Next Article in Journal
Isolation of Mirificin and Other Bioactive Isoflavone Glycosides from the Kudzu Root Lyophilisate Using Centrifugal Partition and Flash Chromatographic Techniques
Previous Article in Journal
Therapeutic and Nutraceutical Effects of Polyphenolics from Natural Sources
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ionization of Decamethylmanganocene: Insights from the DFT-Assisted Laser Spectroscopy

1
G.A. Razuvaev Institute of Organometallic Chemistry of the Russian Academy of Sciences, 49 Tropinin St., 603950 Nizhny Novgorod, Russia
2
Institute of Atomic and Molecular Sciences, Academia Sinica, 1 Section 4, Roosevelt Road, Taipei 10617, Taiwan
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(19), 6226; https://doi.org/10.3390/molecules27196226
Submission received: 1 September 2022 / Revised: 15 September 2022 / Accepted: 17 September 2022 / Published: 22 September 2022
(This article belongs to the Section Organometallic Chemistry)

Abstract

:
Metallocenes represent one of the most important classes of organometallics with wide prospects for practical use in various fields of chemistry, materials science, molecular electronics, and biomedicine. Many applications of these metal complexes are based on their ability to form molecular ions. We report the first results concerning the changes in the molecular and electronic structure of decamethylmanganocene, Cp*2Mn, upon ionization provided by the high-resolution mass-analyzed threshold ionization (MATI) spectroscopy supported by DFT calculations. The precise ionization energy of Cp*2Mn is determined as 5.349 ± 0.001 eV. The DFT modeling of the MATI spectrum shows that the main structural deformations accompanying the detachment of an electron consist in the elongation of the Mn-C bonds and a change in the Me out-of-plane bending angles. Surprisingly, the DFT calculations predict that most of the reduction in electron density (ED) upon ionization is associated with the hydrogen atoms of the substituents, despite the metal character of the ionized orbital. However, the ED difference isosurfaces reveal a complex mechanism of the charge redistribution involving also the carbon atoms of the molecule.

1. Introduction

In 2022, chemists celebrate the 70th anniversary of the first publications on a unique sandwich structure of ferrocene, Cp2Fe (Cp = η5-C5H5) [1,2]. With the discovery of this iconic metal complex, a new era of organometallic chemistry has started. Currently, metallocenes are of ever-increasing interest among researchers in fundamental and applied science, since they are relevant to the modern theories of metal–ligand bonding [3,4,5,6], electro- and magnetochemistry [7,8,9,10,11,12,13], metallopolymers [14,15,16,17,18], biomedicine [19,20,21,22], catalysis [23,24,25], nanoelectronics [26,27,28,29,30,31,32] and molecular machines [33,34,35]. Many applications of metallocenes take advantage of their pronounced reducing properties, which allow for the easy formation of stable or reactive sandwich ions [7,8,9,10,14,15,16,23,24,25,26,27,28,29,30,33,34,35,36,37,38,39]. These properties are associated with the low ionization energies, I, corresponding to the detachment of an electron from a sandwich high occupied molecular orbital (MO) with substantial metal d character [40,41,42,43,44,45,46,47]. Comprehensive studies of ionization processes in metallocenes are, therefore, especially important for understanding their reactivity.
Unprecedented opportunities for studying the details of the ionization of molecules appeared with the development of high-resolution methods using laser irradiation. Precise measurements of ionization energies are provided by the zero kinetic energy (ZEKE) or mass-analyzed threshold ionization (MATI) spectroscopy [48,49] using the laser excitation of molecules to high Rydberg levels (so-called ZEKE states) which are then ionized by an electric pulse. The ZEKE states of polyatomic molecules with high I values are usually populated by multiphoton laser excitation through intermediate vibronic levels. For organometallics, this approach encounters experimental difficulties associated with poor resolution of vibronic states and fast photodissociation of the excited molecules [50,51]. However, the ZEKE states of metallocenes and related sandwich complexes with I < 6 eV can be populated in a one-photon experiment using standard laser sources.
Nevertheless, the experimental measurement of the MATI and ZEKE spectra of sandwich compounds remains a nontrivial task due to the low volatility and thermal stability of these metal complexes as well as their air and moisture sensitivity. Among metallocenes, the MATI spectrum has so far been recorded only for Cp2Co [52]. The MATI/ZEKE data have been obtained also for some bisarene complexes [53,54,55,56,57,58,59,60,61,62,63,64,65] and few mixed sandwiches [66]. These studies demonstrated new possibilities provided by the MATI and ZEKE techniques for detecting substituent effects and structural changes of sandwich molecules upon ionization [67]. In the present work, we report the first MATI spectrum of jet-cooled decamethylmanganocene, Cp*2Mn (Cp* = η5-C5Me5). The manganocene-methylated derivatives are of particular fundamental interest, since the introduction of the CH3 groups to the Cp rings of Cp2Mn leads to a “switching” of the ground electronic state from the 6A1 sextet to the 2E2 doublet [40,68,69,70,71,72,73] (we use the irreducible representations of the D5 point group for designation of the electronic states and MOs). In contrast to many other metallocenes, manganocene does not form stable cationic derivatives. On the other hand, its permethylated low-spin analogue, Cp*2Mn, is readily oxidized to yield the Cp*2Mn+ cation [11,73]. Such systems can display properties of bulk molecular ferromagnets [74,75]. Therefore, the details of structural and energetic changes accompanying the ionization of Cp*2Mn attract special attention. Our combined MATI/DFT study provides new insights into these details.

2. Results and Discussion

Similar to other theoretical studies of metallocenes [72,76,77,78,79], our B3PW91/6-311++G(d,p) DFT calculations show that high-energy occupied MOs of Cp*2Mn have a predominantly metal d character (Figure 1). The electronic configurations of the Cp*2Mn0 neutral and the Cp*2Mn+ cation are …(dx2-y2)1(dz2)2(dxy)2 and …(dx2-y2)1(dz2)1(dxy)2, respectively. These configurations correspond to the 2E2 doublet ground state of the D5 neutral and the 3E2 triplet ground state of the ion [40,71]. However, the degeneracy of MO dx2-y2 and dxy, which would take place in the D5 point group, is actually lifted in Cp*2Mn (Figure 1) due to the Jahn–Teller (JT) distortion.
The ground-state 3E2 Cp*2Mn+ cation is formed by a detachment of the non-bonding dz2 β-electron from the neutral molecule. The I value can be determined in a laser ionization experiment by recording the photoionization efficiency (PIE) curve or, much more accurately, by measuring the MATI spectrum.

