• Aucun résultat trouvé

Electronic structure and absorption spectra of biferrocenyl and bisfulvalenide diiron radical cations: detection and assignment of new low-energy transitions

N/A
N/A
Protected

Academic year: 2022

Partager "Electronic structure and absorption spectra of biferrocenyl and bisfulvalenide diiron radical cations: detection and assignment of new low-energy transitions"

Copied!
14
0
0

Texte intégral

(1)

Electronic Structure and Absorption Spectra of Biferrocenyl and Bisful ACHTUNGTRENNUNG valenide Diiron Radical Cations: Detection and Assignment of New

Low-Energy Transitions

Ralf Warratz,

[a]

Hanane Aboulfadl,

[b]

Thomas Bally,*

[b]

and Felix Tuczek*

[a]

Introduction

Mixed-valent transition-metal dimers have been and contin- ue to be of interest for the investigation of inner-sphere electron transfer. On the basis of their tremendous impact on this research area, FeII–FeIIIbimetallocenes[1–8]can in fact be considered as the organometallic counterparts of the Creutz–Taube ion. A phenomenon that is exclusively associ- ated with this class of compounds and that has aroused con- siderable interest over the years is the transition from a charge-localized state at low temperatures to a delocalized

state at higher temperatures. This effect, which has been dis- covered and primarily investigated by means of Mçssbauer spectroscopy,[2]renders these systems particularly well suited to study the dependence of electron-transfer rates on small geometric changes, a subject that is of general interest to all areas of chemistry, biology, and physics.

The charge-localization–delocalization transition may be exemplified on ferrocenyl–ferrocenium (biferrocenyl) triio- dide, Fc2I3(actually Fc2C+I3;1) and 1’,1’’’-diethylbiferrocen- yl triiodide, (FcEt)2I3(2), which have been studied in detail by Mçssbauer spectroscopy and other techniques.[1–8]

In the solid state, Fc2I3 is charge localized between 4.2 and 300 K, and exhibits two discrete quadrupole doublets Abstract: UV-visible/near-IR (NIR)/

mid-IR (MIR) solution, solid-state, and matrix-isolation electronic absorption spectra of the FeII–FeIII mixed-valent homobimetallic compounds biferrocen- yl triiodide (1) and 1’,1’’’-diethylbiferro- cenyltriiodide (2) reveal the presence of a low-energy transition in the MIR region that has not been reported before. The new absorption feature and the known NIR band are both as- signed to intervalence charge-transfer (IVCT) transitions. To obtain insight

into the electronic structures of 1 and 2, DFT calculations with the BP86 functionals and different basis sets have been performed. Based on the molecular orbital scheme of cation 1, one band corresponds to the transition between the highest occupied dx2y2 or-

bitals on the two iron centers, whereas the other one is assigned to a transition from a lower-lying dz2 orbital to the dx2y2 orbital. For comparison, the doubly bridged bisfulvalenide diiron cation (3) has been investigated by op- tical absorption spectroscopy and DFT calculations. The experimental and the- oretical results are discussed with re- spect to the degree of electron localiza- tion/delocalization in these systems.

Keywords: electronic structure · iron · mixed-valent compounds · optical absorption spectroscopy · radical ions

[a] Dr. R. Warratz, Prof. Dr. F. Tuczek Institut fr Anorganische Chemie Christian Albrechts Universitt Kiel Otto Hahn Platz 6/7, 24098 Kiel (Germany) Fax: (+49) 431-880-1520

E-mail: ftuczek@ac.uni-kiel.de [b] H. Aboulfadl, Prof. T. Bally

Dpartement de Chimie Universit de Fribourg

Chemin du Muse 9, 1700 Fribourg (Switzerland) Fax: (+41) 26-300-97-37

E-mail: Thomas.Bally@unifr.ch

Published in "Chemistry - A European Journal 15(7): 1604 - 1617, 2009"

which should be cited to refer to this work.

http://doc.rero.ch

(2)

for FeII and FeIII. At a critical temperature (Tc) of 360 K, a transition to a charge-delocalized state occurs, which is evi- dent from the appearance of a single quadrupole dou- blet.[9, 10] In contrast, for (FcEt)2I3 the charge-localization–

delocalization transition occurs atTc=275 K. At lower tem- perature the system is localized (two quadrupole doublets) and at higher temperature delocalized (one quadrupole dou- blet). Importantly, the Mçssbauer spectra do not exhibit line broadening at any temperature, in particular also not in the vicinity of Tc. This indicates that electronic relaxation is always rapid on the Mçssbauer time scale, a result that has been difficult to understand for a long time.[11]

A variety of theoretical approaches have been developed to interpret these phenomena.[12, 13, 14] In the vibronic model of Varret and co-workers, which has been further elaborated by Klokishner et al.,[15] the charge-localization–delocaliza- tion transitions are attributed to cooperative interactions be- tween the bimetallocene molecules in the crystalline solid, which modulate the energy difference (Wd) between the local minima of the double-well potentials of the individual dimers. At or above Tc the energy difference Wdbecomes zero. The FeII–FeIIIunits thus can be described as symmetri- cal class II–III mixed-valent dimers (according to the Robin and Day classification scheme),[16, 17]and in the presence of fast relaxation a charge-delocalized state (exhibiting, for ex- ample, one Mçssbauer doublet) is observed. Below Tc the potential becomes asymmetric and Wd may become large compared to kT, leading to a charge-ordered state. In this case two distinct Mçssbauer doublets can be observed, one mostly deriving from FeIII and the other one mostly from FeII. The cooperative interactions that trigger the transition from the charge-delocalized to the charge-ordered state sen- sitively depend on the substituents[9, 12] and the counter- ions[18, 19]of the biferrocenyl radical cations, thus explaining the observation of differentTcvalues for different systems.

