INTRODUCTION
An understanding of the interaction of gold with ligands at the molecular level is of fundamental interest in inorganic and organometallic chemistry.1-3 In particular, the interaction of excited states of metal clusters with ligands is important because these states may interact differently with the ligands than the ground state, opening up new possibilities for molecular activation and/or reactivity changes that could be selectively switched on or off by light. Another example is the concept of two (or multiple) state reactivity that based on the distinct state-dependent reactivity of transition metals, together with a spin-crossover during the course of a reaction.4 Correlations between cluster reactivity and electronic properties, such as ionization and excitation energies or HOMO-LUMO gaps, have been reported long ago, for example, for the binding of or , and are suggested to indicate intermediate electronic transitions during the reaction.5-7 Here, we present vibrationally resolved electronic spectra of the and complexes that provide deep insight into the electronic properties of the ground and first excited electronic states.
The importance of gold lies not only in its catalytic properties but also in its role as a model system for developing an appropriate theoretical description of strong relativistic effects. These relativistic effects and their influence on ligands and their binding motifs have been the topic of numerous studies.1,8,9 specifically was shown to catalyze the formation of ethylene,10,11 an observation that was later called into question and it was speculated that the observation could be explained by an excited state participating in the reaction.12-14 The ligands considered herein are also of fundamental interest. is a potent greenhouse gas and a better understanding of its metal-based catalytic reduction is desirable.15,16 is abundantly available and efficient reduction to its bioavailable form, , is an important chemical challenge.17-19
The optical properties of and have been reported previously.20,21 The first excited state of (A2Σu+) is formed from the ground state (X2Σg+) by excitation from an antibonding 5dσ* orbital into the half-filled bonding 6sσ orbital. As a result, the Au-Au bond distance decreases significantly upon electronic A ← X excitation. The X and A states of are described well using standard time-dependent density functional theory (TD-DFT) as is apparent from the similarity between computed and measured excitation spectra.
The manifold bonding and activation mechanisms of by transition metals have been described previously and coordination as well as back donation from the metal d-orbital into the LUMO of were identified to be the most critical components of the interaction.2 The reactions of clusters with have been studied in a Penning trap22 and complexes have been characterized via IR action spectroscopy.23,24 While the reactions in a Penning trap revealed several gas-phase pathways (especially dissociative charge transfer yielding ), the structural IR investigation of the ground state showed no indication of dissociation apart from simple loss.23,24 The latter studies also showed the -bound isomer to be formed predominantly, with the -bound isomer being substantially less stable. Many metal– complexes have been studied,25 and the strong bond was found to be activated in complexes.26 On the other hand, interaction with oxoacetates was found to strengthen the bond.27 This nonclassical behavior was explained by a lack of π back donation and delocalization of the antibonding 4σ() orbital, an effect that has also been described for isoelectronic in metal– complexes.28,29 Our study focuses on the effects of and ligands on the X and A states of , identifying the resulting binding motifs in each state and any changes in N2 and activation upon metal complexation (Figure 1). The detailed insight into the potential energy surface of these fundamental metal–ligand complexes arises from recent instrumental improvements, which allow us to measure vibronic spectra in unprecedented detail.20,30,31
[IMAGE OMITTED. SEE PDF]
RESULTS AND DISCUSSION
Vibronic spectra of ( and ) have been measured in the 0.8–1.5 eV range (Figure 2) via electronic photodissociation (EPD) of mass-selected clusters in a quadrupole/time-of-flight tandem mass spectrometer.20,31 Both clusters decay exclusively via loss of the ligand and the fragment is detected. Parent clusters are generated via laser vaporization (532 nm, 3 mJ/pulse, ∼0.5 mm diameter) of a turning and translating rod. and are added to the expansion gas in fractions of 0.1 %. For cluster generation, the plasma is expanded through a reaction channel (l = 3 cm, d = 4 mm) and a conical nozzle cooled by liquid . After passing through a skimmer, complexes are mass-selected in a quadrupole and guided into the extraction region of a reflectron time-of-flight mass spectrometer. Before extraction, they are overlapped with a laser pulse emitted from a tuneable optical parametric oscillator (bandwidth ≤10 cm–1, 1015–1016 photons/cm2) to induce dissociation. The photodissociation cross section is then determined as a function of photon energy from the relative parent and fragment ion signals, using an overlap factor of ion and photon beam of α = 0.6 as estimated from power-dependent measurements.31 The EPD spectra provide a good measure for the total absorption cross section if competing relaxation processes can be neglected. To aid the spectral assignment and to obtain deeper insight into the electronic structure of the complexes, DFT calculations are performed. The and à states of L are optimized at the CAM-B3LYP/cc-pVTZ level,32,33 including GD3BJ empirical dispersion and ECP60MDF effective core potentials,34 as implemented in GAUSSIAN16 (Figure 1 and Table 1).35 D0 and Ea values reported herein are the zero-point corrected dissociation and adiabatic excitation energies. Previous studies have shown that the lowest excited state of is described very well at this level of theory, giving a difference in Ea of less than 0.1 eV.20 The lowest energy electronic states were found to have doublet multiplicity for all complexes studied here with quartet states lying more than 2.5 eV higher in energy and thus ignored henceforth.
