Next Article in Journal
L-Arginine Enhances Oral Keratinocyte Proliferation under High-Glucose Conditions via Upregulation of CYP1A1, SKP2, and SRSF5
Next Article in Special Issue
Real-Time Extension of TAO-DFT
Previous Article in Journal
The Content of Bioactive Compounds and Technological Properties of Matcha Green Tea and Its Application in the Design of Functional Beverages
Previous Article in Special Issue
On the Potential Role of the (Pseudo-) Jahn–Teller Effect in the Membrane Transport Processes: Enniatin B and Beauvericin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Jahn–Teller Effect in Silver Complexes of Dimethyl Amino Phenyl Substituted Phthalocyanine †

Department of Physical Chemistry, Slovak Technical University, Radlinskeho 9, SK-81237 Bratislava, Slovakia
Dedicated to Professor Andrej Staško in honor of his 85th birthday.
Molecules 2023, 28(20), 7019; https://doi.org/10.3390/molecules28207019
Submission received: 9 September 2023 / Revised: 3 October 2023 / Accepted: 9 October 2023 / Published: 10 October 2023
(This article belongs to the Special Issue Multiconfigurational and DFT Methods Applied to Chemical Systems)

Abstract

:
The structures of Ag complexes with dimethyl amino phenyl substituted phthalocyanine m[dmaphPcAg]q of various charges q and in the two lowest spin states m were optimized using the B3LYP method within the D4h symmetry group and its subgroups. The most stable reaction intermediate in the supposed photoinitiation reaction is 3[dmaphPcAg]. Group-theoretical analysis of the optimized structures and of their electron states reveals two symmetry-descent mechanisms. The stable structures of maximal symmetry of complexes 1[dmaphPcAg]+, 3[dmaphPcAg]+, 2[dmaphPcAg]0, and 4[dmaphPcAg]2− correspond to the D4 group as a consequence of the pseudo-Jahn–Teller effect within unstable D4h structure. Complexes 4[dmaphPcAg]0, 1[dmaphPcAg], 3[dmaphPcAg], and 2[dmaphPcAg]2− with double degenerate electron ground states in D4h symmetry structures undergo a symmetry descent to stable structures corresponding to maximal D2 symmetry, not because of a simple Jahn–Teller effect but due to a hidden pseudo-Jahn–Teller effect (strong vibronic interaction between excited electron states). The reduction of the neutral photoinitiator causes symmetry descent to its anionic intermediate because of vibronic interactions that must significantly affect the polymerization reactions.

1. Introduction

New photosensitive systems with improved efficiency for initiating free-radical and/or cationic polymerizations under light activation are developed using sophisticated chemical treatments. Very recently, Breloy et al. [1] synthesized dimethyl amino substituted phthalocyanine dmaphPcH2 and its Ag(II) complex 2[dmaphPcAg]0 (Figure 1). Photoexcitation of [dmaphPcAg]0 in the presence and absence of an iodonium salt produced acidic and radical species that initiated cationic and free-radical polymerizations, respectively. The photoinitiator properties were investigated via spectroscopic and quantum-chemical methods. The polymerization kinetics in laminate and under air was studied. In the first step, irradiation (385 and 405 nm) caused a reduction of Ag(II) to Ag(I), and nitrogen-centered radicals were formed. Subsequently, Ag nanoparticles and carbon-centered radicals developed. Intermediates [dmaphPc], Ag0 nanoparticles, and [dmaphPcAg]q complexes, q = −1 → +1, are supposed within the proposed reaction mechanism. DFT calculations indicate a D4 → D2 symmetry descent in [dmaphPcAg]q complexes which could be ascribed to the Jahn–Teller (JT) effect. The aim of our recent study is to shed more light on this problem and to verify the above explanation using a group-theoretical analysis of the DFT-optimized structures of [dmaphPcAg]q complexes within the highest possible D4h symmetry group and its subgroups.