2.1. PIE and MATI Spectra of Cp*2Mn

The PIE curve of Cp*2Mn (Figure 2a) shows a rather sharp step in the 43,000–43,200 cm−1 region corresponding to the ionization energy of 5.33–5.36 eV, which agrees very well with the I = 5.33 ± 0.10 eV value determined earlier from the “classical” photoelectron spectrum [71]. The MATI spectrum of Cp*2Mn (Figure 2b) reveals a clearly defined peak arising from a detachment of an electron from the ZEKE levels lying a few wavenumbers below the cation ground vibronic state. The I value of 43,146 ± 8 cm−1, or 5.349 ± 0.001 eV, corresponds to a narrow MATI signal descend range. Therefore, the MATI spectrum of Cp*2Mn provides an extremely accurate ionization energy, the error being two orders of magnitude smaller than that of standard photoelectron spectroscopy.
The precise I value of Cp*2Mn obtained from the MATI spectrum provides new accurate information on the difference between the Mn-Cp* bond dissociation energies in the decamethylmanganocene cation and neutral [67]:
D(Cp*2Mn +) − D(Cp*2Mn 0) = I(Mn0) − I(Cp*2Mn0),
where D(Cp*2Mn +/0) corresponds to the energy change in the reaction:
Cp*2Mn+/0 → Mn+/0 + 2C5Me5.
Taking into account the known ionization energy of the manganese atom, I(Mn0) = 4.43402 eV [80], and the I(Cp*2Mn0) value derived from our MATI spectrum we obtain the dissociation energy difference (Equation (1)) as 2.085 ± 0.001 eV, or 201.2 ± 0.1 kJ mol−1. Therefore, the mean Mn-Cp* bond dissociation energy in the decamethylmanganocene cation, equal to 0.5 D(Cp*2Mn +), is 100.6 kJ mol−1 larger than that in the neutral molecule. The accuracy in the determination of this difference from the MATI data presented in this work (0.1 kJ mol−1) is much better than that obtained from the UV photoelectron spectrum of Cp*2Mn [71] (ca. 10 kJ mol−1). As a result, the use of laser threshold ionization spectroscopy provides new possibilities for studying fine substituent effects in manganocenes, similar to those found for bisarene sandwich systems [67].
The MATI spectrum of Cp*2Mn (Figure 2b) reveals no lower-energy vibronic features accompanying the strong peak with the maximum at 43,100 cm−1. Therefore, this peak corresponds to the origin (the 000 transition between the ground vibrational states of Cp*2Mn0 and Cp*2Mn +). Then, the vertical and adiabatic ionization energies of Cp*2Mn coincide. A different situation holds with the MATI spectrum of Cp2Co [51], where the vertical I value is 0.115 eV higher than the adiabatic ionization energy. This difference can be associated with the antibonding nature of the ionized MO in Cp2Co [40,51,77]. On the other hand, the chromium bisarene complexes with the non-bonding dz2 electron ionized produce MATI spectra [52,53,54,55,56,58,64,65,66] showing the 000 vibronic component as the strongest feature like Cp*2Mn. The vibrationally resolved high-resolution MATI spectra provide much more information on the sandwich molecules than structureless “classical” photoelectron spectra or PIE curves.

2.2. Vibronic Structure of the Cp*2Mn MATI Spectrum

Detailed analysis of the Cp*2Mn MATI spectrum recorded at a higher concentration of the organometallic molecules in the supersonic jet reveals the weaker high-energy vibronic components accompanying the origin (Figure 3a). Their assignment can be made on the basis of the DFT modeling. The B3PW91/6-311++G(d,p) simulation (Figure 3b) gives a MATI vibronic structure that agrees very well with the experiment. The relative positions of the vibronic components (Table 1) correspond to the vibrational frequencies of the free Cp*2Mn+ cation. Only two normal vibrations of Cp*2Mn+ are responsible for the MATI features of noticeable intensity (Figure 3). They are denoted in this work as ν1 and ν2. According to the DFT calculations, the 130 cm−1 ν1 mode corresponds to the Me out-of-plane bend, while the 352 cm−1 ν2 vibration represents the symmetric Mn-ring stretch without displacement of the methyl carbon nuclei. Both the ν1 and ν2 modes involve the change of the Mn-Cring-CMe angle. The relative atomic shifts for the ν1 and ν2 vibrations are shown in Figure 4, and the corresponding animations are presented in the Supplementary Materials (Videos S1 and S2).
Both ν1 and ν2 would be totally symmetric modes in a D5 decamethylmetallocene molecule like closed-shell Cp*2Fe. The experimental and simulated MATI spectra of Cp*2Mn (Figure 3) do not contain vibronic components that could be associated with the JT distortion of the neutral or cationic species in the degenerate E2 electronic state. This testifies for similar shifts of the Cp*2Mn0 and Cp*2Mn+ molecular geometries from the D5 point group along the JT active vibrational coordinates. The good agreement between the experimental and simulated MATI vibronic structure (Figure 3, Table 1) shows that the B3PW91/6-311++G(d,p) level of DFT provides reliable information about the structural changes in Cp*2Mn upon ionization.

2.3. Changes in the Geometry of Decamethylmanganocene upon Ionization

The structural transformations accompanying ionization of Cp*2Mn can be analyzed on the basis of the neutral and ion geometries obtained from the X-ray diffraction experiments [68,74]. However, in this case the changes caused by the intramolecular effects would be superimposed on those resulted from the intermolecular and cation–anion interactions. The Cp*2Mn structure in the gas phase has been determined by electron diffraction [69], but these data provide only averaged geometry of the sandwich molecule corresponding to the high symmetry of the D5d point group. Therefore, the DFT calculations verified by the simulation of the experimental gas-phase MATI spectrum would be preferable as a basis for studying the fine details of the structural changes accompanying ionization of the Cp*2Mn molecule.
The atom numbering as well as selected bond lengths and angles in one Cp*Mn fragment are given in Figure 5 and Figure 6. The atomic coordinates are provided in the Supplementary Materials (Tables S1 and S2). The geometry of the other fragment is identical since the molecule with a staggered conformation of the Cp* ligands possesses a center of inversion. The optimized staggered geometry of Cp*2Mn0 is in accord with the X-ray structure [68,81] (Tables S3 and S4, see the Supplementary Materials). The calculated interatomic distances in Cp*2Mn0 (Figure 5a) agree well with the single-crystal X-ray diffraction data [68] (r(C-C) = 1.409–1.434 Å for the carbocycle and r(Mn-C) = 2.105–2.118 Å) and gas-phase electron diffraction results [69] (1.440 and 2.130 Å, respectively). The DFT-based structural parameters of the Cp*2Mn+ ion (Figure 5b) also correlate with the crystal structure [74], though the experimental solid-phase parameters depend strongly on the counter ion.
The optimized structure of the neutral Cp*2Mn molecule or the corresponding cation is nearly symmetric relative to the MnC1C6 plane. Together with the inversion center at the Mn atom, this means that the molecular symmetry is close to the C2h point group. The deviation from the ideal D5d symmetry is caused by the JT distortion of the Cp*2Mn neutral or ion in the degenerate E2 ground electronic state. The JT distortion in both Cp*2Mn0 and Cp*2Mn+ includes mainly: (1) a slippage of the Cp* ligands leading to a shift of the Mn atom projection on the ring plane from the ring centroid and to different Mn-C bond lengths; (2) an in-plane ring deformation (a contraction of the C1-C2 and C1-C5 bonds, an elongation of the C3-C4 bond); (3) unequal bending of the methyl groups away from the metal atom (the C8 and C9 atoms lie at the longest distance from the ring plane). The first and the second trend were also revealed by the X-ray study [68].
It should be noticed that the JT distortion of Cp*2Mn0 differs from that in Cp2Co0 where the Cp rings become non-planar [51,82]. This difference arises from the different nature of the single-occupied MO in these sandwich molecules. In Cp*2Mn0, the unpaired electron occupies MO dx2-y2 belonging to the e2 set in a non-distorted molecule. In Cp2Co0, the single-occupied MO is derived from the Co 3dxz or 3dyz and Cp π(e1) wavefunctions.
The structural transformation of Cp*2Mn0 upon ionization consists mainly in lengthening the Mn-Cp* distances and changing the Me bending angles (Table 2). As a result of the electron detachment, the distance between the ligand ring planes increases by 0.048 Å, from 3.494 Å to 3.542 Å. This increase in the ligand–ligand distance exceeds that computed earlier for (η6-C6H6)2Cr (0.038 Å [66]). The symmetry of the Cp*2Mn0 neutral and Cp*2Mn+ ion is the same, so only totally symmetric vibrations ν1 and ν2 corresponding to the Me bend and Mn-Cp* stretch, respectively, are excited in the MATI experiment. In contrast to decamethylmanganocene, ionization of cobaltocene leads to the highly symmetric closed-shell Cp2Co+ ion. Accordingly, the MATI spectrum of Cp2Co reveals the e2 and e1 vibrational modes which are active in the JT effect and pseudo-JT effect occurring in the neutral molecule [51,83]. The ionized MO of Cp*2Mn0 represents an almost “pure” 3dz2 orbital of the Mn atom (Figure 1). However, the electron density (ED) analysis reveals a much more complex charge redistribution upon ionization of decamethylmanganocene than would be expected from the simple subtraction of the dz2 ED.