[IMAGE OMITTED. SEE PDF]
TABLE 1 Experimental properties of the ground and first excited state of , , and compared to computed values
Experiment | CAM-B3LYP/cc-pVTZa | |||||||||||
N2 | N2O | N2 | N2O | N2 | N2O | |||||||
State | Σ+ | Ã2Σ+ | Σ+ | Ã2Σ+ | Σ+ | Ã2Σ+ | Σ+ | Ã2Σ+ | X1Σg+ | 1Σg+ | X2Σg+ | A2Σu+ |
Ea / eV | 0.910(1) | 0.90(5) | 0.975 | 0.953 | ||||||||
Ev / eV | 1.07(2) | |||||||||||
Re(Au-L) / Å | ΔRe(, Ã) = 0.25(4)b | 2.067 | 2.264 | 2.068 | 2.219 | |||||||
re(Au-Au) / Å | Δre(, Ã) = −0.09(4)b | 2.606 | 2.459 | 2.607 | 2.458 | 2.635 | 2.447 | |||||
D0(-L) / eV | <0.91 | 1.05(3) | 0.89 | 0.64c | 1.06 | 0.86c | ||||||
ω1 / cm–1 | 2501 | 2514 | 2496 | 2476 | 2501d | 1359 | 140 | 193 | ||||
ω2 / cm–1 | 290(2) | 216(5) | 285 | 219 | 1452 | 1460 | 637 | |||||
ω3 / cm–1 | 147(2) | 179(2) | 230(10) | 138 | 180 | 245 | 220 | 2404 | ||||
ω4 / cm–1 | Δ(ω4‘‘,ω4‘) = 39(2) | 170(10) | 243 | 205 | 133 | 170 | ||||||
ω5 / cm–1 | 44 | 49 | 617 | 633 | ||||||||
ω6 / cm–1 | 104 | 106 | ||||||||||
ω7 / cm–1 | Δ(ω7‘‘,ω7‘) = –4(2) | 26 | 27 | |||||||||
ω2χ2 / cm–1 | 0.89(5) | |||||||||||
ω3χ3 / cm–1 | 0.8(1) | |||||||||||
χ23 / cm–1 | 0.85(5) | |||||||||||
kAu-Aub / 102 N m–1 | 2.3(1) | 3.9(2) | 2.04 | 4.01 | 2.01 | 3.86 | ||||||
kAu-Lb / 102 N m–1 | 2.0(1) | 1.0(1) | 1.95 | 1.00 | 2.06 | 1.41 |
The EPD spectra of and are compared in Figure 2 to the spectrum of Ar reported previously.20 shows a long progression starting at Ea = 7340(5) cm–1. Isolated peaks have a full width at half maximum (FWHM) of 5 cm–1, which corresponds to the laser bandwidth and allows to give a lower limit for the excited state lifetime. Together with the information that no delayed decay is observed in the reflectron after laser excitation, we can estimate that the excited state lifetime lies in between 2×10–6 and 1×10–12 s. About 13 quanta in the main progression (ν2 = 213 cm–1) are observed together with 2–3 quanta in a second mode (ν3 = 176 cm–1), as well as some peaks arising from hot bands. By comparison with the spectrum and the TD-DFT calculations, the progressions are assigned to the symmetric and antisymmetric stretch vibrations ν3 and ν2, their hot bands, and some small contributions from hot bands of the degenerate ν4 mode (SOMO ← HOMO-1, Figure 3 and Table S2). A Dunham fit yields harmonic frequencies and (cross) anharmonicities of the à state (Table 1). The A state of is formed by moving an electron from the antibonding 5dσ∗ orbital into the half-filled bonding 6sσ orbital (Figure 1), effectively increasing the bond order from 0.5 to 1.5. The resulting à states of L, therefore, show a 0.15 Å shorter Au-Au bond length than the state. At the same time, the -L bond strength decreases, resulting in an increase in intermolecular distance (+0.2 and +0.15 Å for L = N2 and N2O). The harmonic frequencies and atomic masses can be used to estimate the force constants of the Au-Au and the Au-N2 bonds (Table 1) as described in the Supporting Information. The Au-Au force constant increases by a factor of 1.7(2) upon excitation, consistent with an increased bond order. The -N2 force constant halves upon excitation. Both changes result in substantial geometry changes that are reflected in extended Franck–Condon (FC) progressions of the vibration corresponding to this change, which is mainly the antisymmetric Au-Au-L stretch mode (ν2 for and ν3 for ) and to a lesser extent the symmetric stretch (ν3 for and ν4 for , Figure 1).