Besides Mçssbauer spectroscopy, a range of physical and spectroscopic methods have been employed to study the electronic and geometric structure as well as structural fea- tures of these intriguing systems.[1] In particular, infrared and EPR spectroscopies were applied to detect tempera- ture-dependent phenomena. Thereby it was found that the transition between static and fluxional behavior could occur at different temperatures, depending on the characteristic time scale of the method. Moreover, heat capacity measure- ments were conducted to monitor the enthalpic changes as- sociated with the charge-localization–delocalization transi- tion.[20]Important information on the structural consequen- ces of this transition was also obtained by X-ray crystallog- raphy. This technique conclusively demonstrated the pres- ence of an asymmetric distortion of the individual dimers belowTcthat disappears atTcand higher temperatures.[21]

In spite of the wealth of structural, physical, and spectro- scopic data being available for FeII–FeIIIbimetallocenes, rel- atively little information exists with respect to their optical absorption properties. This is surprising because for other mixed-valence systems the detection and detailed analysis of the intervalence charge transfer (IVCT) transition has

played a decisive role in their physico-chemical characteriza- tion.[22, 23]An IVCT transition has been detected for biferro- cenyl cations and has been interpreted on the basis of the Piepho–Krausz–Schatz (PKS) model.[24] However, no tem- perature dependence was recorded and no good fit of the room-temperature data was obtained.[25]We therefore decid- ed to reinvestigate the radical cations of Fc2and (FcEt)2by means of optical absorption spectroscopy, both in the solid- state and in solution. Here the results of this study are pre- sented. To obtain the full spectroscopic information, the wavelength range is extended to the mid-IR (MIR) range where, in addition to the known near-IR (NIR) transition, a new absorption feature is detected.

To interpret the optical spectra, in particular the unex- pected presence of two absorption features in the low- energy range, electronic structure calculations were per- formed. These calculations allowed us to assign these fea- tures to two spectrally overlapping transitions. For compari- son with the singly bridged biferrocenyl cations Fc2C+ and (FcEt)2C+ the corresponding doubly bridged bisfulvalenide diiron cation Fc=FcC+was investigated by the same methods.

Importantly, this dimer also exhibits a two-peak MIR transi- tion, a phenomenon that has been known and interpreted before.[6, 25, 26, 27] Moreover, Mçssbauer spectroscopy results showed that the bisfulvalenide cation is always delocalized and no localization–delocalization transitions exist in this type of molecule. Herein, the electronic structure of this dimer is also evaluated by means of DFT calculations and compared with that of the singly bridged counterparts. A forthcoming study will be devoted to a detailed analysis of the band shapes of the observed low-energy absorption bands, and in particular, their temperature dependence.[28]

Experimental Section

Materials: Biferrocene (Fc2),[12, 29] 1’,1’’’-diethylferrocenyl biferrocene ((FcEt)2),[9]and bisfulvalenide diiron (Fc=Fc),[30]as well as the triiodide and hexafluorophosphate salts of the corresponding radical cations, were prepared according to literature methods and checked for purity by ele- mental analysis (see the Supporting Information).

Analysis: UV-visible-NIR spectra were recorded on a Cary 5 UV-Vis- NIR spectrometer using quartz cuvettes (QX) withd=0.5 cm for solu- tion spectra or KBr pellets (10 tons cm2). Resolution was set to 1 nm.

Polystyrene films were prepared by codissolving the biferrocenyl salts and polystyrene and subsequently evaporating the solution slowly. MIR spectra were recorded with a Genesis Type I spectrometer (ATI Mattson, Madison) in the range of 400 to 6000 cm1with an accuracy of 2 cm1. Substances were pressed under vacuum with KBr at 10 tons cm2or mea- sured in solution within quartz cuvettes (QX, 0.5 cm path length) or a CaF2cell (0.1 cm path length). Elemental analysis was performed with a Euro Vector CHNS-O-element analyzer Euro EA 3000. Samples were burned in sealed tin containers with a stream of oxygen.

In the matrix-isolation experiments, neutral Fc2or Fc=Fc was sublimed from a heatable tube and codeposited with a stream of argon containing CH2Cl2(1 mol %) onto a CsI plate held at 20 K in a closed-cycle cryostat.

After deposition and recording of initial spectra, the sample was exposed to X-irradiation from a tungsten target (40 kV/40 mA) for 90 min. This results in the creation of electron–hole pairs in the Ar matrix, some of which separate and move through the matrix. The electrons undergo dis- sociative attachment to CH2Cl2whereas the holes are ultimately trans-

http://doc.rero.ch

(3)

ferred to the substrate, Fc2or Fc=Fc, which is thereby ionized by electron transfer to ArC+.[31]Spectra were recorded on a Perkin–Elmer Lambda 900 spectrometer (200–3000 nm) and on a Bomem DA3 interferometer (IR range).

Density functional theory calculations: Spin-unrestricted DFT calcula- tions were performed by the BP86[32, 33]functional, which had proven to give satisfactory results for neutral metallocenes.[34]In calculations using the Gaussian program package[35]the LANL2DZ basis set was used, which implies Dunning/Huzinaga full double-zeta (D95) basis func- tions[36]on the first row and Los Alamos effective core potentials plus double-zeta (DZ) functions on all other atoms.[37]In calculations with the Amsterdam density functional (ADF) program package[38]a triple-zeta polarized (TZP) set of Slater-type atomic orbitals (1 set of polarization functions, frozen core) optimized for use with the zero-order regular ap- proximation (ZORA)[39]to the relativistic Dirac equation was used. All structures were fully optimized with each method before performing single-point time-dependent (TD) DFT excited-state calculations.[40]Mo- lecular orbitals were plotted with the ADFView facility.