2.4. Charge Density Redistribution after Electron Detachment from Cp*2Mn0

The ED difference (EDD) isosurfaces (Figure 7) corresponding to the ED change on going from Cp*2Mn0 to Cp*2Mn+ at the geometry of the neutral (the vertical ionization) reveal areas of both a decrease and an increase in the ED. The ED gain near the metal is the result of charge transfer from the Cp* ligands to the Mn atom during ED relaxation after ionization. The 0.005 a.u. EDD isosurface (Figure 7a) shows the ED loss arising from the detachment of the Mn dz2 electron and from the consequent transfer of the ligand π-electron density to the Mn atom. The largest area of the positive EDD visualizes the acquisition of ED by the metal atom through the Mn(dxz,yz)/Cp(π) e1-occupied orbital. A smaller ED shift from C(Me) to C(ring) along the C-C σ-bonds is also observed (Figure 7a). The 0.003 a.u. and 0.001 EDD isosurfaces (Figure 7b) reveal an additional ED loss by the H atoms of the Me groups accompanied by a ED increase near C(Me) around the C-H bonds. The positive EDD region appears inside the carbocycles near the C-C σ-bonds.
The inhomogeneous ED redistribution upon ionization of decamethylmanganocene leads to nontrivial changes in the atomic charges. They were calculated in this work by integration of the ED over the atomic basins within the frames of the Quantum Theory of Atoms in Molecules (QTAIM) [84,85]. Due to the transfer of the ED from the ligands to the metal, after the removal of the dz2 electron, the Mn charge increases by only 0.198 a.u. on going from Cp*2Mn0 to Cp*2Mn+ (Table 3). Correspondingly, each Cp* ligand becomes less negative by 0.401 a.u. However, the carbon charges change very little (Table 3). Moreover, the C atoms of both carbocycles and methyl groups become more negative upon ionization of Cp*2Mn0. The total change in the C(ring) and C(Me) charges in the molecule is −0.044 and −0.037 a.u., respectively. The loss of the ED by the ligands is provided, therefore, entirely by the hydrogen atoms of the methyl groups. Indeed, the total increase of the H positive charges is 0.883 a.u. It is worth mentioning that the larger charge shift (0.044 a.u.) corresponds to the H atoms oriented away from the metal (Table 3). This agrees well with the EDD isosurfaces (Figure 7b,c) which demonstrate, however, that, despite the decisive role of H atoms in the total loss of the ED by the ligands during Cp*2Mn0 ionization, carbon atoms are also involved in the ED redistribution mechanism. The loss of the π-electron density by the carbocyles, which transfer a negative charge to the metal, is compensated by the ED gain through the C(ring)-C(Me) σ-bonds.

3. Materials and Methods

3.1. PIE and MATI Experiments

Decamethylmanganocene was purchased from Alfa Aesar and purified by sublimation in the PIE and MATI experiments. All manipulations with the air- and moisture-sensitive compound were carried out under an inert atmosphere in a glovebox. The MATI setup was described earlier [64,86]. The Cp*2Mn sample was placed into a stainless steel tube heated up to 170–180 °C. The sublimed vapor-phase organometallics diluted with helium (1.5 bar) was expanded into an evacuated chamber using a heated pulsed valve. The molecular beam selected by a skimmer was directed into a high-vacuum (10−8 Torr) ionization chamber where the Cp*2Mn neutrals were excited (MATI) or ionized (PIE) by the UV laser radiation. Tunable laser pulses in the 225–235 nm region were formed when doubling the dye-laser frequency (Lambda-Physik, Scanmate UV) by the BBO-I crystal. The Nd:YAG laser (Quanta-Ray PRO-190-10) was used for the pumping of the dye laser. The laser wavelengths were calibrated with a wavemeter (Coherent, WaveMaster). In the PIE experiment, the molecular ions formed by ionization of Cp*2Mn were detected by a microchannel plate particle detector of the time-of-flight mass spectrometer. The photoionization mass spectrum revealed an intense molecular ion signal at m/z = 325 a.u. (C20H30Mn+). The ion signal at variable laser wavelengths was collected and analyzed. In the MATI experiment, the neutrals were excited by a UV laser pulse to the long-lived ZEKE states. To remove the prompt ions, a pulsed electric field of ~1 V/cm was switched on about 20 ns after the laser pulses. Then, another pulse of the +200 V/cm electric field was applied with a 11 µs delay to ionize the ZEKE-state molecules. The obtained Cp*2Mn+ ions were detected and analyzed like those in the PIE measurements.

3.2. DFT Calculations and Electron Density Analysis

The geometries of the neutral and cationic decamethylmanganocene were optimized at the B3PW91/6-311++G(d,p) level of DFT. The combination of the hybrid B3PW91 functional [87,88] and triple-ζ split valence basis set 6-311++G(d,p) including diffuse and polarization functions on hydrogen and the heavy atoms [89,90] demonstrated earlier good performance when modeling the MATI spectra of bisarene complexes [64,65] and mixed sandwiches [66]. The experimental X-ray structure [68] was used as a starting geometry in the optimization procedure. Harmonic vibrational frequencies were calculated and the results were used to compute the Franck-Condon factors and produce the model MATI spectra. No imaginary frequencies were found. The calculations were carried out with the Gaussian 16 program package [91]. The QT AIM analysis of the ED distribution was performed with the AIM ALL suite [92]. The atomic charges were obtained by integration of ED over the atomic basins using the Promega1 algorithm [93]. The EDD isosurfaces were built as a result of the neutral ED subtraction from the ion ED at the ground-state geometry using the Multiwfn code [94].