[IMAGE OMITTED. SEE PDF]
Comparison of the simulated FC spectrum of to the measured EPD spectrum shows good agreement in the main progression (ν2) after correction for anharmonicity (Figure 3) and a small origin shift of –0.065 eV. The calculated geometry change is slightly overestimated, resulting in larger intensity toward the high-energy side of the spectrum. The simulations also slightly overestimate the contribution from ν3. The observed hot bands stem from ν3 and ν4. We observe the 2n31m transitions with m = 0 and 1 and the 2n411 transitions, from which we determine ω3'' = 143(2) cm–1 and Δ(ω4'', ω4') = 39(2) cm–1. These compare well to the computed values of 138 and 38 cm–1, respectively. Interestingly, we do not observe a hot band in the lowest energy vibration, ν5(π). If this made a substantial contribution to the spectrum, the peaks would be broader because a difference of 5 cm–1 between ω5' and ω5'' is at the lower end of the instrumental resolution. Possibly, lower frequency modes cool more efficiently than those with higher energy, an effect also seen in the Ar spectra,20 but not in the spectra of . Comparison between EPD spectra of recorded at 100 and 300 K (Figure S1), indeed, reveals significant broadening, albeit not as much as would be expected at 300 K. Another explanation is that Δ(ω5'', ω5') is significantly smaller than the calculated value of 5 cm–1. Significantly, the whole NIR spectrum of lies above its D0 value and thus is obtainable via single-photon EPD. A Birge–Sponer plot of the main progression (ν2) of N2 (Figure S2) results in ω2 = 216.3 cm–1 (the value given in Table 1 comes from the Dunham expansion) and D0 = 1.5 eV. The ω2 frequency agrees well with calculations but D0 is overestimated probably due to coupling of the Au-Au and Au-L local modes and strong deviation from the Morse potential assumed (Figure S3).20
The EPD spectrum of in Figure 2 is not as easily assigned as that of N2. It shows fewer and broader peaks (FWHM ∼30 cm–1) with lower peak intensity, and a clear band origin cannot be identified. It is noteworthy that the shape of the first peaks changes with nozzle temperature (Figures 2 and S1). The calculated –N2O dissociation energy (D0 = 1.06 eV, 8550 cm–1) occurs within the observed band system, and several experimental observations confirm this result. First, the pronounced vibronic progression is significantly shorter than for N2. Second, its total integrated intensity is significantly weaker, despite the same calculated oscillator strength, indicating that the low-frequency part of the absorption spectrum is not observed by EPD. This could be because the ligand binding energy exceeds the photon energy and thus prevents single-photon EPD from vibrationally cold ions. Third, the strong temperature dependence near the onset of the progression suggests that we observe mainly hot bands in this range. These results allow us to determine the experimental dissociation energy to D0 = 1.05(3) eV, lying somewhere between the temperature-dependent peak at 8486 cm–1 and the lowest observable signal at around 8200 cm–1, in excellent agreement with the predicted value (1.06 eV, 8550 cm–1). The band origin (Ea) of the à ← transition cannot be determined as precisely. However, an estimate can be made by comparing to the L spectra with L = Ar and N2. Their observed Ea values lie 523 and 610 cm–1 below the calculated energies. Assuming a similar difference for , one can compare the vibrational progressions of the two complexes by plotting the trace such that both calculated Ea’s overlap (blue trace in Figure S5, the 178 cm–1 shift is the difference in Ea’s). One can see that the progression then resembles the last part of the progression. An even better agreement between the two spectra is achieved by shifting the progression further (red trace in Figure 4 for a 691 cm–1 shift). We thus estimate Ea to lie between 7000 and 7500 cm–1 (i.e., Ea = 0.90(4) eV).