2. Theoretical Background

According to the Jahn–Teller theorem [2], any nonlinear configuration of atomic nuclei in a degenerate electron state is unstable. Hence, there is at least one such configuration of lower symmetry where the above degeneracy is removed. In other words, the multidimensional representation in the high-symmetric (HS) structure is split into nondegenerate representations in the low-symmetric (LS) structure.
The JT active coordinate QJT describes the abovementioned symmetry descent. If the potential energy surface E = f(Qi) of N atoms is a function of 3N − 6 independent nuclear coordinates Qi, for the JT ‘unstable’ HS structure, the following relation for any QJT is valid:
E Q JT HS 0
The HS → LS geometry change described by QJT is connected with an energy decrease which is denoted as the JT stabilization energy EJT:
EJT = EHSELS
An analogous symmetry descent and energy decrease for pseudodegenerate electron states is known as the pseudo-Jahn–Teller (PJT) effect. It can be observed for a sufficiently strong vibronic interaction and relatively small energy difference Δij between the interacting electron states Ψi and Ψj [3].
In some cases, the existence of LS structures cannot be explained by the JT effect of the degenerate ground electron state nor by the PJT interactions of the non-degenerate ground electron state with low-lying excited states in the HS structure (e.g., because of their different spin multiplicity). (P)JT effects in these cases are “hidden” in the excited states, which can undergo vibronic interactions as well. The vibronic interaction within the degenerate excited state or between two (or more) excited states of suitable symmetries can be so strong that the lower energy surface penetrates the potential surface of the ground electron state corresponding the HS structure and becomes the lowest state in the LS structure (i.e., its ground electron state). These consequences of the strong JT and PJT vibronic interactions within excited electron states are known as the hidden JT (HJT) and hidden PJT (HPJT) effects, respectively [4].
(P)JT potential surfaces can be described using an analytical function based on perturbation theory. In the simplest case of single distortion coordinate Q and double electron degeneracy or two interacting electron states of different symmetries, we obtain their potential energy surface E(Q) in the form
E Q = 1 2 K Q 2 ± [ ij 2 4 + F 2 Q 2 ] 1 / 2
where Δij = 0 for double electron degeneracy (JT effect) or Δij > 0 is the energy difference between both interacting electron states (PJT effect) in the undistorted geometry, K is the primary force constant (without vibronic coupling), and F is the vibronic coupling constant [3]. However, the search for their extremal points corresponding to the ‘stable’ or ‘unstable’ (P)JT structures is too complicated for large molecular systems. In such cases, a group-theoretical treatment must be used.
The epikernel principle method [5,6] is based on the QJT symmetry in the HS structure. A nonzero value of the integral
< Ψ i H ^ Q JT Ψ j > 0
for the interacting electron states Ψi and Ψj and the full-symmetric Hamilton operator H ^ demands that the direct product
Γi* ⊗ ΛJT ⊗ Γj
where Γi, ΛJT, and Γj are representations of Ψi, QJT, and Ψj, respectively (the asterisk denotes a complex conjugate value), must contain a full-symmetric representation. Alternatively:
ΛJT ∈ Γi* ⊗ Γj
In the case of the JT effect and the degenerate states Ψi = Ψj, i.e., Γi = Γj, we obtain the even stronger condition
ΛJT ∈ [Γi ⊗ Γi]+
where […]+ denotes the symmetric direct product.
The epikernel principle states that the extrema of a JT energy surface correspond to the kernel K(G, ΛJT) or epikernel E(G, ΛJT) subgroups of the HS parent group G. Kernels contain the symmetry operations of G that leave the ΛJT representation invariant. The symmetry operations of epikernels leave invariant only some components of the degenerate ΛJT representation.
The epikernel principle was originally formulated for the systems in degenerate electron states only [5,6], but it has been successfully extended (except Equation (7)) to pseudodegenerate electron states as well [7]. Its drawback is in the restriction to JT active coordinates that have been derived within perturbation theory for the linear Taylor expansion of the perturbation only. This method might offer incomplete results in some cases (e.g., in systems with C5 rotations) [7].
The method of step-by-step symmetry descent [8,9] is based on consecutive splitting of a degenerate electron state within a JT symmetry descent. The vibronic interaction causes instability of the HS structure, and thus, in the sense of the JT theorem [2], some symmetry elements are removed. The probability of the symmetry elements removal decreases with the number of these elements. It implies that the symmetry decrease to the immediate subgroups with the lowest number of the removed symmetry elements are more probable than to other subgroups. The multidimensional representation corresponding to the degenerate electron state can be split within symmetry descent to an immediate subgroup of the parent HS group G (see Table S1 in Supplementary Materials). If the structure corresponding to this subgroup is in a nondegenerate electron state, it is JT stable and the symmetry descent stops. If the structure corresponding to an immediate subgroup of G is in a degenerate electron state described by a multidimensional representation (a JT unstable structure), the symmetry descent continues to its immediate subgroups and the whole procedure is repeated. As every group can have several immediate subgroups, various symmetry descent paths are possible and the JT stable structures can correspond to various LS symmetry groups. The only condition is its nondegenerate electron state (i.e., one-dimensional representation), obtained through splitting the degenerate electron state (i.e., multidimensional representation) of the HS structure of the parent group G.