4. Conclusions

The high-resolution MATI spectroscopy provides unprecedented accuracy in determination of the ionization energy of decamethylmanganocene. The I value of 43,146 ± 8 cm−1, or 5.349 ± 0.001 eV is obtained on the basis of the Cp*2Mn MATI spectrum. The adiabatic and vertical ionization energies of Cp*2Mn coincide. The precise I value makes it possible to determine with high accuracy the difference between the Mn-Cp* mean bond energies in the Cp*2Mn+ cation and Cp*2Mn0 neutral (100.6 ± 0.1 kJ mol−1). The bond strengthening in the cation can be associated with a weaker repulsion between the less negative Cp* ligands and stronger Mn-Cp* covalent interactions that resulted from the increase in the dxz,yz electron density. The B3PW91/6-311++G(d,p) model of the MATI vibronic structure agrees very well with the experiment. Although the geometry of both Cp*2Mn0 and Cp*2Mn+ is distorted because of the JT effect, only the symmetric Cp*-Mn-Cp* stretching vibration and Mn-C(ring)-C(Me) bending mode are detected in the MATI structure. Correspondingly, the main structural deformations accompanying the electron detachment include an elongation of the Mn-C bonds and changes in the Me out-of-plane bending angle. Interestingly, atomic charge calculations using the QTAIM method predict that the largest part of the reduction in ED during ionization is associated with substituent hydrogen atoms, despite the metallic nature of the ionized orbital. The ED difference isosurfaces reveal, however, a complex mechanism of the charge redistribution involving also the carbon atoms of the molecule. Thus, DFT-assisted MATI spectroscopy represents a powerful instrument that unveils the details of the structural and electronic transformations accompanying ionization of the permethylated metallocene molecule.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27196226/s1, Table S1: Atomic coordinates of the optimized Cp*2Mn0 molecule; Table S2: Atomic coordinates of the optimized Cp*2Mn+ cation; Table S3: Calculated and experimental selected bond lengths in the Cp*2Mn0 molecule; Table S4: Calculated and experimental selected bond angles in the Cp*2Mn0 molecule; Video S1: Animated ν1 vibration of Cp*2Mn+; Video S2: Animated ν2 vibration of Cp*2Mn+. References [68,81] are cited in the supplementary materials.

Author Contributions

Conceptualization, S.K.; methodology, S.K. and W.-B.T.; formal analysis, E.R. and S.-Y.T.; investigation, S.-Y.T.; writing—original draft preparation, S.K.; writing—review and editing, W.-B.T. All authors have read and agreed to the published version of the manuscript.

Funding

The PIE and MATI experiments were supported by grants from the Russian Foundation for Basic Research and the Ministry of Science and Technology of Taiwan (RFBR Project No. 15-53-52008 and MOST Grant No. 110-2113-M-001-032). The DFT studies and ED analysis were funded by the Russian Science Foundation (Project No. 18-13-00356).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data reported in this study are available in the present article and in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

The sample of Cp*2Mn is available from commercial sources.