Results Biferrocenyl cations

Electronic and geometric structure: The electronic structure of Fc2C+ was modeled by DFT calculations. In the solid state the two units of the singly linked Fc2C+ are arranged in a transconfiguration, as in neutral Fc2, and the cation exhibits perfect C2hsymmetry;[12] so the calculations were also per- formed in this point group. Figure 1 shows the bond lengths

resulting from geometry optimizations of Fc2 and Fc2C+ by the Gaussian program (data for the geometries from the cal- culations with ADF can be found in the Supporting Infor- mation, Figure S1). It is seen that oxidation of the neutral dimer leads to an increase of the Fe–Cp distances and to a slight shortening of the bond linking the two Fc moieties, whereas the bonds in the Cp rings are barely affected. These changes indicate that the MO from which ionization occurs must have some ligand character and, in particular, is slight- ly antibonding along the bond linking the two Fc halves.

A section of the molecular orbital scheme of Fc2contain- ing the relevant MOs is shown in Figure 2. The underlying coordinate system is oriented such that the x axis is along the CC linkage and thezaxis is along the metal–Cp direc- tion. The MOs of the dimer can be derived as linear combi- nations of those of Fc that are shown on the left-hand side of Figure 2 (Kohn–Sham orbitals in D5hsymmetry). On the right-hand side the resulting orbitals of Fc2 are shown. In the radical cation all orbitals of the closed-shell dimer shift and split into aandborbitals. For reasons of simplicity this is not shown in Figure 2. Moreover, only the closed-shell MOs are given as the shape of the orbitals was not found to change much upon ionization of the neutral dimer.

The metal-centered d MOs of Fc combine to give the six highest-lying MOs of Fc2. The HOMO 18agof Fc2(i.e., the singly occupied molecular orbital (SOMO) of Fc2C+), corre- sponds to the positive combination of the dx2y2AOs of Fc, which extends partially to the ligand and is slightly anti- bonding along the CC bond that links the two Fc moieties.

The corresponding negative combination, 17bu, exhibits much less ligand character. The two linear combinations of the dxyAOs of Fc are also located between these two levels.

At lower energy the two dz2 combinations 17ag/16bu are found, which are barely split. Below these d-type levels the eight +/ combinations of the Fc ligandporbitals 4e1’and Fc metal-ligand bonding orbitals 3e1’’are located, with 4e1’ being at higher energy than 3e1’’. In the dimer some hybridi- zation between these doubly degenerate sets of orbitals occurs such that the highest-energy combination 16ag(which is antibonding with respect to the CC linkage) gets some bonding metal dxzorbital contribution. The same applies to its counterpart at lower energy, 15bu, which is slightly bond- ing with respect to the CC linkage. The two combinations in between, 10auand 10bg, have very little metal character.

At even lower energy the corresponding dimer combina- tions of the 3e1’’Fc orbitals are found.

Above the occupied metal orbitals of Fc2, the four vacant MOs 18bu, 19ag, 12au, and 12bg are located, which are +/ combinations of the Fc metal-ligand antibonding 4e1’’ dxz and dyzorbitals, respectively. The LUMO, 18bu, has bonding character with respect to the CC linkage. At even higher energy, a ligand p*-type orbital is found (19bu), which also has some metal character and is bonding with respect to the CC linkage.

As is usual with DFT calculations, the valence orbital en- ergies of neutral compounds lie about 2 eV too high in energy, but the general pattern is in line with the photoelec- tron (PE) spectrum of Fc.[41] In particular, the gap of ap- proximately 1.5 eV between the metal-centered d MOs of Fc (3e2’and 5a1’; 3e2gand 5a1g, respectively, in the “native”

D5d symmetry of Fc) and the ligand-based 4e1’ and 3e1’’

MOs (4e1uand 3e1g, respectively, inD5d) is reproduced well.

Our calculations for Fc2are also in excellent agreement with the photoelectron spectrum of this dimer,[41]which shows a double-humped first band, the first component of which has a width of 0.56 eV (0.22 eV more than in Fc), whereas that of the second peak is only 0.22 eV (0.11 eV more than in Figure 1. Structure of Fc2and its radical cation, calculated by the BP86/

LANL2DZ method, with interatomic distances in indicated in normal font for the neutral dimer and in italic for the radical cation.

http://doc.rero.ch

(4)

Fc). According to Lichtenbergers analysis,[41]the first, wide peak is due to the dxy and dx2y2 AOs, whereas the second, narrower peak is due to the dz2 AOs of the Fc units. The present DFT calculations suggest that the width of the first peak is due mainly to a splitting of the dx2y2 levels (18ag/ 17buin Fc2), whereas the dxy levels remain almost degener- ate. The narrow width of the second peak is due to the small splitting of the lower-lying dz2 levels (17ag/16bu). For the latter set of MOs, our calculations predict the bonding combination (17ag) to lie approximately 0.05 eV higher than the antibonding one (16bu), which might appear strange, but is due to subtle rehybridization among those combinations of d AOs, which become symmetry equivalent in the C2h

point group of Fc2.

Optical absorption spectroscopy: Figure 3 shows the optical absorption spectrum of the biferrocenyl (Fc2C+) triiodide along with that of neutral Fc2and the spectra of the constit- uents ferrocene (Fc) and ferrocenium (FcC+) (as a hexafluor- ophosphate salt) in the range from 2500 to 50 000 cm1. In the inset the spectra are scaled up in the range below 20 000 cm1. Band positions and intensities are collected in Table 1. Importantly, Fc2I3exhibits an absorption feature in the NIR/MIR region (band I) that is not present in all other systems (see Figure 3) and thus must be an intrinsic property

of Fc2C+. The higher-energy component of this feature (I b) had been observed previously and was attributed to an Figure 2. Kohn–Sham molecular orbitals of Fc (left-hand side) and Fc2(right-hand side) from BP86/DZACHTUNGTRENNUNG(ZORA) calculations with the ADF program.

