121.59 |
19 | e′1 | 162.25 | 185.67 | 160.41 | 154.22 |
20 | a″2 | 165.23 | 185.89 | 163.82 | 157.52 |
Method | (A) | (B) | (C) | (D) | (E) |
---|---|---|---|---|---|
RIMP2/TZVPP | 97.2 | 70.8 | 160.8 | 54.1 | 58.9 |
BP86/RI/TZVP | 49.5 | 20.9 | 10.6 | −5.6 | −10.3 |
BP86/TZVPP | 48.3 | 20.6 | 10.1 | −6.1 | −10.6 |
B3LYP/TZVPP | 55.1 | 19.3 | 9.8 | 2.7 | −3.8 |
B3LYP(0.30)/TZVPP | 60.4 | 23.2 | 12.8 | 9.3 | 2.4 |
B3LYP(0.50)/TZVPP | 69.5 | 31.0 | 14.2 | 20.0 | 13.1 |
According to transition state theory, an approximation for the rate constant for the rotation of the icosahedral structure to another icosahedral structure via the D5h symmetric transition state is given by
![]() | (3.5) |
![]() | (3.6) |
![]() | ||
Fig. 3 Rate constant k
(given as log10(k)) as a function of the inverse temperature for the transition Ih-WAu12![]() ![]() |
Method | ΔG‡ | k |
---|---|---|
BP86/RI/TZVP | −28.09 | 7.45![]() ![]() |
BP86/TZVPP | −26.84 | 12.32![]() ![]() |
B3LYP/TZVPP | −27.74 | 8.60![]() ![]() |
RIMP2/TZVPP | −65.90 | 1.77![]() ![]() |
Searching for a simpler model for fluxional motions of small gold clusters, we calculated the barrier for the distortion of tetrahedral50 Au42+ towards a D4h-symmetric transition structure. The data are included in Table 9. No low barriers are found in this particular case. For the Au135+, the D5h and Oh structures actually lie within a few kJ mol−1 from the Ih.
We furthermore investigated the influence of the exact exchange contribution to all relative energies by setting the exact exchange in the B3LYP functional to 30% and 50%, since, for certain reactions,51,52 the hybrid density functionals with a higher amount of exact exchange lead to improved descriptions of reaction barriers. In our case, the effect is 3 to 4 kJ mol−1 for 30% exact exchange and 4 to 12 kJ mol−1 for 50% exact exchange in case of the WAu12 barriers, while the changes are larger for the highly charged Au135+ system (6 to 7 kJ mol−1 for 30% exact exchange; about 17 kJ mol−1 for 50% exact exchange). However, the results obtained with an increased percentage of exact exchange are still more in the range of the other DFT data than to those from RIMP2.
The fluxionality of these icosahedral clusters may enhance their eventual catalytical activity. Such an increase was recently noticed for the adsorption and activation of O2 on gold clusters.53 Similarly, Oviedo and Palmer54 found a large number of isomers for Au13. The group of Kappes55 observed experimentally for Au9+ two isomers below 140 K, corresponding to an activation energy of 10–20 kJ mol−1.
A further, intriguing phenomenon that may or may not have a qualitative connection with the present fluxionality is the non-existence of frozen interstitial atoms in gold, down to 0.3 K.56 For copper and silver they exist, and are activated at 38 and 28 K, respectively.57 Sakai et al.58 measured for the grain-boundary activation energy of nanocrystalline gold a value of 0.2 eV or 19 kJ mol−1. Furthermore, a surface reorganization takes place for metallic gold. For all these aspects related to the general ‘softness’ of gold, see a review by Pyykkö.59
We note that all 197Au species with closed electron shells are identical heavy fermions. How low would the barriers have to be, for appreciable tunneling to take place, a process unlikely for a particle as heavy as the gold atom? Consider a one-dimensional potential step of height V and width 2a and let the gold atom, of mass m, vibrate against it with a frequency ν, near the bottom of the barrier. Then the average tunneling frequency becomes
![]() | (3.7) |
f![]() ![]() ![]() ![]() | (3.8) |
We conclude from the corresponding Fig. 4 that the barriers would have to be below 0.2 kJ mol−1 for appreciable tunneling of a single gold atom to take place.