References

  1. Wilkinson, G.; Rosenblum, M.; Whiting, M.C.; Woodward, R.B. The Structure of Iron Bis-Cyclopentadienyl. J. Am. Chem. Soc. 1952, 74, 2125–2126. [Google Scholar] [CrossRef]
  2. Fischer, E.O.; Pfab, W. Cyclopentadien-Metallkomplexe, Ein Neuer Typ Metallorganischer Verbindungen. Z. Naturforsch. 1952, 7, 377–379. [Google Scholar] [CrossRef]
  3. Elschenbroich, C. Organometallics, 3rd ed.; Wiley-VCH: Weinheim, Germany, 2006; pp. 528–549. [Google Scholar]
  4. Frenking, G. Chemical Bonding in Transition Metal Compounds. In The Chemical Bond: Chemical Bonding across the Periodic Table; Frenking, G., Shaik, S., Eds.; Wiley-VCH: Berlin/Weinheim, Germany, 2014; Chapter 7; pp. 175–217. [Google Scholar]
  5. Cloke, F.G.N.; Green, J.C.; Kilpatrick, A.F.R.; O’Hare, D. Bonding in Pentalene Complexes and Their Recent Applications. Coord. Chem. Rev. 2017, 344, 238–262. [Google Scholar] [CrossRef]
  6. Gritzner, G. Reference Redox Systems in Nonaqueous Systems and the Relation of Electrode Potentials in Nonaqueous and Mixed Solvents to Standard Potentials in Water. In Handbook of Reference Electrodes; Inzelt, G., Lewenstam, A., Scholz, F., Eds.; Springer: Berlin/Heidelberg, Germany, 2013; Chapter 2; pp. 25–31. [Google Scholar] [CrossRef]
  7. Jörissen, J.; Speiser, B. Preparative Electrolysis on the Laboratory Scale. In Organic Electrochemistry: Revised and Expanded, 5th ed.; Hammerich, O., Speiser, B., Eds.; CRC Press: Boca Raton, FL, USA, 2015; Chapter 7; pp. 263–330. [Google Scholar] [CrossRef]
  8. Ding, Y.; Zhang, C.; Zhang, L.; Zhou, Y.; Yu, G. Molecular Engineering of Organic Electroactive Materials for Redox Flow Batteries. Chem. Soc. Rev. 2018, 47, 69–103. [Google Scholar] [CrossRef] [PubMed]
  9. Liu, K.; Liu, Y.; Lin, D.; Pei, A.; Cui, Y. Materials for Lithium-Ion Battery Safety. Sci. Adv. 2018, 4, eaas9820. [Google Scholar] [CrossRef]
  10. Yee, G.T.; Miller, J.S. Metallocene-Based Magnets. In Magnetism: Molecules to Materials V; Miller, J.S., Drillon, M., Eds.; Wiley-VCH: Weinheim, Germany, 2005; Chapter 7; pp. 223–260. [Google Scholar]
  11. Cirera, J.; Ruiz, E. Electronic and Steric Control of the Spin-Crossover Behavior in [(CpR)2Mn] Manganocenes. Inorg. Chem. 2018, 57, 702–709. [Google Scholar] [CrossRef]
  12. Shuku, Y.; Hirai, Y.; Semenov, N.A.; Kadilenko, E.; Gritsan, N.P.; Zibarev, A.V.; Rakitin, O.A.; Awaga, K. 3D Molecular Network And Magnetic Ordering, Formed By Multi-Dentate Magnetic Couplers, Bis(Benzene)Chromium(I) and [1,2,5]Thiadiazolo[3,4-c][1,2,5]Thiadiazolidyl. Dalton Trans. 2018, 47, 9897–9902. [Google Scholar] [CrossRef]
  13. Ormaza, M.; Abufager, P.; Verlhac, B.; Bachellier, N.; Bocquet, M.-L.; Lorente, N.; Limot, L. Controlled Spin Switching in a Metallocene Molecular Junction. Nat. Commun. 2017, 8, 1974. [Google Scholar] [CrossRef] [Green Version]
  14. Braunschweig, H.; Damme, A.; Demeshko, S.; Dück, K.; Kramer, T.; Krummenacher, I.; Meyer, F.; Radacki, K.; Stellwag-Konertz, S.; Whittell, G.R. A Paramagnetic Heterobimetallic Polymer: Synthesis, Reactivity, and Ring-Opening Polymerization of Tin-Bridged Homo- and Heteroleptic Vanadoarenophanes. J. Am. Chem. Soc. 2015, 137, 1492–1500. [Google Scholar] [CrossRef]
  15. Musgrave, R.A.; Russell, A.D.; Hayward, D.W.; Whittell, G.R.; Lawrence, P.G.; Gates, P.J.; Green, J.C.; Manners, I. Main-chain Metallopolymers at the Static–Dynamic Boundary Based on Nickelocene. Nat. Chem. 2017, 9, 743–750. [Google Scholar] [CrossRef]
  16. Astruc, D. Why is Ferrocene so Exceptional? Eur. J. Inorg. Chem. 2017, 2017, 6–29. [Google Scholar] [CrossRef]
  17. Eloi, J.-C.; Chabanne, L.; Whittell, G.R.; Manners, I. Metallopolymers with Emerging Applications. Mater. Today 2008, 11, 28–36. [Google Scholar] [CrossRef]
  18. Wang, Y.; Astruc, D.; Abd-El-Aziz, A.S. Metallopolymers for Advanced Sustainable Applications. Chem. Soc. Rev. 2019, 48, 558–636. [Google Scholar] [CrossRef] [PubMed]
  19. Van Staveren, D.R.; Metzler-Nolte, N. Bioorganometallic Chemistry of Ferrocene. Chem. Rev. 2004, 104, 5931–5986. [Google Scholar] [CrossRef]
  20. Ihara, T. Oranometallic Complexes for Biosensing. In Advances in Bioorganometallic Chemistry, 1st ed.; Hirao, T., Moriuchi, T., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; Chapter 14; pp. 277–303. [Google Scholar]
  21. Patra, M.; Gasser, G. The medicinal chemistry of ferrocene and its derivatives. Nat. Rev. Chem. 2017, 1, 0066. [Google Scholar] [CrossRef]
  22. Albada, B.; Metzler-Nolte, N. Organometallic–Peptide Bioconjugates: Synthetic Strategies and Medicinal Applications. Chem. Rev. 2016, 116, 11797–11839. [Google Scholar] [CrossRef]
  23. Astruc, D. Organometallic Chemistry and Catalysis; Springer: Berlin/Heidelberg, Germany, 2007; pp. 251–288. [Google Scholar]
  24. Collins, S.; Linnolahti, M. Activation of Substituted Metallocene Catalysts Using Methylaluminoxane. ChemCatChem 2022, 14, e202101918. [Google Scholar] [CrossRef]
  25. Bochmann, M. Organometallics and Catalysis: An Introduction; Oxford University Press: Oxford, UK, 2015; 432p. [Google Scholar]
  26. Liu, R.; Ke, S.-H.; Baranger, H.U.; Yang, W. Organometallic Spintronics: Dicobaltocene Switch. Nano Lett. 2005, 5, 1959–1962. [Google Scholar] [CrossRef]
  27. Engtrakul, C.; Sita, L.R. Ferrocene-Based Nanoelectronics: Regioselective Syntheses and Electrochemical Characterization of α-Monothiol and α,ω-Dithiol, Phenylethynyl-Conjugated, 2,5-Diethynylpyridyl- and Pyridinium-Linked Diferrocene Frameworks Having an End-to-End Distance of ∼4 nm. Organometallics 2008, 27, 927–937. [Google Scholar] [CrossRef]
  28. Morari, C.; Rungger, I.; Rocha, A.R.; Sanvito, S.; Melinte, S.; Rignanese, G.-M. Electronic Transport Properties of 1,1′-Ferrocene Dicarboxylic Acid Linked to Al(111) Electrodes. ACS Nano 2009, 3, 4137–4143. [Google Scholar] [CrossRef]
  29. Lawson, B.; Zahl, P.; Hybertsen, M.S.; Kamenetska, M. Formation and Evolution of Metallocene Single-Molecule Circuits with Direct Gold-π Links. J. Am. Chem. Soc. 2022, 144, 6504–6515. [Google Scholar] [CrossRef] [PubMed]
  30. Inkpen, M.S.; White, A.J.P.; Albrecht, T.; Long, N.J. Avoiding Problem Reactions at the Ferrocenyl-Alkyne Motif: A Convenient Synthesis of Model, Redox-Active Complexes for Molecular Electronics. Dalton Trans. 2014, 43, 15287–15290. [Google Scholar] [CrossRef]
  31. Wilson, L.E.; Hassenrück, C.; Winter, R.F.; White, A.J.P.; Albrecht, T.; Long, N.J. Ferrocene- and Biferrocene-Containing Macrocycles towards Single-Molecule Electronics. Angew. Chem. Int. Ed. Engl. 2017, 56, 6838–6842. [Google Scholar] [CrossRef]