3. Results

The highest possible structures of [dmaphPcAg]q complexes are of the D4h symmetry group with side phenyl groups perpendicular to the central phthalocyanine plane (Figure 2). Because of the great number of possible mutual orientations of dimethyl amino phenyl groups in less symmetric structures, we restricted our study to the stable structures of the maximal symmetry group only. Another restriction is implied by the inability of standard DFT methods to optimize the atomic configurations in degenerate electron states [10]. The results of the geometry optimizations of the m[dmaphPcAg]q complexes with total charges q = +1 → −2 in the two lowest spin states m are presented in Table 1. It is interesting that the [dmaphPcAg]q complexes in the triplet spin state are more stable than these ones with the same charge in the singlet spin state. 3[dmaphPcAg] of D2 symmetry is the most stable complex under study. It must be mentioned that the use of an unrestricted ‘broken symmetry’ treatment [11] results in zero spin populations and does not reduce the energy of the systems studied. Therefore, only the restricted Kohn–Sham formalism has been used in subsequent TD-DFT calculations of our complexes in singlet ground states.
Further inspection of Table 1 shows that the studied complexes can be divided into two categories:
(i)
Category I contains complexes 1[dmaphPcAg]+, 3[dmaphPcAg]+, 2[dmaphPcAg]0, and 4[dmaphPcAg]2− with optimized structures of D4h (unstable) and D4 (stable) symmetry groups.
(ii)
Category II contains complexes 4[dmaphPcAg]0, 1[dmaphPcAg], 3[dmaphPcAg], and 2[dmaphPcAg]2−, where only their stable optimized structures of the D2 symmetry group were found, while the D4h optimized structures were absent. This can be explained by the electron configurations of the D4h complexes in Table 2. We may conclude that the category II complexes should contain partially occupied eg molecular orbitals, which implies Eg or Eu ground electron states. Hence, they are not accessible through standard DFT methods.

3.1. Category I Complexes

The optimized D4h structures of these complexes are unstable due to several imaginary vibrations of Λim representations (Table 1) that coincide with the JT active coordinates. The JT stabilization energies EJT correspond to D4h → D4 symmetry decrease. Stable D4 structures with equally rotated phenyl rings (Figure 3) and without any imaginary vibration correspond to the K(D4h, Λim) kernel subgroups for the coordinate of the a1u representation. As indicated by its wavenumber νim, the energies of the corresponding vibrations are comparable with the highest energy ones in all category I complexes. The optimized structures of D2d and C4v symmetries are not stable (not presented), and hence only the D4 ones fulfill the condition of the stable structure of the highest symmetry group.
For known ground state Γ0 and JT coordinate Λim representations, the Equations (4) and (5) can be used to determine the excited-state representations Γexc (see Table 1) interacting with the ground electron state in D4h structures of our complexes. The comparison of the TD-DFT calculated electron states in the corresponding D4h and D4 structures (Table 3) shows that the energy difference Eexc between the PJT interacting states increases after symmetry descent as a consequence of their vibronic interaction. This confirms the correct assignment of the corresponding states in both groups because the standard treatment based on similarity of their oscillator strengths f is hardly usable in our cases. The ground states in the D4h structures also correspond to those in their D4 subgroups.
We can see (Table 3) that the PJT active excited states in the D4h structures are relatively high. This indicates that their excitation energies are less important for possible vibronic interactions than the energies of JT active coordinates.