![]() | ||
Fig. 4 Average tunneling frequency of a gold atom, hitting a barrier of height V and width 2a![]() ![]() ![]() ![]() |
![]() | (3.9) |
![]() | (3.10) |
No. | Irrep |
![]() |
![]() |
![]() |
A RIMP2 | A BP86 | A B3LYP | ⊥ σ u,BP86 | ⊥ σ u,B3LYP |
---|---|---|---|---|---|---|---|---|---|
1 | hu | 57.88 | 50.26 | 47.00 | — | — | — | — | — |
2 | fg | 87.61 | 68.62 | 63.43 | — | — | — | — | — |
3 | hg | 99.10 | 86.73 | 82.02 | — | — | — | 5113 | 5840 |
4 | t1u | 89.86 | 91.49 | 89.52 | 0.033 | 2.525 | 4.107 | — | — |
5 | fu | 111.72 | 105.56 | 99.24 | — | — | — | — | — |
6 | t2u | 118.19 | 111.42 | 104.35 | — | — | — | — | — |
7 | hg | 158.62 | 133.29 | 123.48 | — | — | — | 3853 | 4397 |
8 | ag | 165.78 | 155.86 | 145.67 | — | — | — | 4334 | 5546 |
9 | t1u | 219.26 | 197.30 | 186.18 | 21.103 | 2.599 | 6.401 | — | — |
No. | Irrep | k BP86(WAu12) | k BP86(WAg12) | μ BP86(WAu12) | μ BP86(WAg12) |
---|---|---|---|---|---|
1 | hu | 6.5 | 10.2 | 196.97 | 106.91 |
2 | fg | 16.3 | 19.1 | 196.97 | 106.91 |
3 | hg | 35.2 | 30.4 | 196.97 | 106.91 |
4 | t1u | 38.7 | 39.5 | 192.37 | 124.64 |
5 | t2u | 48.9 | 50.2 | 196.97 | 106.91 |
6 | fu | 60.0 | 45.1 | 196.97 | 106.91 |
7 | hg | 92.4 | 71.9 | 196.97 | 106.91 |
8 | ag | 129.3 | 98.3 | 196.97 | 106.91 |
9 | t1u | 219.8 | 203.0 | 189.08 | 137.77 |
To complete the picture, Voronoi charges were calculated. The Voronoi partitioning is based on integration of the electron density over the Voronoi cell fragments that define the molecule.39 This method gives a charge of +0.91 for tungsten, using the BP86 functional with the TZVP basis. It has been suggested, that one should not use the Voronoi charges as such, but rather the change in charge from the atomic fragments to the molecule.60 By using this so called Voronoi deformation density, a charge of +0.075 is obtained for tungsten.
All the population analyses above give, however, only an indirect picture of the charge distribution in WAu12, and molecules in general. Their value comes more from giving trends in changes to atomic charges upon various perturbations of the system, not in providing absolute values as such.
For a highly symmetrical, nearly spherical species as WAu12, there is a more direct alternative to population analyses, for extracting the charge of the central atom. By doing a direct integration of the electron density around W, and assuming a spherical atom, a good picture of the charge localisation can be obtained.33,34 A somewhat arbitrary parameter in this case will be the radius of the tungsten atom. As seen from the integration curve in Fig. 5, any reasonable choice for the radius will yield a positive charge for tungsten. Two different radii have been selected for each method: The inflection point of the density accumulation, and half of the bond length. The inflection point, where the increase in electron density goes through a minimum, occurs for all electronic-structure methods at approximately 125 pm. The charges for W, obtained with these radii, are tabulated in Table 14. The tungsten is clearly positive.
![]() | ||
Fig. 5 The charge of the tungsten atom as a function of its radius. Optimized structures for the respective methods have been used. |
Method | 125 pm | Midpoint |
---|---|---|
RI-CC2/TZVPP | 2.28 | 1.59 (135 pm) |
BP86/TZVPP | 2.71 | 1.80 (139 pm) |
B3LYP/TZVPP | 2.76 | 1.79 (141 pm) |
B3LYP/VQZPP | 2.76 | 1.87 (139 pm) |
Both BP86 and B3LYP give very similar results. The most positive charge at a given radius is given by the B3LYP/TZVPP combination, which also has the longest W–Au distance. The closer to tungsten the gold atoms are, the more electron density they donate to the inner regions. The result for the coupled cluster RI-CC2 method is slightly different, implying a clearly less positive charge for the tungsten. For all quantum-chemical methods, to get even a neutral tungsten, the radius would have to be very large.