Blue levels are virtual MOs, red levels are open shells in the radical cations, and black levels are doubly occupied MOs. Dipole-allowed IR transitions are indicated by green arrows.

Figure 3. UV-visible-NIR-MIR spectra of biferrocenyl triiodide (Fc2I3; c) along with those of the constituent subunits ferrocene (Fc;d) and ferrocenium hexafluorophosphate (FcPF6; g), and the neutral parent neutral compound biferrocene (Fc2;b) in [D3]acetonitrile.

http://doc.rero.ch

(5)

ICVT transition (vide supra), whereas the lower-energy component (I a) has never been detected before. Very simi- lar absorption features in the IR range are observed for the diethyl derivative, (FcEt)2C+ (Figure 4). In particular, band I

is again found to exhibit two maxima, one at about 3900 cm1 (a) and the other one at 5000 cm1 (b). When going from dichloromethane to acetonitrile, the solvent shift of the lower-energy component I a is weaker (ca. 300 cm1) than that of the higher-energy one (I b; 600 cm1). From this we conclude that the two maxima correspond to two differ- ent electronic excitations. Further solvent shifts are collected in Table 2. At higher energy (17 000 cm1), a second ab- sorption feature (band II) is observed that also seems to contain at least two transitions. Upon changing to the more polar solvent this feature exhibits a marked change in ap- pearance as well; in particular, the lower-energy component (band II a) appears to decrease, whereas the higher-energy component (band II b) appears to increase in intensity.

At approximately 27 500 and 34 000 cm1 the spectrum of Fc2I3 (Figure 3) shows two intense bands which, however, are due to the triiodide counterion. Efforts to record optical

solution spectra of the mixed-valent dimer with counterions lacking intrinsic absorptions in the visible region (e.g., PF6) were unsuccessful, as in this case the IR features (bands I a and I b) disappeared, which indicated decomposition of the Fc2C+cation. In Figure 5 the spectra of Fc2I3and (FcEt)2I3in

a KBr matrix are given. The spectra show again the pair of NIR/MIR transitions (bands I a and I b) at 3000 and 5000 cm1; moreover, at 14 000 and 18 000 cm1 the double- humped absorption of band II is located. At higher energy the intense triiodide absorptions at approximately 27 000 and 34 000 cm1 are observed. Interestingly, the solid-state absorption spectra seem to contain an additional absorption around 10 000 cm1, which is lacking in the solution spectra.

A peak-fit analysis allows to more exactly locate this band at approximately 8750 cm1; that is, between bands I and II (see Figure S2 in the Supporting Information). This band exists both in triiodide and hexafluorophosphate salts of Fc2C+ and (FcEt)2C+, thus eliminating the possibility of a triiodide absorption that would also lead to a broad band over the whole visible range in the solid-state spectra. The fact that this band is exclusively observed in KBr pellets (which contain crystallites of the compounds) but not in so- lution or matrix-isolation spectra (see below) leads to the conclusion that it is probably due to an intermolecular (in- terdimer) charge-transfer transition in the crystal. Similar observations have been made on single crystals of ferroceni- um trichloroacetate; also in this case an intermolecular Table 1. Electronic absorption bands (in cm1) of Fc2C+.

Band Band position of sample in Assignment no. CH3CN[a] KBr pellet solid Ar/10 K of transition[b]

I a 3900 (1000) 3000 3000 12Bu(17bub!18agb/

16bub!18agb) I b 5500 (1000) 5000 5000 22Bu(17bub!18agb/

16bub!18agb) II a 14 100 (1000) 14 000 15 000 42Bu(15bub!18agb;

“hole transition”) II b 18 200 (2000) 18 000 18 000 72Bu(18aga!18bua;

SOMO–LUMO)

III 27 000 ligand!18bu(LMCT;

mixed with LF) [a]evalues in brackets. [b] Refer to Figure 2 and Table 5.

Figure 4. Low-energy bands of the 1,1’’’-diethylbiferrocenyl triiodide ((FcEt)2I3) in dichloromethane (c) and in acetonitrile (g), using a CaF2cell (300 K).

Table 2. Solvent dependence of low-energy bands ( in cm1) for (FcEt)2I3.

Solvent Band I b Band I a

CD3CN 5400 3900

CD2Cl2 4900 3700

[D8]THF 5300 3900

ACHTUNGTRENNUNG(CD3)2CO 5500 3600

CD3Cl 4900 3700

Figure 5. Visible-NIR-MIR spectra of (FcEt)2I3(g) and Fc2I3(c) in KBr pellets (300 K).

http://doc.rero.ch

(6)

charge transfer has been invoked to explain the presence of this band.[42]

To obtain spectroscopic information about Fc2C+ without interfering effects of the counterion, the solvent, or the crys- talline environment, we turned to matrix-isolation spectros- copy. Figure 6 shows the electronic absorption spectrum of

Fc2C+ obtained by ionization of Fc2in an Ar matrix at 6 K as well as the absorption spectrum of the parent, neutral dimer. As in these experiments the negative charge carriers are separated by one or more layers of Ar from the radical cation, the spectrum of Fc2C+ can be regarded as being due to isolated radical cations, unperturbed by any counterion, solvent (radical ions with such small electron affinities as Fc2C+ interact only very weakly with Ar), or solid-state ef- fects. Nevertheless, the spectrum shows similar bands as the above solution and solid-state spectra, except that there is now a clear gap of approximately 1 eV between the two IR bands I a/I b and the double-humped band system (II) that follows in the visible range. Moreover, the Ar spectrum shows an intrinsic absorption band of the Fc2C+ radical cation (band III) at a similar position (27 000 cm1) at which the first triiodide absorption in the solution and solid-state absorption spectra of Fc2I3is located (vide supra). Above 32 000 cm1 the spectrum in Ar is masked by strong absorptions of Fc2(only 10–15 % of which are converted to the radical cation in this experiment).