3.2. Category II Complexes

As mentioned above, the ground state of the D4h structures of these complexes is double degenerate (Eg or Eu representations), and only the stable D2 structures (Figure 4) were obtained through geometry optimizations. Possible representations of the corresponding JT active coordinates can be obtained via the symmetric direct products for the HS group:
[Eg ⊗ Eg]+ = [Eu ⊗ Eu]+ = A1g ⊕ B1g ⊕ B2g
The full-symmetric a1g coordinate does not change the symmetry of the structure (i.e., cannot be JT active) and the kernels
K(D4h, b1g) = D2h(C2′)
K(D4h, b2g) = D2h(C2″)
do not explain the existence of the optimized D2 structure. Therefore, the method of the epikernel principle [5,6] cannot explain the D4h → D2 symmetry descent.
The method of step-by-step symmetry descent [8,9] for the structures of the D4h symmetry group in double degenerate electron states is based on the scheme in Figure 5. Except D2h, all immediate subgroups of the D4h group (i.e., D4, D2d, C4v, and C4h) are JT unstable because of preserved electron degeneracy. After the removal of the C4 axis from the D4 group, we obtain its immediate subgroup D2, which is JT stable because the degeneracy is removed here. Alternatively, the D2 group can be obtained via symmetry descent through the JT unstable D2d group with preserved electron degeneracy. The structures of the C4 and S4 symmetry groups preserve electron degeneracy and therefore cannot be stable. We have not found any stable structure of D2h, C2v, or C2h symmetry groups through geometry optimization. Therefore, the stable D2 structures meet the condition of maximal groups for category II complexes.
During the symmetry descent from D4h to D2, the double degenerate representation is split into the nondegenerate B2 and B3 ones:
Eg or Eu (D4h) → E (D4 or D2d) → B2 ⊕ B3 (D2)
None of these representations corresponds to the calculated ground electron state of the stable D2 structures (Table 4). This discrepancy can be explained using the above-mentioned HPJT effect, and the ground state of the D2 structure corresponds to the lower PJT interacting excited state in its supergroup. In our case, the situation is complicated by the fact that D2 is not an immediate subgroup of the D4h group. Therefore, there are three possible two-step symmetry descent paths:
D 4 h   a 1 u   D 4   b 1   o r   b 2   D 2
D 4 h   b 1 g   o r   b 2 g   D 2 h   a u   D 2
D 4 h   b 1 u   o r   b 2 u   D 2 d   b 1   D 2
Based on group–subgroup relations, we can assign the electron state representations of the D2 symmetry group to the corresponding D4, D2h, or D2d ones despite the several alternatives. The same holds for the D4h to D4, D2h, or D2d symmetry descent. For known representations of ground states and JT active coordinates, it might be possible to determine the HPJT excited-state representations according to Equation (4). For the two-step symmetry descent, there are too many possibilities where almost all excited-state representations (except two-dimensional) might be involved. Therefore, we shall not deal with this problem.

4. Methods

We have performed geometry optimization of the complexes m[dmaphPcAg]q with charges q = +1 → −2 in the two lowest spin states (singlet to quartet) with spin multiplicities m within the D4h symmetry and its subgroups. In agreement with our previous study [1], B3LYP hybrid functional [12], GD3 dispersion correction [13], cc-pVDZ-PP pseudopotential and basis set for Ag [14], and cc-pVDZ basis sets for remaining atoms [15] were used. The optimized structures were tested on the number of imaginary vibrations. Excited states (up to 50) were calculated for every optimized structure using time-dependent DFT (TD-DFT) treatment [16,17], analogously to our previous studies [1,18,19]. All calculations were performed with Gaussian16 (Revision B.01) software [20]. MOLDRAW (Release 2.0, https://www.moldraw.software.informer.com, accessed on 9 September 2019) software [21] was used for visualization and geometry modification purposes. Finally, a group-theoretical analysis of the obtained results was carried out using the methods of the epikernel principle [5,6,7] and of step-by-step symmetry descent [8,9].
The B3LYP functional is most frequently used in quantum-chemical studies and produces relatively reliable excitation energies. The basis sets used are restricted by our technical capabilities.