The assumption of a spherical central atom is naturally only an approximation. The electron cloud of tungsten is deformed by the gold atoms. This leads to a dodecahedral charge distribution and thereby to too small a number of electrons for tungsten. To correct for this, one can calculate the volume difference between a sphere and its inscribed dodecahedron. Then one can integrate over a thin shell to get the electron density in this region. Doing this at the point of inflection (125 pm) shows that approximately 0.5 electrons exist between the sphere and the polyhedron (BP86/TZVP). This gives a charge for tungsten of about +1.4 compared to +1.9 for the spherical integration. A direct integration of the dodecahedron was also performed. For this we determined the minimum of the gradient of the electron density. This defines a basin for tungsten that has the shape of a dodecahedron. Integrating the electron density in this basin also gives a charge of +1.4 for tungsten (BP86/TZVP).
Another chemically useful, but somewhat fluid, concept is the oxidation state of atoms in molecules. For gold, these have been determined using electron spectroscopy for chemical analysis (ESCA).61 The energy of the 4f7/2 orbital plotted against the oxidation state of gold in different compounds gives approximately a straight line. The oxidation state of a new compound can be obtained from measured 4f7/2 orbital energies. Here we applied this method in a computational way. By calculating the energy of the 4f7/2 orbital for a few gold and tungsten compounds, whose oxidation states are known, we constructed straight lines representing the 4f7/2 energies of gold and tungsten as a function of the oxidation state of the metal. The uncertainty in the fit is unfortunately a bit higher than desired, as can be seen in Figs. 6 and 7. Even so, the results suggest an oxidation state of about +1.5 for W and +0.7 for Au. Also this analysis points to a more positive tungsten. Positive oxidation states for all atoms in the system might seem somewhat contradictory for a chemist, while in metal physics it is common to regard for instance all atoms of metallic Eu or La as divalent and trivalent, respectively. Such a positive oxidation state of both tungsten and gold atoms can be given an explanation. The electron density difference between the WAu12 molecule, and its constituent atoms shows how the electrons move upon formation of the cluster. The integrated electron density differences spheres around the tungsten and gold atoms are shown in Fig. 8. Near the atoms, the electron density difference between WAu12 and the free atoms is negative, i.e., both W and Au in WAu12 are more positively charged than the neutral atoms. The electron density close to the nucleus probably affects the 4f7/2 orbital energy most, as these orbitals are quite close to the nuclei themselves. The fact that the ESCA analysis gives positive oxidation states for all atoms is not, after all, contradictory.
![]() | ||
Fig. 6 Plot of the 4f7/2 electron energy versus the oxidation state of gold in different compounds. Calculated with BP86/TZVP using ADF. The arrow marks the value for gold in WAu12. |
![]() | ||
Fig. 7 Plot of the 4f7/2 electron energy versus the oxidation state of tungsten in different compounds. Calculated with BP86/TZVP using ADF. The arrow marks the value for tungsten in WAu12. |
![]() | ||
Fig. 8 The integrated electron density difference between the WAu12 molecule and the free atoms it is built from. Curves based on spheres around tungsten (big picture) and gold (inset) are shown. The density difference in it self is indicated by the whole line, its derivative as a dashed line. For clarity, the areas with a negative derivative are painted red (lower part), whereas areas with a positive derivate are blue (upper part). Thus, upon formation of the molecule, electrons move from the “red” areas to the “blue”. The vertical line at 282 pm shows the position of the gold atoms. Curve based on B3LYP/TZVPP-level calculations. |
Fig. 8 also shows the differentiated changes in electron density (dashed lines). Areas with a negative derivative (shown in red) suffer a loss of electron density upon formation of the molecule, whereas areas with a positive derivative (in blue) gain electrons. A very large electron gathering between the tungsten and the gold atoms can be seen: A bond is formed.
Other features of the electron distribution also deserve commenting. As seen in Fig. 8, the derivative function has a long negative tail outside the WAu12 molecule; at large distances, the electron cloud around the gold atoms is less diffuse compared to the electron density of the free atoms. In WAu12, electrons are pulled inwards towards the centre. However, just outside the gold atoms, there is a sphere of electron gain. This reveals presence of a delocalised electron cloud, and might be an indication of molecular aromaticity. In fact, the nuclear magnetic shielding calculations show a quite strong long-range magnetic shielding effect where no electron density is present. The long-range magnetic shielding function has been used for estimating the strength of the magnetically induced currents which can be used as a measure of the degree of molecular aromaticity.62 Thus the magnetic shielding calculations also suggest that WAu12 is aromatic.