  32. Ghosh, A.; Mishra, S.; Giri, S.; Mobin, S.M.; Bera, A.; Chatterjee, S. Electrolyte-Free Dye-Sensitized Solar Cell with High Open Circuit Voltage Using a Bifunctional Ferrocene-Based Cyanovinyl Molecule as Dye and Redox Couple. Organometallics 2018, 37, 1999–2002. [Google Scholar] [CrossRef]
  33. Scottwell, S.O.; Elliott, A.B.S.; Shaffer, K.J.; Nafady, A.; McAdam, C.J.; Gordon, K.C.; Crowley, J.D. Chemically and Electrochemically Induced Expansion and Contraction of a Ferrocene Rotor. Chem. Commun. 2015, 51, 8161–8164. [Google Scholar] [CrossRef] [PubMed]
  34. Scottwell, S.O.; Crowley, J.D. Ferrocene-Containing Non-Interlocked Molecular Machines. Chem. Commun. 2016, 52, 2451–2464. [Google Scholar] [CrossRef]
  35. Scottwell, S.O.; Barnsley, J.E.; McAdam, C.J.; Gordon, K.C.; Crowley, J.D. A Ferrocene Based Switchable Molecular Folding Ruler. Chem. Commun. 2017, 53, 7628–7631. [Google Scholar] [CrossRef]
  36. Tritto, I.; Sacchi, M.C.; Locatelli, P.; Li, S.X. Metallocene ion pairs: A direct insight into the reaction equilibria and polymerization by 13C NMR spectroscopy. Macromol. Symp. 1995, 89, 289–298. [Google Scholar] [CrossRef]
  37. Abel, E.W.; Stone, F.G.A.; Wilkinson, G. Comprehensive Organometallic Chemistry II: A Review of the Literature 1982–1994, 1st ed.; Pergamon: Oxford, UK, 1995; Volume 5–9. [Google Scholar]
  38. Deck, P.A.; Marks, T.J. Cationic Metallocene Olefin Polymerization Catalysts. Thermodynamic and Kinetic Parameters for Ion Pair Formation, Dissociation, and Reorganization. J. Am. Chem. Soc. 1995, 117, 6128–6129. [Google Scholar] [CrossRef]
  39. Bochmann, M. Cationic Group 4 Metallocene Complexes and Their Role in Polymerisation Catalysis: The Chemistry of Well Defined Ziegler Catalysts. J. Chem. Soc. Dalton Trans. 1996, 3, 255–270. [Google Scholar] [CrossRef]
  40. Green, J.C. Gas Phase Photoelectron Spectra of d- and f-Block Organometallic Compounds. Struct. Bond. 1981, 43, 37–112. [Google Scholar]
  41. Rabalais, J.W.; Werme, L.O.; Bergmark, T.; Karlsson, L.; Hussain, M.; Siegbahn, K. Electron Spectroscopy of Open-Shell Systems: Spectra of Ni(C5H5)2, Fe(C5H5)2, Mn(C5H5)2, and Cr(C5H5)2. J. Chem. Phys. 1972, 57, 1185–1192. [Google Scholar] [CrossRef]
  42. Green, J.C.; Green, M.L.H.; Kaltsoyannis, N.; Mountford, P.; Scott, P.; Simpson, S.J. Synthesis and Electronic and Molecular Structures Of η-Cycloheptatrienyl η-Cyclopentadienyl Derivatives of Vanadium, Niobium, and Tantalum: Photoelectron Spectroscopic, Electrochemical, and X-Ray Crystallographic Study. Organometallics 1992, 11, 3353–3361. [Google Scholar] [CrossRef]
  43. Ketkov, S.Y.; Domrachev, G.A. Rydberg Transitions in the Near-Ultraviolet Vapor-Phase Absorption Spectra of 3d Metallocenes. Inorg. Chim. Acta 1990, 178, 233–242. [Google Scholar] [CrossRef]
  44. Ketkov, S.Y. Photoabsorption Study of Electronic Excited States of (η5-Cyclopentadienyl)(η6-Arene)Manganese. J. Organomet. Chem. 1994, 465, 225–231. [Google Scholar] [CrossRef]
  45. Révész, A.; Szepes, L.; Baer, T.; Sztáray, B.J. Binding Energies and Isomerization in Metallocene Ions from Threshold Photoelectron Photoion Coincidence Spectroscopy. J. Am. Chem. Soc. 2010, 132, 17795–17803. [Google Scholar] [CrossRef]
  46. Green, J.C. Periodic Trends Revealed by Photoelectron Studies of Transition Metal and Lanthanide Compounds. Struct. Bond. 2019, 181, 81–110. [Google Scholar] [CrossRef]
  47. Green, J.C.; Decleva, P. Photoionization cross-sections: A guide to electronic structure. Coord. Chem. Rev. 2005, 249, 209–228. [Google Scholar] [CrossRef]
  48. Schlag, E.W. ZEKE Spectroscopy; Cambridge University Press: Cambridge, UK, 1998; 277p. [Google Scholar]
  49. Yang, D.-S. Photoelectron spectra of metal-containing molecules with resolutions better than 1 meV. Coord. Chem. Rev. 2001, 214, 187–213. [Google Scholar] [CrossRef]
  50. Clara, M.; Neusser, H.J. Ultrafast Dissociation Dynamics of Ferrocene and [3]-Ferrocenophan. J. Phys. Chem. A 2001, 105, 5577–5585. [Google Scholar] [CrossRef]
  51. Ketkov, S.Y.; Selzle, H.L. Threshold Ionization of Cobaltocene: The Metallocene Molecule Revealing Zero Kinetic Energy States. Angew. Chem. Int. Ed. 2012, 51, 11527–11530. [Google Scholar] [CrossRef] [PubMed]
  52. Ketkov, S.Y.; Selzle, H.L.; Schlag, E.W. Study of Jet-Cooled Bis(η6-Benzene)Chromium by Multiphoton Zero Kinetic Energy–Mass-Analyzed Threshold Ionization Spectroscopy. Isr. J. Chem. 2004, 44, 65–69. [Google Scholar] [CrossRef]
  53. Ketkov, S.Y.; Selzle, H.L.; Schlag, E.W. High-Resolution Mass-Analyzed Threshold Ionization Study of Deuterated Derivatives of Bis(η6-Benzene)Chromium. Organometallics 2006, 25, 1712–1716. [Google Scholar] [CrossRef]
  54. Choi, K.-W.; Choi, S.; Baek, S.J.; Kim, S.K. Pulsed-Field Ionization Spectroscopy of High Rydberg States (n = 50–200) of Bis(η6-Benzene)Chromium. J. Chem. Phys. 2007, 126, 034308. [Google Scholar] [CrossRef]
  55. Ketkov, S.Y.; Selzle, H.L.; Schlag, E.W. Nanosecond Multiphoton Ionization Spectroscopy of Transition-Metal Sandwich Compounds: A Comparative Study of Nickelocene, Ferrocene and Bis(η6-Benzene)Chromium. Mol. Phys. 2004, 102, 1749–1757. [Google Scholar] [CrossRef]
  56. Choi, K.-W.; Kim, S.K.; Ahn, D.-S.; Lee, S. One-Photon Mass-Analyzed Threshold Ionization Spectroscopy of Bis(η6-Benzene)Chromium and Its Benzene and Ar Clusters. J. Phys. Chem. A 2004, 108, 11292–11295. [Google Scholar] [CrossRef]
  57. Sohnlein, B.R.; Yang, D.-S. Pulsed-Field Ionization Electron Spectroscopy of Group 6 Metal (Cr, Mo, and W) Bis(Benzene) Sandwich Complexes. J. Chem. Phys. 2006, 124, 134305. [Google Scholar] [CrossRef]
  58. Ketkov, S.Y.; Selzle, H.L.; Cloke, F.G.N. Direct Detection of Individual Bis(arene) Rotational Isomers in the Gas Phase by Mass-Analyzed Threshold Ionization Spectroscopy. Angew. Chem. Int. Ed. 2007, 46, 7072–7074. [Google Scholar] [CrossRef]
  59. Sohnlein, B.R.; Lei, Y.; Yang, D.-S. Electronic States of Neutral and Cationic Bis(Benzene) Titanium and Vanadium Sandwich Complexes Studied by Pulsed Field Ionization Electron Spectroscopy. J. Chem. Phys. 2007, 127, 114302. [Google Scholar] [CrossRef]
  60. Lee, J.S.; Kumari, S.; Yang, D.-S. Conformational Isomers and Isomerization of Group 6 (Cr, Mo, and W) Metal-Bis(Toluene) Sandwich Complexes Probed by Variable-Temperature Electron Spectroscopy. J. Phys. Chem. A. 2010, 114, 11277–11284. [Google Scholar] [CrossRef]