Importantly, the spectrum of the neutral dimer is very dif- ferent from that of the radical cation, showing an absorption at approximately 23 000 cm1and a shoulder at 30 000 cm1. (Table 3). The solution spectrum of Fc2 given in Figure 3

further indicates the presence of an intense absorption band at 33 000 cm1. We also recorded the IR spectra of Fc2 and its radical cation (see Figure S3 in the Supporting Informa- tion), and both are in good agreement with the calculated spectra for the two species (see the Supporting Informa- tion).

A comparison of the spectra of Fc2C+ in solution (CH3CN), KBr, and Ar is given in Figure 7. As clearly shown by this figure, qualitatively similar spectra are ob- tained. However, there are conspicuous differences in band positions and intensities. In particular, it is evident that the IR bands I a and I b are markedly shifted to higher energies in the solution spectrum, which can be attributed to solvent effects. Moreover, the intensity distribution of the two com- ponents of band I in the Ar spectrum is reversed in solution:

whereas in the former case the high-energy transition is much less intense than the transition at lower energy, the high-energy transition becomes more intense than the low- energy transition in solution. Similar changes are observed in the double-humped band at 18 000 cm1 (band II) when going from the Ar matrix to the solution spectrum. The spectral differences at higher energy are due to the fact that matrix-isolation data show an intrinsic absorption band of the Fc2C+ radical cation (band III) at a similar position ( Figure 6. Electronic absorption spectra of Fc2 (b) and Fc2C+ (c) in

Ar at 10 K. The filled squares (&) denote electronic transitions predicted by TD-BP86 calculations using the TZP basis set and the ZORA method as implemented in the ADF program, whereas the open squares (&) refer to calculations using the LANL2DZ basis set with the Gaussian program as listed in Tables 4 (for Fc2) and 5 (for Fc2C+).

Table 3. Electronic absorption bands (in cm1) of Fc2. Band position of sample in Assignment CH3CN[a] solid Ar/10 K of transition[b]

23 000 (500) 23 500 11Bu(18ag!18bu;

HOMO–LUMO) 30 000 (sh.)[c] 30 000 (sh.)[c] 71Bu(16ag!18bu; LMCT)[d]

33 000 (6500) 81Bu(18ag!19bu; MLCT)

38 500 (7300) ?

[a]evalues in brackets. [b] Refer to Figure 2 and Table 4. [c] sh.=should- er. [d] In ADF mixed with 18agACHTUNGTRENNUNG(HOMO)!19bu, MLCT.

Figure 7. Comparison of optical absorption spectra of Fc2C+ in KBr, CH3CN (also multiplied by 10 for clarity), and an Ar matrix.

http://doc.rero.ch

(7)

27 000 cm1) as the first triiodide absorption in the solu- tion and solid-state absorption spectra of Fc2I3(vide supra).

Upon lowering the temperature from 300 to 10 K, the band shapes change in the spectra taken in KBr pellets, but the overall change in intensity is low. In contrast, the ab- sorption spectrum recorded in a polystyrene matrix is almost temperature independent (data not shown); as a matter of fact this spectrum is more related to that taken in dichloromethane (cf. Figure 4). The temperature depend- ence of the solid-state spectra is considered in more detail in a forthcoming paper.

TD-DFT calculations and band assignments: To assign the electronic absorption bands of Fc2 and Fc2C+, TD-DFT cal- culations were performed. As before, two different basis sets and programs were employed (ADF, G03; cf. the Ex- perimental Section). The results of the calculations are col- lected in Tables 4 and 5 for Fc2 and Fc2C+, respectively.

Moreover, the positions and intensities of the electronic transitions calculated by means of the ADF and Gaussian 03 programs are indicated schematically in Figure 6. The result- ing assignments of experimentally observed bands are in- cluded in Tables 1 and 3.

For neutral Fc2(cf. spectra in Figures 3 and 6) the assign- ments are fairly clear cut: the band at approximately 23 000 cm1most probably is associated with the 18ag!18bu

excitation (11Bu state), which corresponds to the HOMO–

LUMO transition (dx2y2!dxz) in this dimer that is predicted a bit too low in energy by both methods (Table 4). The shoulder at approximately 30 000 cm1, which also is visible

in the solution spectrum of Fc2 is probably associated with the transition to the 71B1u state, which corresponds to the

16ag!18buACHTUNGTRENNUNG(LUMO) excitation, a ligand!metal CT

(LMCT) transition. In the ADF calculations this excitation is mixed with the 18agACHTUNGTRENNUNG(HOMO)!19butransition, a metal!

ligand CT (MLCT) transition. The intense band at approxi- mately 33 000 cm1corresponds to the 81Bustate, and is as- sociated with the same charge-transfer transitions, but this time with a preponderance of the 18ag!19buLMCT transi- tion. Between the 11Bu and the 71Bustate, five ligand-field excited states (21Buto 61Bu) are located that are associated with transitions of dx2y2!dxz, dxy!dyz, and dz2!dxz charac- ter. The corresponding intensities, however, are predicted to be weak by both methods and are therefore not expected to give rise to distinct bands.

Table 5 shows the results of TD-DFT calculations on Fc2C+. We had to include over 100 roots in these calculations to reach 30 000 cm1(most of the transitions are not listed in Table 5 because they are either dipole forbidden or carry vanishingly small oscillator strengths), so we refrained from carrying out calculations of higher-lying excited states, be- cause transitions to such states could not be observed. In line with expectations from the MO scheme in Figure 2, these calculations predict three dipole-allowed transitions to occur in the low-energy range (below 7000 cm1), all of which involve mainly excitation of an electron from one of Table 4. Electronic transitions (n˜in cm1) of Fc2calculated by the TD-

BP86 method.