Another method for looking at the bonding situation in WAu12, is to calculate the electron-localisation function (ELF).30 An isosurface of the ELF for the whole system is shown in Fig. 9. As seen in that figure, the gold atoms are far from spherical and there is an area of increased localised electrons between the tungsten and gold atoms. Fig. 10 shows a cut plane of the ELF curves for WAu12. In this plane, the gold ligands look nearly triangular and between the gold and tungsten atoms one sees an area with high ELF values showing the covalent part of the gold–tungsten bond. The W–Au bond could be thought to lead to a local electron maximum between the atoms. Actually such a maximum rather seems to be pushed to the outer part of the cluster, leading to the peculiar shape of the gold atoms. This increase of electron density outside the cluster is also seen in Fig. 8. A comparison of the ELF plots in Fig. 10 with a corresponding ELF plot (not shown) for a neutral tungsten atom reveals the large size of the free tungsten atom and explains some of the difficulties of the electron partitioning schemes in assigning partial charges to tungsten and gold. Because the aurophilic attraction is predominantly of the dispersion type, it will not necessarily give any strong density deformations.
![]() | ||
Fig. 9 Plot of an isosurface of the ELF of WAu12. |
![]() | ||
Fig. 10 Cut plane through the ELF of WAu12. |
The proposed 18-electron bonding1 was confirmed by different analyses. Table 15 shows contributions of the 5d, 6s and 6p orbitals of tungsten to the molecular orbitals. Also, looking at the radial distributions of orbitals of ag, t1u, and hg symmetry, corresponding to the s, p, and d symmetries of W, shows their participation in bonding. Fig. 11 shows the normalised orbital density accumulation for selected, especially strong bonding orbitals.
![]() | ||
Fig. 11 The radial integration of the norm for various molecular orbitals of WAu12. |
E/eV | MO No. | MO | % | AO (W) | E/eV | Occ |
---|---|---|---|---|---|---|
−12.975 | 88 | 12a1g | 25.25% | 6s | −6.549 | 2.0 |
−11.127 | 89 | 14e2g | 28.53% | 5dx2−y2,xy | −5.495 | 0.8 |
−11.127 | 90 | 13a1g | 28.53% | 5dz2 | −5.495 | 0.8 |
−11.126 | 91 | 14e1g | 28.53% | 5dyz,xz | −5.495 | 0.8 |
−10.892 | 92 | 16e1u | 9.34% | 6px,y | −1.288 | 0.0 |
−10.891 | 93 | 13a2u | 9.34% | 6pz | −1.288 | 0.0 |
−10.288 | 94 | 14a1g | 14.86% | 7s | 0.736 | 0.0 |
−9.254 | 103 | 16e2g | 2.84% | 5dx2−y2,xy | −5.495 | 0.8 |
−9.254 | 104 | 15a1g | 2.84% | 5dz2 | −5.495 | 0.8 |
−9.253 | 105 | 16e1g | 2.83% | 5dyz,xz | −5.495 | 0.8 |
−8.704 | 109 | 17e1g | 2.66% | 5dyz,xz | −5.495 | 0.8 |
−8.704 | 110 | 17e2g | 2.66% | 5dx2−y2,xy | −5.495 | 0.8 |
−8.704 | 111 | 16a1g | 2.65% | 5dz2 | −5.495 | 0.8 |
−7.449 | 123 | 22e1u | 2.69% | 6px, y | −1.288 | 0.0 |
−7.448 | 124 | 17a2u | 2.68% | 6pz | −1.288 | 0.0 |
−6.252 | 127 | 17a1g | 21.02% | 5dz2 | −5.495 | 0.8 |
−6.252 | 128 | 20e2g | 21.03% | 5dx2−y2,xy | −5.495 | 0.8 |
−6.251 | 129 | 20e1g | 21.04% | 5dyz, xz | −5.495 | 0.8 |
HOMO | ||||||
LUMO | ||||||
−4.464 | 130 | 18a1g | 46.74% | 5dz2 | −5.495 | 0.8 |
−4.464 | 131 | 21e2g | 46.74% | 5dx2−y2,xy | −5.495 | 0.8 |
−4.464 | 132 | 21e1g | 46.71% | 5dyz,xz | −5.495 | 0.8 |
−4.205 | 133 | 19a1g | 18.49% | 6s | −6.549 | 2.0 |
−4.205 | 133 | 19a1g | 10.01% | 7s | 0.736 | 0.0 |
−2.140 | 136 | 19a2u | 22.68% | 6pz | −1.288 | 0.0 |
−2.140 | 137 | 23e1u | 22.66% | 6px,y | −1.288 | 0.0 |
Finally, the shared electron number (SEN) analysis shows a very pronounced multicentre bonding in the molecule. At B3LYP/VQZPP-level, the W–Au pairs share as many as 4.1 electrons. The SEN between two adjacent golds is also high, 2.2. Furthermore, the three-centre analysis for the triangles W–Au–Au and Au–Au–Au gives SEN's of 1.8 and 1.3, respectively. Even for the four-centre trigonal pyramid, W–Au3, the SEN is 1.2. The SEN analysis also further corroborates the 18-electron binding; the only positive contributors to the shared 4.1 electrons between W and the Au's are the ag, t1u, and hg orbitals.