[IMAGE OMITTED. SEE PDF]
An FC simulation at 100 K is compared to the measured spectrum in Figure 4. The FC spectrum is shifted by –1060 cm–1, such that the relative peak heights at the higher energy side overlap with the experimental spectrum. However, because the calculation might overestimate the geometry change as in N2, this shift is somewhat arbitrary and a firm assignment of vibrational quanta has to remain open. Nevertheless, some safe conclusions can be drawn. The observed peaks stem from the main symmetric and antisymmetric stretch vibrations ν3 and ν4, in accordance with N2. Their harmonic frequencies are not as precisely determined as those for because of the missing initial part of the progression. Assuming a similar anharmonicity as that for (Figure S3) and a start of the observed progression at around v3 = 10–12, we obtain ω3 = 230(10) and ω4 = 170(10) cm–1. The larger width of the peaks compared to those of is explained by the fact that the lowest energy vibration (ν1) has a harmonic frequency of only ν1 = 26 cm–1 (compared to 44 cm–1 for N2) and is thus more strongly populated at the same internal temperature. Additionally, the difference in ω5'' and ω5' is probably larger than the calculated 1 cm–1. Using a Boltzmann distribution at 100 K, a difference in ω5'' and ω5' of –4(2) cm–1 can explain the observed line widths.
The binding energies of 0.89 and 1.06 eV computed for and in the state are relatively low compared to those for related complexes.1 As can be seen from the orbitals in Figure 1, a significant part of bonding comes from electron delocalization. In common with and , both complexes show a strong decrease in distance upon à ← excitation because the transition is mostly located on .20,21 Usually, one would use the shift of the stretch frequency to estimate a change in binding energy or force constant. Here, this is not easily possible because of very strong mixing with the stretch mode, which changes with ligand and electronic excitation. Considering the calculated distance, the bond becomes stronger upon excitation consistent with the increased bond order of . The ligand changes the situation somewhat. While the distance decreases in the state by ligation from 2.635 to 2.623, 2.606, and 2.608 Å in , , , and , respectively, the trend is less pronounced and reversed in the à state (2.447, 2.455, 2.459, and 2.459 Å). This result can be understood by considering that L donates electron density into the half-filled, bonding 6sσ orbital of , thereby stabilizing the dimer. In the à state, however, this orbital is filled and L cannot stabilize the bond further. Instead, any electron density transfer from L to either contributes to the partially occupied antibonding 5dσ∗ orbital or the empty 6sσ∗ orbital of , resulting in an increase in the distance upon ligation. This effect is weaker than the stabilizing effect in the state, probably due to the larger energy mismatch between the donating ligand orbitals and the antibonding orbitals. These interactions also explain the lower ligand binding energy in the à state compared to the state (Table 1).
We can also consider the impact of on the properties of the ligands. The calculated stretch frequency in free is the same as that in the state of (2501 cm–1). In the à state of , it is slightly higher at 2514 cm–1. A similar 17 cm–1 increase was observed in oxoacetate complexes and explained by delocalization of the antibonding 4σ*() orbital.27 The state wavenumber can be explained by the fact that only the nonbonding lone pair of participates in the bond. The small à state increase could be caused by the aforementioned delocalization or by the small induced dipole.