  61. Ketkov, S.Y.; Selzle, H.L.; Cloke, F.G.N.; Markin, G.V.; Schevelev, Y.A.; Domrachev, G.A.; Schlag, E.W. Zero Kinetic Energy Spectroscopy: Mass-Analyzed Threshold Ionization Spectra of Chromium Sandwich Complexes with Alkylbenzenes, (η6-RPh)2Cr (R = Me, Et, i-Pr, t-Bu). J. Phys. Chem. A 2010, 114, 11298–11303. [Google Scholar] [CrossRef] [PubMed]
  62. Yang, D.-S. High-Resolution Electron Spectroscopy of Gas-Phase Metal−Aromatic Complexes. J. Phys. Chem. Lett. 2011, 2, 25–33. [Google Scholar] [CrossRef] [PubMed]
  63. Kumari, S.; Yang, D.-S. High-Resolution Electron Spectroscopy and Rotational Conformers of Group 6 Metal (Cr, Mo, and W) Bis(mesitylene) Sandwich Complexes. J. Phys. Chem. A 2013, 117, 13336–13344. [Google Scholar] [CrossRef] [PubMed]
  64. Ketkov, S.Y.; Markin, G.V.; Tzeng, S.-Y.; Tzeng, W.-B. Fine Substituent Effects in Sandwich Complexes: A Threshold Ionization Study of Monosubstituted Chromium Bisarene Compounds. Chem. Eur. J. 2016, 22, 4690–4694. [Google Scholar] [CrossRef]
  65. Ketkov, S.Y.; Tzeng, S.-Y.; Wu, P.-Y.; Markin, G.V.; Tzeng, W.-B. DFT Supported Threshold Ionization Study of Chromium Biphenyl Complexes: Unveiling the Mechanisms of Substituent Influence on Redox Properties of Sandwich Compounds. Chem. Eur. J. 2017, 23, 13669–13675. [Google Scholar] [CrossRef]
  66. Ketkov, S.Y.; Tzeng, S.-Y.; Rychagova, E.A.; Markin, G.V.; Makarov, S.G.; Tzeng, W.-B. Laser Spectroscopic and Computational Insights into Unexpected Structural Behaviours of Sandwich Complexes on Ionization. Dalton Trans. 2021, 50, 10729–10736. [Google Scholar] [CrossRef]
  67. Ketkov, S.Y. Substituent Effects on the Electronic Structures of Sandwich Compounds: New Understandings Provided By DFT-Assisted Laser Ionization Spectroscopy of Bisarene Complexes. Dalton Trans. 2020, 49, 569–577. [Google Scholar] [CrossRef]
  68. Freyberg, D.P.; Robbins, J.L.; Raymond, K.N.; Smart, J.C. Crystal and Molecular Structures of Decamethylmanganocene and Decamethylferrocene. Static Jahn-Teller Distortion in a Metallocene. J. Am. Chem. Soc. 1979, 101, 892–897. [Google Scholar] [CrossRef]
  69. Fernholt, L.; Haaland, A.; Seip, R.; Robbins, J.L.; Smart, J.C. The Molecular Structure of Decamethylmanganocene, [η-C5(CH3)5]2Mn, Determined by Gas Phase Electron Diffraction. J. Organomet. Chem. 1980, 194, 351–355. [Google Scholar] [CrossRef]
  70. Walter, M.D.; Sofield, C.D.; Booth, C.H.; Andersen, R.A. Spin Equilibria in Monomeric Manganocenes: Solid-State Magnetic and EXAFS Studies. Organometallics 2009, 28, 2005–2019. [Google Scholar] [CrossRef]
  71. Cauletti, C.; Green, J.C.; Kelly, M.R.; Powell, P.; van Tilborg, J.; Robbins, J.; Smart, J. Photoelectron Spectra of Metallocenes. J. Electr. Spectrosc. Rel. Phenom. 1980, 19, 327–353. [Google Scholar] [CrossRef]
  72. Wang, G.; Slebodnick, C.; Butcher, R.J.; Tam, M.C.; Crawford, T.D.; Yee, G.T. A Family of Decamethylmetallocene Charge-Transfer Salt Magnets Using Methyl Tricyanoethylenecarboxylate (MTCE) as the Electron Acceptor. J. Am. Chem. Soc. 2004, 126, 16890–16895. [Google Scholar] [CrossRef]
  73. Robbins, J.L.; Edelstein, N.M.; Cooper, S.R.; Smart, J.C. Syntheses and Electronic Structures of Decamethylmanganocenes. J. Am. Chem. Soc. 1979, 101, 3853–3857. [Google Scholar] [CrossRef]
  74. Faulmann, C.; Riviere, E.; Dorbes, S.; Senocq, F.; Coronado, E.; Cassoux, P. Ferromagnetism in [Mn(Cp*)2]+-Derived Complexes: The “Miraculous” Stacking in [Mn(Cp*)2][Ni(dmit)2]. Eur. J. Inorg. Chem. 2003, 2003, 2880–2888. [Google Scholar] [CrossRef]
  75. Broderick, W.E.; Thompson, J.A.; Day, E.P.; Hoffman, B.M. A Molecular Ferromagnet with a Curie Temperature of 6.2 Kelvin: [Mn(C5(CH3)5)2]+[TCNQ]. Science 1990, 249, 401–403. [Google Scholar] [CrossRef]
  76. Clack, D.W.; Warren, K.D. MetaI-Ligand Bonding in 3 d Sandwich Complexes. Struct. Bond. 1980, 39, 1–41. [Google Scholar] [CrossRef]
  77. Xu, Z.-F.; Xie, Y.; Feng, W.-L.; Schaefer, H.F. Systematic Investigation of Electronic and Molecular Structures for the First Transition Metal Series Metallocenes M(C5H5)2 (M = V, Cr, Mn, Fe, Co, and Ni). J. Phys. Chem. A 2003, 107, 2716–2729. [Google Scholar] [CrossRef]
  78. Matsuzawa, N.; Seto, J.; Dixon, D.A. Density Functional Theory Predictions of Second-Order Hyperpolarizabilities of Metallocenes. J. Phys. Chem. A 1997, 101, 9391–9398. [Google Scholar] [CrossRef]
  79. Hrobárik, P.; Reviakine, R.; Arbuznikov, A.V.; Malkina, O.L.; Malkin, V.G.; Köhler, F.H.; Kaupp, M. Density Functional Calculations of NMR Shielding Tensors for Paramagnetic Systems with Arbitrary Spin Multiplicity: Validation On 3d Metallocenes. J. Chem. Phys. 2007, 126, 024107. [Google Scholar] [CrossRef]
  80. Lias, S.G.; Bartmess, J.E.; Liebman, J.F.; Holmes, J.L.; Levin, R.D.; Mallard, W.G. Ion Energetics Data. In NIST Chemistry WebBook, NIST Standard Reference Database Number 69; Linstrom, P.J., Mallard, W.G., Eds.; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2022. [Google Scholar] [CrossRef]
  81. Augart, N.; Boese, R.; Schmid, G. Phasenübergang von Decamethylmanganocen. Z. Anorg. Allg. Chem. 1991, 595, 27–34. [Google Scholar] [CrossRef]
  82. Zlatar, M.; Schläpfer, C.-W.; Fowe, E.P.; Daul, C. Density Functional Theory Study of the Jahn-Teller Effect in Cobaltocene. Pure Appl. Chem. 2009, 81, 1397–1411. [Google Scholar] [CrossRef]
  83. Ketkov, S.Y.; Rychagova, E.A.; Zhigulin, G.Y.; Tzeng, S.Y.; Tzeng, W.B. Quantum-Chemical Modeling of the Mass-Analyzed Threshold Ionization Spectra of Ferrocene and Cobaltocene. High Energ. Chem. 2020, 54, 414–420. [Google Scholar] [CrossRef]
  84. Bader, R.F.W. Atoms in Molecules: A Quantum Theory; Oxford University Press: Oxford, UK, 1990; 438p. [Google Scholar]
  85. Cortés-Guzmán, F.; Bader, R.F.W. Complementarity of QTAIM and MO theory in the study of bonding in donor-acceptor complexes. Coord. Chem. Rev. 2005, 249, 633–662. [Google Scholar] [CrossRef]
  86. Xu, Y.; Tzeng, S.Y.; Shivatare, V.; Takahashi, K.; Zhang, B.; Tzeng, W.B. Identification of Four Rotamers of M-Methoxystyrene by Resonant Two-Photon Ionization and Mass Analyzed Threshold Ionization Spectroscopy. J. Chem. Phys. 2015, 142, 124314. [Google Scholar] [CrossRef] [PubMed]
  87. Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  88. Perdew, J.P.; Burke, K.; Wang, Y. Generalized Gradient Approximation for the Exchange-Correlation Hole of a Many-Electron System. Phys. Rev. B Condens. Matter 1996, 54, 16533–16539. [Google Scholar] [CrossRef]