The presented bonding analyses allow us to comment on the way of writing the molecular formula as W@Au12.2 This notation implies a tungsten atom trapped inside an Au12 cage. As clear bonding between the tungsten and gold atoms is found to exist, the W@Au12 mental picture for the WAu12 cluster is not really correct. The Au12 cage does not exist as such but gets heavily distorted if the tungsten is removed. Hence the tungsten is not an addition to an otherwise stable molecule, in the same sense as in a M@C60 system, which does have a strongly bound cage.
Much stronger absorption lines are found in the ultraviolet region. Both B3LYP and BP86 show a high intensity peak at about 250 nm. The overall correspondence between B3LYP and BP86 is, however, not very high. B3LYP gives quite similar spectra with both the TZVPP and VQZPP basis sets. BP86, on the other hand, seems to still be far from the basis set limit at TZVPP level; the VQZPP results differ significantly in the higher energy region of the spectra. Fig. 12 shows the spectra as calculated with various methods. Two of the lowest energy excitations calculated with various methods are compared in Table 16.
![]() | ||
Fig. 12 Calculated UV/VIS spectra for WAu12. The spectra from CC2/TZVPP (full line), B3LYP/VQZPP (dashed line) and BP86/VQZPP (dotted line) are compared. For RI-CC2, only excitations above 272 nm were obtained. All calculations used optimized structures. |
Method | exc 1 | exc 2 |
---|---|---|
CCS/TZVPP | 284 (0.01) | 267 (0.12) |
CC2/TZVPP | 442 (0.02) | 381 (0.18) |
BP86/TZVPP | 466 (0.05) | 408 (0.10) |
BP86/VQZPP | 455 (0.06) | 401 (0.09) |
B3LYP/TZVPP | 420 (0.11) | 346 (0.05) |
B3LYP/VQZPP | 418 (0.11) | 345 (0.03) |
It is worth noting that the CCS calculation gives much too energetic excitations; the double excitation character accounted for by CC2 lowers the energy considerably. For both CC2 peaks, the doubles contribution is over 18%. Such a high doubles contribution indicates that the uncertainties in the CC2 excitation energies can be significant.
Finally we have verified that the vertical excitation energy from the singlet ground state to a triplet excited state is 1.68 eV at B3LYP/VQZPP level. The triplet excited states were also considered by Li et al.2
183W has a natural abundance of 14.31%, I=
½ and a nuclear magnetic moment of 0.11778 μN. It is a frequently utilized NMR nucleus. The results of the NMR computations are listed in Table 17.
Property | rel.a | nrel.b |
---|---|---|
a Relativistic ZORA calculation with ZORA-TZP basis, including spin–orbit coupling unless otherwise noted.
b Nonrelativistic calculation with nonrelativistic TZP basis.
c ZORA-TZP+s basis for W (shieldings), or W and Au (coupling constants). See Sec. 2.5 for details
d Scalar relativistic ZORA computation.
e Corresponding to (5 1K![]() ![]() ![]() ![]() |
||
σ(183)W | 15![]() |
13![]() |
16![]() |
||
δ(183)W | −13![]() |
−12![]() |
−13![]() |
||
K(WAu) | 7350cd | 3322 |
1 K(AuAu) | 11![]() |
6000 |
2 K(AuAu) | 15![]() |
329 |
3 K(AuAu) | 693cd | −4596 |
K av(AuAu) | 12![]() |
2459e |
NQCC (197Au) | 549c | 275 |
482cd |
Tungsten chemical shifts have previously been calculated with the ADF ZORA NMR method and the TZP basis by Ziegler et al.47 They have obtained a good overall agreement with experimental data, with a mean absolute deviation of approximately 3% of the investigated 183W chemical shift range of 7200 ppm. Spin–orbit contributions were found to be of minor importance for 183W chemical shifts in ref. 47.