The computed N2O symmetric and antisymmetric stretch frequencies shift blue by almost equal amounts in the and à states of compared to those of bare N2O. A similar blue shift in the antisymmetric stretch of Au+N2OAr was explained by σ-donation from the antibonding 7σ* orbital of N2O to the metal center.24 Our calculations predict a small increase in the N-N distance upon ligation and a stronger decrease in the N-O distance in the complex compared to bare N2O. The calculated force constants for the N-N and N-O stretch vibrations increase by 6% and 12% upon ligation, respectively, showing that the strengths of both the N-N and the N-O bonds slightly increase but the effect is stronger on the N-O side. The difference in the N2O structure in the and à states of the complex is small and thus indicates that the binding mechanism of N2O to does not change upon excitation. This behavior can be explained in accordance with the case of Au+N2OAr.24 Electron density from the 7σ* N-N antibonding orbital of N2O is donated into the half-filled bonding σ orbital of . The 6σ* N-O antibonding orbital also loses electron density, probably into the 7σ N-N orbital, resulting in a strengthening of the N-N, N-O, and Au-Au bonds. Note that we do not observe the higher lying O-bound ON2 isomer23,24 and thus do not discuss it here.
Comparing the two ligands, both N2 and N2O have a similar effect on , that is, strengthening the Au-Au bond. The effect is strongest in the state of N2, resulting in an increase in the adiabatic excitation energy compared to Ar and despite an increase in binding energy in the and à states of Ar, to (Figure S4). The ligands, however, are affected differently. While the bond strength in N2 barely changes, the N-O and the N-N bonds in N2O are strengthened in both the and à states.
CONCLUDING REMARKS
In summary, gas-phase and complexes have been characterized by measuring their high-resolution vibronic à ← spectra occurring in the near infrared (NIR) range. The analysis provides unprecedented insight into the effects of electronic excitation on the chemical bonding in these prototypical metal–ligand complexes at the molecular level. An extended vibrational progression in the Au-Au-N2 antisymmetric stretch indicates that the Au-Au bond order increases markedly upon à ← excitation, while the Au-N2 bond length increases. For , the band origin of the à state lies below the dissociation energy of the state, providing an accurate experimental value for the important dissociation energy of D0 = 1.05(3) eV. Interestingly, the Au-Au bond is also strengthened upon ligation, in both cases due to electron density transfer into a bonding σ orbital of originating either from a lower lying antibonding orbital or from the ligand. Interestingly, the ligand effect is reversed in the excited state, because the bonding σ orbital is filled and electron transfer from the ligand occurs into antibonding orbitals, thereby, weakening the Au-Au bond upon ligation. The ligand binding energy decreases upon à ← excitation due to weaker bonding interactions in the excited state.
The N-N bond in N2 is nearly unaffected by binding to , either in the and à state of because the electron density donated to stems from a nonbonding N2 lone pair orbital. Although we have no direct experimental indication for back donation (we do not observe any internal ligand vibration), our calculations indicate that both N2O bonds in are strengthened. As a result, no molecular activation of N2 or N2O arises from binding to in either the or à state. In fact, the metal–ligand interaction in the à state is even weaker than in the state. The situation will be different for ligands that can readily donate electron density from bonding orbitals into the half-filled SOMO of or, when the à state is considered, into the energetically displaced antibonding orbitals of . To this end, this high-resolution spectroscopic approach will allow us to characterize reactive complexes, such as CnHm,10-14 providing detailed insight into photocatalysis of metal-organic reactions at the spectroscopic level.
AUTHOR CONTRIBUTIONS
Marko Förstel: conceptualization; formal analysis; investigation; methodology; project administration; software; supervision; validation; visualization; writing—original draft; writing—review and editing. Nima-Noah Nahvi: formal analysis; investigation; software; validation; writing—review and editing. Kai Pollow: conceptualization; formal analysis; investigation; software; validation; writing—review and editing. Taarna Studemund and Alice E. Green: investigation; writing—review and editing. André Fielicke: methodology; resources; writing—review and editing. Stuart R. Mackenzie: conceptualization; funding acquisition; methodology; writing—review and editing. Otto Dopfer: conceptualization; funding acquisition; investigation; methodology; project administration; resources; supervision; validation; writing—original draft; writing—review and editing.
ACKNOWLEDGMENTS
This project was supported by Deutsche Forschungsgemeinschaft (DFG, DO 729/9). We thank Roland Mitric for helpful discussions. Financial support is also gratefully acknowledged from the Oxford-Berlin Research Partnership (OXBER_STEM5: A Collaborative Approach to Understanding Nitrogen Oxide Reduction at Metal Centres). AEG and SRM gratefully acknowledge EPSRC programme grants EP/L005913 and EP/T021675. Corrections added on October 7, 2022 after first online publication: Ethic Statement was missing.