  89. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-Consistent Molecular Orbital Methods. XX. A Basis Set for Correlated Wave Functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  90. Clark, T.; Chandrasekhar, J.; Spitznagel, G.W.; Schleyer, P.v.R. Efficient Diffuse Function-Augmented Basis Sets for Anion Calculations. III. The 3-21+G Basis Set for First-Row Elements, Li-F. J. Comput. Chem. 1983, 4, 294–301. [Google Scholar] [CrossRef]
  91. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  92. Keith, T.A. AIMAll, Version 13.05.06; TK Gristmill Software: Overland Park, KS, USA, 2013. Available online: https://aim.tkgristmill.com(accessed on 20 September 2015).
  93. Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  94. Lu, T.; Chen, F. Quantitative Analysis of Molecular Surface Based on Improved Marching Tetrahedra Algorithm. J. Mol. Graph. Model. 2012, 38, 314–323. [Google Scholar] [CrossRef]
Figure 1. MO correlation diagram of the Cp*2Mn0 neutral and Cp*2Mn+ cation. The MO isosurfaces are given for the wavefunction value of 0.04. Notice the different energy scales for Cp*2Mn0 (left) and Cp*2Mn+ (right).
Figure 1. MO correlation diagram of the Cp*2Mn0 neutral and Cp*2Mn+ cation. The MO isosurfaces are given for the wavefunction value of 0.04. Notice the different energy scales for Cp*2Mn0 (left) and Cp*2Mn+ (right).
Molecules 27 06226 g001
Figure 2. The PIE (a) and MATI (b) spectra of jet-cooled Cp*2Mn. The ionization energy (I) corresponds to the narrow range of the MATI signal descend.
Figure 2. The PIE (a) and MATI (b) spectra of jet-cooled Cp*2Mn. The ionization energy (I) corresponds to the narrow range of the MATI signal descend.
Molecules 27 06226 g002
Figure 3. The experimental (a) and DFT-simulated (b) vibronic structure of the Cp*2Mn MATI spectrum.
Figure 3. The experimental (a) and DFT-simulated (b) vibronic structure of the Cp*2Mn MATI spectrum.
Molecules 27 06226 g003
Figure 4. Relative atomic shifts corresponding to the ν1 (a) and ν2 (b) vibrations of the Cp*2Mn+ cation.
Figure 4. Relative atomic shifts corresponding to the ν1 (a) and ν2 (b) vibrations of the Cp*2Mn+ cation.
Molecules 27 06226 g004
Figure 5. Mn-C(ring) and C(ring)-C(ring) distances (Å) in the optimized Cp*2Mn0 neutral (a) and Cp*2Mn+ cation (b). The hydrogen atoms are omitted for clarity.
Figure 5. Mn-C(ring) and C(ring)-C(ring) distances (Å) in the optimized Cp*2Mn0 neutral (a) and Cp*2Mn+ cation (b). The hydrogen atoms are omitted for clarity.
Molecules 27 06226 g005
Figure 6. Centroid(ring)-C(ring)-C(Me) angles in the optimized Cp*2Mn0 neutral (a) and Cp*2Mn+ cation (b). The hydrogen atoms are omitted for clarity.
Figure 6. Centroid(ring)-C(ring)-C(Me) angles in the optimized Cp*2Mn0 neutral (a) and Cp*2Mn+ cation (b). The hydrogen atoms are omitted for clarity.
Molecules 27 06226 g006
Figure 7. Isosurfaces of the electron density difference between the Cp*2Mn+ cation and neutral Cp*2Mn0 at the optimized geometry of the latter (vertical ionization of Cp*2Mn). The isovalues are 0.005 a.u. (a), 0.003 a.u. (b) and 0.001 a.u. (c). The blue and red colors correspond to the loss and gain of ED, respectively.
Figure 7. Isosurfaces of the electron density difference between the Cp*2Mn+ cation and neutral Cp*2Mn0 at the optimized geometry of the latter (vertical ionization of Cp*2Mn). The isovalues are 0.005 a.u. (a), 0.003 a.u. (b) and 0.001 a.u. (c). The blue and red colors correspond to the loss and gain of ED, respectively.
Molecules 27 06226 g007
Table 1. Relative positions (cm−1) of the vibronic components in the Cp*2Mn experimental MATI spectrum and its DFT simulation.
Table 1. Relative positions (cm−1) of the vibronic components in the Cp*2Mn experimental MATI spectrum and its DFT simulation.
Vibronic Component aMATIDFT
00000
101140 b130
102-261
201340352
201101480482
a The vibronic transitions are notated as Nv″v′, where N is the vibration number and v″ and v′ are the vibrational quantum number in the neutral and ionic state, respectively. b Shoulder.
Table 2. Calculated changes in selected bond lengths and atom-plane(ring) distances ∆r (Å) as well as angles ∆ϕ (°) upon ionization of Cp*2Mn0.
Table 2. Calculated changes in selected bond lengths and atom-plane(ring) distances ∆r (Å) as well as angles ∆ϕ (°) upon ionization of Cp*2Mn0.
Distance ΔrAngleΔϕ
C1-C20.004Mn-centroid(ring)-C1−0.1
C2-C30.002Mn-centroid(ring)-C20.0
C3-C40.002Mn-centroid(ring)-C30.1
Mn-C10.019Mn-C1-C60.1
Mn-C20.022Mn-C2-C7−0.3
Mn-C30.024Mn-C3-C8−0.7
C6-plane(ring)0.009centroid(ring)-C1-C6−0.3
C7-plane(ring)0.001centroid(ring)-C2-C7−0.1
C8-plane(ring)−0.012centroid(ring)-C3-C80.6
Mn-plane(ring)0.024
Table 3. Charges q (a.u.) on selected atoms a and fragments in the optimized Cp*2Mn0 neutral and Cp*2Mn+ cation and their change upon ionization ∆q.
Table 3. Charges q (a.u.) on selected atoms a and fragments in the optimized Cp*2Mn0 neutral and Cp*2Mn+ cation and their change upon ionization ∆q.
Atom or Fragmentq (Neutral)q (Cation)q
Mn0.8741.0720.198
C1−0.103−0.105−0.002
C2−0.114−0.118−0.004
C3−0.123−0.129−0.006
C60.000−0.003−0.003
C70.003−0.001−0.004
C80.0050.001−0.004
H(C6) b0.0060.0500.044
H(C6) c0.0050.0280.023
H(C6) c0.0060.0290.023
H(C7) b0.0090.0530.044
H(C7) c0.0070.0290.022
H(C7) c0.0080.0310.023
H(C8) b0.0120.0560.044
H(C8) c0.0090.0310.022
H(C8) c0.0100.0320.022
Me(C6)0.0230.1080.085
Me(C7)0.0270.1120.085
Me(C8)0.0320.1160.084
Cp*−0.437−0.0360.401
a See the text and Figure 5 and Figure 6 for the atom numbering; the carbon atoms of the methyl groups are indicated in parentheses. b The H atoms oriented away from the metal. c The H atoms oriented towards the metal.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ketkov, S.; Tzeng, S.-Y.; Rychagova, E.; Tzeng, W.-B. Ionization of Decamethylmanganocene: Insights from the DFT-Assisted Laser Spectroscopy. Molecules 2022, 27, 6226. https://doi.org/10.3390/molecules27196226

AMA Style

Ketkov S, Tzeng S-Y, Rychagova E, Tzeng W-B. Ionization of Decamethylmanganocene: Insights from the DFT-Assisted Laser Spectroscopy. Molecules. 2022; 27(19):6226. https://doi.org/10.3390/molecules27196226

Chicago/Turabian Style

Ketkov, Sergey, Sheng-Yuan Tzeng, Elena Rychagova, and Wen-Bih Tzeng. 2022. "Ionization of Decamethylmanganocene: Insights from the DFT-Assisted Laser Spectroscopy" Molecules 27, no. 19: 6226. https://doi.org/10.3390/molecules27196226

Article Metrics

Back to TopTop