As to the chemical shifts, compared to the reference system WO42−, the 183W nucleus in WAu12 is very strongly shielded. The computations thus predict an extraordinarily large chemical shift for W in WAu12 of more than −13000 ppm. For comparison, the largest chemical shifts that have been calculated by Ziegler et al. in ref. 47 and compared to experiment were 3638 ppm for WS42− and −3876 ppm for W(CO)6, respectively, with respect to WO42−. A comparison with the nonrelativistic results demonstrates that the chemical shift in WAu12 is only to a comparatively small extent due to relativistic effects. It results from a large positive paramagnetic shielding contribution of 3243 ppm as compared to the reference (WO42−: −8088 ppm). In addition, the spin–orbit coupling contributions of 3850 ppm to the 183W shielding in WAu12 are also significantly larger than for the reference (1937 ppm). Thus the spin–orbit contributions to the chemical shift here exceed the ones reported by Ziegler et al.47 by a large margin. The diamagnetic shielding contributions for WAu12 and the reference are very similar (8649 ppm for WAu12 as compared to 8613 ppm for WO42−). It appears that the Au cage quite effectively shields the external magnetic field.
When converted to frequency units, the spin–spin coupling constants are quite small, reflecting the smallness of the W and Au nuclear magnetic moments, and the small s character of the bonds, compared with Hg–Hg bonds, see for the latter for instance ref. 64. The calculated J(WAu) is 65 Hz and the averaged J(AuAu) over all gold positions, corresponding to a rapid intramolecular hindered rotation, is 45 Hz. A comparison with the nonrelativistic data shows that the magnitudes of the W–Au as well as the Au–Au reduced coupling constants exhibit large positive contributions due to relativistic effects, in particular for 2K(AuAu). However, already at the nonrelativistic level large magnitudes for K(WAu) and 1K(AuAu) and 3K(AuAu) are found. The relativistic increase of the coupling constants is comparatively small for a coupling constant between two heavy nuclei. The case of WAu12 here appears to be similar to Pt–Pt and Pt–ligand coupling constants previously investigated in refs. 43,65,66. For these coupling constants typically a sizeable, but not huge, relativistic increase is obtained. Its magnitude depends sensitively on the chemical environment of the coupled nuclei and is balanced by relativistic effects on the hyperfine integrals, on the chemical bond between the coupled atoms, and on their interaction with other neighboring atoms. Interestingly, the two-bond Au–Au coupling constant is predicted to be larger than the one-bond coupling constant. They both exceed the three-bond coupling by an order of magnitude in the relativistic case.
The calculated NQCC values for the 197Au nuclear ground state are of the order of 0.5 GHz. This is a typical order of magnitude for gold compounds. Note that the quadrupole splittings will then exceed the Zeeman splittings in typical laboratory fields. More specifically, in pure gold clusters, the Mössbauer quadrupole splittings of unbound surface atoms and surface atoms bound to trimethyl phosphines are observed to be67 1.40 and 3.5 mm/s, which translates to 87 and 218 MHz of splitting, respectively.
At most temperatures the spin–lattice relaxation, characterized by the time T1, of the Au spin system is expected to be faster than the calculated J(WAu). Due to these fast spin flips, it is expected that the corresponding splitting of the W line will be averaged out. A single, narrow 183W signal is expected, about 13500 ppm upfield from the WO42− standard. In the unlikely case of long T1(Au) but fast molecular rotation, averaging out the quadrupole and direct spin–spin couplings, the W signal would be a 37-component weighted trisdekaheptuplet, with predicted intervals of 65 Hz, due to the coupling of the 183W nucleus to the total nuclear spin of the twelve gold atoms, with a total spin I
=
18.
The Mössbauer spectrum of every 197Au nucleus will be dominated by its quadrupole splitting. If the inner and outer hindered rotations are both slower than about 100 MHz, it will be observed, otherwise not. In gas-phase magnetic resonance the predominant quantization will come from the twelve weakly coupled quadrupolar systems, altogether coupled to the molecular rotation states. The solid-state nuclear quadrupole resonance should show a single signal corresponding to the splitting between mI=
±½ and ±
. For spin
that resonance will occur at NQCC/2, or 275 MHz.
This journal is © the Owner Societies 2004 |