CONFLICT OF INTEREST
The authors declare no conflict of interest.
DATA AVAILABILITY STATEMENT
The data that support the findings of this study are available from the corresponding author upon reasonable request.
ETHICS STATEMENT
The authors confirm that they have followed the ethical policies of the journal.
PEER REVIEW
The peer review history for this article is available at
Schröder D, Schwarz H, Hrušák J, Pyykkö P. Cationic gold (I) complexes of xenon and of ligands containing the donor atoms oxygen, nitrogen, phosphorus, and sulfur. Inorg Chem. 1998;37:624‐632.
Tolman WB. Binding and activation of N2O at transition‐metal centers: recent mechanistic insights. Angew Chem Int Ed. 2010;49:1018–1024.
Luo ZX, Castleman AW, Khanna SN. Reactivity of metal clusters. Chem Rev. 2016;116:14456‐14492.
Schröder D, Shaik S, Schwarz H. Two‐state reactivity as a new concept in organometallic chemistry. Acc Chem Res. 2000;33:139‐145.
Whetten R, Cox D, Trevor D, Kaldor A. Correspondence between electron binding energy and chemisorption reactivity of iron clusters. Phys Rev Lett. 1985;54:1494.
Conceicao J, Laaksonen R, Wang L‐S, Guo T, Nordlander P, Smalley R. Photoelectron spectroscopy of transition‐metal clusters: correlation of valence electronic structure to reactivity. Phys Rev B. 1995;51:4668.
Berces A, Hackett P, Lian L, Mitchell S, Rayner D. Reactivity of niobium clusters with nitrogen and deuterium. J Chem Phys. 1998;108:5476‐5490.
Pyykkö P. Theoretical chemistry of gold. Angew Chem Int Ed. 2004;43:4412‐4456.
Shayeghi A, Johnston RL, Rayner DM, Schäfer R, Fielicke A. The nature of bonding between argon and mixed gold–silver trimers. Angew Chem Int Ed. 2015;54:10675‐10680.
Lang SM, Bernhardt TM, Bakker JM, Yoon B, Landman U. The interaction of ethylene with free gold cluster cations: infrared photodissociation spectroscopy combined with electronic and vibrational structure calculations. J Phys: Condens Mat. 2018;30: [eLocator: 504001].
Lang SM, Bernhardt TM, Chernyy V, Bakker JM, Barnett RN, Landman U. Selective C−H bond cleavage in methane by small gold clusters. Angew Chem Int Ed. 2017;56:13406‐13410.
Shuman NS, Ard SG, Sweeny BC, Viggiano AA, Owen CJ, Armentrout PB. Methane adducts of gold dimer cations: thermochemistry and structure from collision‐induced dissociation and association kinetics. J Phys Chem A. 2020;124:3335‐3346.
Shuman NS, Ard SG, Sweeny BC, et al. cannot catalyze conversion of methane to ethene at low temperature. Cat Sci Tech. 2019;9:2767‐2780.
Wang M‐M, Zhao Y‐X, Ding X‐L, Li W, He S‐G. Methane activation by heteronuclear diatomic AuRh+ cation: comparison with homonuclear and Rh2+. Phys Chem Chem Phys. 2020;22:6231‐6238.
Prather MJ. Time scales in atmospheric chemistry: coupled perturbations to N2O, NOy, and O3. Science. 1998;279:1339‐1341.
Ravishankara A, Daniel JS, Portmann RW. Nitrous oxide (N2O): the dominant ozone‐depleting substance emitted in the 21st century. Science. 2009;326:123‐125.
Chalkley MJ, Drover MW, Peters JC. Catalytic N2 to NH3 (or N2H4) conversion by well‐defined molecular coordination complexes. Chem Rev. 2020;120:5582‐5636.
Yandulov DV, Schrock RR. Catalytic reduction of dinitrogen to ammonia at a single molybdenum center. Science. 2003;301:76‐78.
Kuriyama S, Arashiba K, Nakajima K, et al. Catalytic transformation of dinitrogen into ammonia and hydrazine by iron‐dinitrogen complexes bearing pincer ligand. Nat Comm. 2016;7:1‐9.
Förstel M, Pollow K, Studemund T, Dopfer O. Near‐Infrared spectrum of the first excited state of . Chem Eur J. 2021;27:15075‐15080.