States ADF[a] Gaussian[b] Configurational [cm1] f [cm1] f composition[c]

11Bu 20 168 0.0443 19 361 0.0320 18ag!18bu(0.86/0.61) 11au!12bg(–/0.25) 21Bu 21 458 0.0048 21 761 0.0004 11bg!12au(0.56/0.60)

11au!12bg(0.42/0.39) 31Bu 22 184 0.0036 22 104 0.0108 17ag!18bu(0.89/0.52)

11au!12bg(–/0.40) 41Bu 23 365 0.0067 22 991 0.0103 17bu!19ag(0.48/0.52)

17ag!18bu(–/0.37) 11au!12bg(0.26/0.21) 11bg!12au(0.22/–) 51Bu 25 503 0.0026 25 653 0.0038 16bu!19ag(0.70/0.56)

17bu!19ag(0.20/0.30) 17ag!18bu(–/0.30) 61Bu 28 405 0.0020 28 396 0.0020 16bu!19ag(0.25/0.41)

17bu!19ag(0.26/0.31) 71Bu 30 968 0.0442 30 655 0.0839 16ag!18bu(0.65/0.59)

18ag!19bu(0.65/–) 81Bu 33 355 0.2421 33 409 0.1970 18ag!19bu(0.56/0.56)

16ag!18bu(0.25/0.26) [a] BP86/TZP/ZORA. [b] BP86/LANL2DZ. [c] In terms of the MOs de- picted in Figure 2. The coefficients of the CIS-type wave function are from the (ADF/Gaussian) calculations, respectively.

Table 5. Electronic transitions (in cm1) of Fc2C+calculated by the TD- BP86 method.

States ADF[a] Gaussian[b] Configurational [cm1] f [cm1] f composition[c]

12Au 4760 <104 3791 <104 11aub!18agb(–/1.03)

12Bu 5405 0.013 5258 0.043 17bub!18agb(0.52/0.66) 16bub!18agb(0.47/0.27) 22Bu 6775 0.037 6678 0.008 17bub!18agb(0.45/0.13) 16bub!18agb(0.52/1.01) 22Au 16 375 0.001 16 428 0.001 11aub!18agb(0.99/0.90) 42Bu 18 635 0.009 16 495 0.017 15bub!18agb(0.94/0.42) 17agb!18bub(0.02/0.37) 17aga!18bua(0.02/0.21) 17bub!19agb(0.01/0.55) 52Bu 20 005 0.009 18 237 0.004 17bub!19agb(0.61/0.34)

17aga!18bua(0.17/

<0.01)

15bub!18agb(0.04/0.56) 72Bu 21 092 0.034 20 945 0.034 18aga!18bua(0.44/0.57) 17agb!18bub(0.36/0.57) 12/

132Bu

25 994 0.036 25 109 0.023 16aga!18bua(0.51/0.66) 16agb!18bub(0.22/0.43) 16bua!19aga(–/0.52) 14bub!18agb(0.22/–) 14/

152Bu

26 719 0.002 25 653 0.009 16bua!9aga(0.70/–) 14bub!18agb(–/0.70) 14/

152Bu

28 216 0.058 15aga!18bua(–/0.57) 15agb!18bub(–/0.41) 162Bu 29 126 0.117 28 859 0.125 16agb!18bub(0.40/0.67)

15agb!18bub(–/0.33) 10aua!12bga(–/0.31) [a] BP86/TZP/ZORA. [b] BP86/LANL2DZ. [c] In terms of the MOs de- picted in Figure 2. The coefficients of the CIS-type wave functions are from the (ADF/Gaussian) calculations, respectively.

http://doc.rero.ch

(8)

the metal-centered MOs to the 18ag SOMO. The first of these transitions leads to a2Auexcited state, but because it is polarized perpendicular to the metal–metal axis, its oscil- lator strength is too weak for it to be observed experimen- tally under our conditions (it corresponds to the low-lying

2A1g!2E2gIR transition in FcC+, which has also never been observed experimentally.[42]The following two transitions to

2Bu excited states that lie 0.65–0.85 eV above the ground state are more intense, with one being three to five times stronger than the other. The two basis sets (and the associat- ed modeling of the Fe core AOs) that we have employed in our study give slightly different quantitative predictions with regard to these states (the ordering of the two Bu states even appears to change between the two models), but the calculations leave little doubt that the two mid-IR bands I a and I b that we have observed are due to these two2Ag!2Bu

transitions, one of which (I b) shows more solvatochromism than the other (I a).

Following these essentially metal–centered transitions there is a gap of over 1 eV in the calculated spectrum, again in agreement with the PE spectrum of Fc2 and the optical spectrum obtained in an Ar matrix (which is free of mani- festations due to interactions of Fc2C+ with itself, its counter- ions, or solvents; see above). After that gap we find another very weak transition to an Austate, followed by three more intense transitions to Buexcited states. These involve excita- tions of electrons from doubly occupied metal- (17bu, 17ag) or ligand-based MOs (15bu) to the 18ag SOMO and/or to virtual metal/ligand-type MOs (18bu, 19ag). The largest am- plitudes within these excitations are associated with the transition from the ligand-type orbital 15bu to the SOMO (corresponding to the “hole transition” of Fc+) and the SOMO!LUMO transition (18aga!18bua), but the weight- ing of the excitations differs considerably between the two DFT methods. However, both of them position the latter transition at approximately 21 000 cm1, about 1000 cm1 higher in energy than the (comparatively “pure”) HOMO! LUMO transition in neutral Fc2(vide supra). The hole tran- sition is located by ADF at approximately 18 600 cm1, 2400 cm1 higher than its counterpart in monomeric Fc+ (16 200 cm1). G03 distributes the 15bu!18agexcitation over two bands, one at 16 500 cm1and one at 18 200 cm1(the in- tensity of the first one being, however, much higher than that of the second one).