Förstel M, Pollow KM, Saroukh K, Najib EA, Mitric R, Dopfer O. The optical spectrum of . Angew Chem Int Ed. 2020;123:21587‐21592.
Dietrich G, Lützenkirchen K, Becker S, et al. Au‐induced decomposition of N2O. Ber Bunsen Gesell. 1994;98:1608‐1612.
Cunningham EM, Green AE, Meizyte G, et al. Infrared action spectroscopy of nitrous oxide on cationic gold and cobalt clusters. Phys Chem Chem Phys. 2021;23:329‐338.
Cunningham EM, Gentleman AS, Beardsmore PW, Iskra A, Mackenzie SR. Infrared signature of structural isomers of gas–phase M+(N2O)n (M= Cu, Ag, Au) ion–molecule complexes. J Phys Chem A. 2017;121:7565‐7571.
Hidai M, Mizobe Y. Recent advances in the chemistry of dinitrogen complexes. Chem Rev. 1995;95:1115‐1133.
Shan H, Yang Y, James AJ, Sharp PR. Dinitrogen bridged gold clusters. Science. 1997;275:1460‐1462.
Lang J, Mohrbach J, Dillinger S, Hewer JM, Niedner‐Schatteburg G. Vibrational blue shift of coordinated N2 in [Fe3O(OAc)6(N2)n]+: “non‐classical” dinitrogen complexes. Chem Comm. 2017;53:420‐423.
Dapprich S, Frenking G. Investigation of donor‐acceptor interactions: a charge decomposition analysis using fragment molecular orbitals. J Phys Chem. 1995;99:9352‐9362.
Lupinetti AJ, Frenking G, Strauss SH. Nonclassical metal carbonyls: appropriate definitions with a theoretical justification. Angew Chem Int Ed. 1998;37:2113‐2116.
Förstel M, Schewe W, Dopfer O. Optical spectroscopy of the Au4+ cluster: the resolved vibronic structure indicates an unexpected isomer. Angew Chem Int Ed. 2019;58:3356‐3360.
Förstel M, Jaeger BKA, Schewe W, Sporkhorst PHA, Dopfer O. Improved tandem mass spectrometer coupled to a laser vaporization cluster ion source. Rev Sci Instr. 2017;88: [eLocator: 123110].
Yanai T, Tew DP, Handy NC. A new hybrid exchange–correlation functional using the Coulomb‐attenuating method (CAM‐B3LYP). Chem Phys Lett. 2004;393:51‐57.
Woon DE, Dunning TH Jr. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J Chem Phys. 1993;98:1358‐1371.
Figgen D, Rauhut G, Dolg M, Stoll H. Energy‐consistent pseudopotentials for group 11 and 12 atoms: adjustment to multi‐configuration Dirac–Hartree–Fock data. Chem Phys. 2005;311:227‐244.
Frisch MJ, Trucks GW, Schlegel HB, et al. GAUSSIAN16, Rev. A.03. Wallingford, CT: Gaussian, Inc.; 2016.
Lofthus A, Krupenie PH. The spectrum of molecular nitrogen. J Phys Chem Ref Data. 1977;6:113‐307.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2023. This work is published under http://creativecommons.org/licenses/by/4.0/ (the "License"). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
The vibrationally resolved Ã2Σ+ ← Σ+ transitions of and are reported together with a detailed characterization of important geometric and electronic properties, enabling a deep understanding of the bonding mechanism at the molecular level. Comparison with time‐dependent density functional theory calculations reveals that the ligand stabilizes the entity in the Σ+ state by donating electron density into the half‐filled bonding orbital leading to the strengthening of the , , and bonds. This effect is reversed in the Ã2Σ+ state, where the bonding orbital is already filled and the ligand destabilizes the bond by donating into the antibonding orbitals of . The spectral detail obtained provides a deep understanding of the interplay of multiple electronic states in gas‐phase metal‐complex cations, opening the door for a systematic approach in the study of excited state reactivity in organometallic chemistry.
Key points
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details








1 Institut für Optik und Atomare Physik, Hardenbergstr. 36, Technische Universität Berlin, Berlin, Germany
2 Physical and Theoretical Chemistry Laboratory, Department of Chemistry, University of Oxford, Oxford, UK
3 Fritz‐Haber‐Institut der Max‐Planck‐Gesellschaft, Berlin, Germany