In summary, TD-DFT predicts several transitions in the range of 7000 to 22 000 cm1, the most intense of which is as- sociated with the SOMO!LUMO excitation. These transi- tions must be associated with the double-humped band II, but the calculations do not allow an unambiguous assign- ment. Based on the fact that the calculations predict the SOMO–LUMO transition of Fc2C+ to be close in energy to the HOMO–LUMO transition of Fc2(i.e., at20 000 cm1) and the hole transitions are predicted by both methods at lower energy, we are tempted to assign the lower-energy component of band II (II a) to excitations of the latter type (cf. Table 1). This would also straightforwardly account for the fact that the visible part of the absorption spectrum of

Fc2C+ extends to much lower energy (1 eV) than that of Fc2, due to the existence of a hole in the valence shell. In comparison with the experimental spectrum all predictions appear 1000–2000 cm1 too high in energy, which also ap- plies to the two IR transitions I a and I b.

Between 22 000 and 30 000 cm1TD-DFT predicts excita- tions with nonvanishing intensities to three to four addition- al2Buexcited states. These involve excitations from ligand- type orbitals to the metal-type LUMO (18bu); namely, LMCT transitions, mixed with ligand-field (LF) transitions (16bu!19ag), and probably are associated with band III and its shoulder in the Ar matrix spectrum of Fc2C+. In contrast to neutral Fc2, no mixing with MLCT transitions is predicted for this part of the spectrum, probably as a consequence of the higher charge of the iron centers.

Bisfulvalenide diiron cation

Geometric and electronic structure: Although closely related to the singly bridged dimers1and2, the doubly bridged bis- fulvalenide diiron cation, Fc=FcC+ (3), differs from the former systems insofar as oxidation of the doubly bridged biferrocenylene has a larger impact on the molecular struc- ture, especially the metal–metal distance. Whereas in the neutral compound this distance is 3.98 , it decreases in the hexafluorophosphate[43] salt of the monocation to 3.68 and in the picrate hemihydroquinone[44] system to 3.63 . These findings indicate a stronger metal–metal interaction than that present in Fc2C+, which also is reflected by the fact that the planes of the Cp rings of each fulvalenide ligand are rotated by about 108 around an axis perpendicular to the CC linkage such that the cross section of the Fc=FcC+ cation through the centers of the fulvalenide ligands and the Fe atoms is distorted from rectangular to ellipsoidal.[44] All these features are clearly borne out by the DFT calculations (see Figure 8 for the G03 geometries; the ADF geometries are shown in Figure S4 in the Supporting Information). The coordinate system is chosen such that thexaxis is along the metal–metal direction and thezaxis along the metal–Cp di- rection.

Figure 8. Structure of Fc=Fc and its radical cation, calculated by the BP86/LANL2DZ method, with interatomic distances [] indicated in normal font for the neutral dimer and in italic for the radical cation.

http://doc.rero.ch

(9)

A section of the molecular orbital scheme of Fc=Fc con- taining the relevant MOs (calculated at the optimized struc- ture of Fc=FcC+) is shown in Figure 9. Again dimer MOs can be derived as linear combinations of those of Fc, which are shown on the left-hand side of Figure 2 above. In particular, the metal-centered MOs of Fc combine to give the six high- est occupied MOs of Fc=Fc. The HOMO 9b2u (i.e., the singly occupied MO of Fc=FcC+) corresponds to the negative combination of the dx2y2 AOs of Fc, which again extends somewhat to the ligand and is slightly antibonding along the CC bonds linking the two Fc moieties, which explains why their distance shrinks on ionization. The corresponding bonding combination is the 10agMO, which exhibits almost no ligand character. Between these two levels we find the four closely spaced metal-centered MOs that arise through linear combination of the dz2 and the dxy AOs of Fc. At

lower energy linear combinations of the Fc ligand p MOs 4e1’ and the Fc metal-ligand bonding MOs 3e1’’ are found.

At higher energy the unoccupied linear combinations of Fc dxzand dyzorbitals are located, which are slightly antibond- ing with respect to the Cp ligands. The LUMO (9b2u) de- rives from the out-of-phase combination of the Fc dxz orbi- tals. At even higher energy, ligand p* orbitals are located (11ag, 10b3u).

Optical absorption spectroscopy: Figure 10 shows the ab- sorption spectra of the doubly bridged ferrocenylene ferro- cenylenium dimer (=bisfulvalenide diiron cation) Fc=FcC+,

along with those of the parent neutral dimer Fc=Fc and the constituent subunits ferrocene (Fc) and ferrocenium hexa- fluorophosphate, FcC+PF6. As shown at a larger scale in the inset, an important new feature of the oxidized compound is the intense band peaking at 6500 cm1(band I a), which ex- hibits a shoulder at approximately 8400 cm1 (band I b; cf.

Table 6). These features correspond to the MIR/NIR bands of the singly bridged Fc dimers and have been interpreted

Figure 9. MO scheme of Fc=Fc from BP86/DZACHTUNGTRENNUNG(ZORA) calculations with the ADF program. Blue levels are virtual MOs, red levels are open shells in the radical cations, and black levels are the doubly occupied MOs.

Dipole-allowed IR transitions are indicated by green arrows.

Figure 10. UV-visible-NIR-MIR spectra of Fc=FcPF6 (c) along with those of the constituent subunits ferrocene (Fc;d) and ferrocenium hexafluorophosphate (FcPF6;g) in acetonitrile and the neutral parent compound biferrocenylene (Fc=Fc;b) in KBr, 300 K.

Table 6. Electronic absorption bands (n˜in cm1) of Fc=FcC+. Band Band position of sample in Assignment no. CH3CN[a] KBr pellet solid Ar/10 K of transition[b]

I a 6500 (3000) 6200 6500 12Ag(10agb!9b2ub/

9agb!9b2ub) I b 8400 (2000) 8200 8500 22Ag(10agb!9b2ub/

9agb!9b2ub) 17 000 (1000) ACHTUNGTRENNUNG(17 000) 22B3g(5b3gb!9b2ub;

(“hole transition”)

22 000 (2000) 22 000 MLCT/LMCT

30 000 (8000) 30 000 42Ag(8agb!9b2ub;

LMCT)

37 000 (16 000) 37 000 CT

[a]evalues are in brackets. [b] Refer to Figure 9 and Table 9.

http://doc.rero.ch

(10)

as IVCT transitions in the literature.[25] This assignment is supported by the TD-DFT calculations, which attribute these bands to transitions from the doubly occupied metal- type orbitals to the SOMO 9b2u(see below).

The solution spectra of Fc=FcC+ in CD3CN and CD2Cl2

are compared in Figure 11. In contrast to the spectrum of the monobridged radical cations, bands I a and I b do not ex- hibit solvatochromic behavior.[45] The spectrum in KBr shows these bands at 6200 and 8200 cm1(data not shown).

As also evident from this figure, the temperature depen-

ACHTUNGTRENNUNG

dence of these bands in KBr pellets is less pronounced than in case of the Fc2C+ dimers; only a slight decrease in line width is observed upon lowering the temperature. At higher energy in the solution spectra (Figure 11) two medium-in- tense bands are found (17 000, 22 000 cm1), and at 30 000 cm1 the first high-intensity band appears. In any case, the transitions of the radical cation Fc=FcC+ are much more intense than those of its neutral counterpart.

To obtain more information about these transitions we also attempted matrix-isolation experiments with Fc=Fc.

However, these experiments were less successful than in the above case of Fc2in that an impurity of slightly higher vola- tility than Fc=Fc was always present. Although this impurity showed no tangible absorption below 40 000 cm1, it gave rise to rather intense bands in the range of 20 000–

35 000 cm1 after X-ray irradiation (i.e., ionization). There- fore we had to resort to subtraction of the spectra obtained in two ionization experiments in which different amounts of Fc=Fc were present in the matrix to obtain the spectrum of Fc=FcC+ presented in Figure 12. This spectrum shows similar transitions as were found in the solution spectra of Fc=FcC+, and in the following it will be used to assign the electronic transitions of Fc=FcC+, based on TD-DFT calculations.

TD-DFT calculations and band assignments: Tables 8 and 9 show the results of TD-DFT calculations on Fc=Fc and Fc=

FcC+, respectively. The calculated band positions and intensi-

ties are also represented schematically in the matrix-isola- tion spectra of these compounds (Figure 12). The resulting assignments of experimentally observed bands are included in Tables 6 and 7. As is already evident from this figure, the situation is by far more complex for the bisfulvalenide diiron systems than for the singly bridged biferrocene dimers. Both for the neutral and the radical cation DFT pre- dicts a multitude of densely spaced excited states out of which only a few, which are not dipole forbidden or are as- sociated with vanishingly small transition dipole moments, are listed in the tables (120 states had to be calculated to reach 45 000 cm1!).

In neutral Fc=Fc it is difficult to reach definite assign- ments for the observed bands (dotted line in Figure 12). The Figure 11. Visible-NIR spectra of Fc=FcPF6in acetonitrile (c) and di-

chloromethane (a).

Figure 12. Electronic absorption spectra of Fc=Fc in Ar at 10 K (bottom panel;c) and Fc=Fc2C+in Ar at 10 K (top panel;b) and in CH3CN (top panel;c) The filled squares (&) denote electronic transitions pre- dicted by TD-BP86 calculations using the TZP basis set and the ZORA method as implemented in the ADF program, whereas the open squares

(&) refer to calculations using the LANL2DZ basis set with the Gaussian

program, as listed in Tables 8 (for Fc=Fc) and 9 (for Fc=Fc2C+), respec- tively.

Table 7. Electronic absorption bands (in cm1) of Fc=Fc.

Band Band position of sample in Assignment

no. CH3CN[a] solid Ar/10 K of transition[b]

I a 21 200 (500) 22 000 (weak) LF (mixed)

I b 28 000 (shoulder) LF/CT

II a 35 000 ?

II b 39 000 (4000) 38 000 ?

[a]evalues are in brackets. [b] Refer to Figure 9 and Table 8.

http://doc.rero.ch

Références

Documents relatifs

The last World Bank World Development Report 2008 has clearly emphasized the coexistence of three developing worlds characterized by their level of economic diversification

and the observed and calculated energy levels along with the U4 + eigenvectors in the D2d symmetry are given in table VI. This table shows the fluorescing level of the

If females are considered to have, on average, lower societal status than males, they are more likely to be adversely affected by unsuccessful epi- sodes but also to experience

These are the first spin polarized ab initio self-consis- tent energy band studies of ferromagnetic transi- tion metal filme that are thick enough C9-layers for Ni(001) and 7

The spectrum obtained with an excitation energy of 570 eV is very similar to the spectrum obtained by electron bombardment ( Fig. Figure 4b shows the h3 x-ray emission

Probably only this mixing causes initial absorption in ionic and covalent crystals, because in spectra of water soluted ions [4] without the effect of the crystal-

range covered by the experiment two main absorption regions are observed, consisting of one single peak at low energies and a superposition of 3 to 6 peaks a t higher

R6sumB. - On discute les modkles thhriques employes pour l'interpretation des spectres elec- troniques de mol6cules triatomiques. I1 est possible qu'un molecule plie est