5. Conclusions

To shed more light on the photopolymerization action of the Ag(II) complex with dimethylamino phenyl-substituted phthalocyanine 2[dmaphPcAg]0, we have performed a quantum-chemical model study of possible reaction intermediates [dmaphPcAg]q with charges q = +1 → −2, with the aim to explain the possible role of the JT effect. Group-theoretical analysis of the obtained results shows that the complexes under study are of two categories.
The stable structures of maximal symmetry of 1[dmaphPcAg]+, 3[dmaphPcAg]+, 2[dmaphPcAg]0, and 4[dmaphPcAg]2− complexes (category I) correspond to the D4 group as a consequence of the PJT effect within the unstable D4h structure. On the other hand, complexes 4[dmaphPcAg]0, 1[dmaphPcAg], 3[dmaphPcAg], and 2[dmaphPcAg]2− (category II) with double degenerate electron ground states in (JT unstable) D4h symmetry structures undergo a symmetry descent to stable structures corresponding to maximal D2 symmetry, not because of a simple JT effect but due to HPJT effect. The most stable reaction intermediate in the supposed photoinitiation reaction [1] is surprisingly 3[dmaphPcAg] (a singlet electron state was expected). Therefore, the reduction of the 2[dmaphPcAg]0 photoinitiator (D4 symmetry) to the 3[dmaphPcAg] intermediate (D2 symmetry) must be significantly affected by vibronic interactions (PJT and HPJT effects), primarily the reaction barrier height and reaction equilibrium. Moreover, the reactivity of 3[dmaphPcAg] is supported by the non-equal spin density at nitrogen atoms of the D2 structure (doubled at one pair of these atoms; see Table S2 in Supplementary Materials). Nevertheless, their exact influence on reaction rates and thermodynamics must be investigated in solutions.
The presented study shows how group-theoretical treatment can be profitable through solving chemical problems for large molecules. The method of epikernel principle [5,6,7] is restricted to JT active coordinates based on perturbation theory and thus grants incomplete results, but it can also be used for the PJT effect. The method of step-by-step symmetry descent [8,9] based on splitting degenerate electron states during symmetry descent is more universal but not suitable for the PJT effect. We have shown the usefulness of the combination of both methods. Further theoretical studies in this field are welcome.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28207019/s1, Table S1. Relations between the irreducible representations of the parent group D4h and of its D4 and D2 subgroups. Table S2. Charge q, spin multiplicity m, symmetry group G, Mulliken atomic charges and spin populations of Ag and pyrrole N atoms in m[dmaphPc]q complexes under study. Tables S3–S14. Atom coordinates of m[dmaphPc]q in D4h, D4 or D2 symmetry (in Å).

Funding

The work has been supported by the Slovak Research and Development Agency (no. APVV-19-0087), by the Slovak Scientific Grant Agency VEGA (no. 1/0139/20), and by Ministry of Education, Science, Research and Sport of the Slovak Republic within the scheme “Excellent research teams”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors thank the HPC center at the Slovak University of Technology in Bratislava, which is a part of the Slovak Infrastructure of High Performance Computing (SIVVP project ITMS 26230120002, funded by European Region Development Funds) for the computational time and resources made available.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Sample Availability

Not applicable.

References

  1. Breloy, L.; Alcay, Y.; Yilmaz, I.; Breza, M.; Bourgon, J.; Brezová, V.; Yagci, Y.; Versace, D.-L. Dimethyl amino phenyl substituted silver phthalocyanine as a UV- and visible-light absorbing photoinitiator: In situ preparation of silver/polymer nanocomposites. Polym. Chem. 2021, 12, 1273–1285. [Google Scholar] [CrossRef]
  2. Jahn, H.A.; Teller, E. Stability of polyatomic molecules in degenerate electronic states. I. Orbital degeneracy. Proc. R. Soc. London A 1937, 161, 220–235. [Google Scholar]
  3. Bersuker, I.B. Pseudo-Jahn–Teller Effect—A Two-State Paradigm in Formation, Deformation, and Transformation of Molecular Systems and Solids. Chem. Rev. 2013, 113, 1351–1390. [Google Scholar] [CrossRef] [PubMed]
  4. Bersuker, I.B. Recent Developments in the Jahn–Teller Effect Theory. In The Jahn-Teller Effect. Fundamentals and Implications for Physics and Chemistry; Köppel, H., Yarkony, D.R., Barentzen, H., Eds.; Springer: Berlin/Heidelberg, Germany, 2009; pp. 3–24. ISBN 978-3642034312. [Google Scholar]
  5. Ceulemans, A.; Beyens, D.; Vanquickenborne, L.G. Symmetry aspects of Jahn-Teller activity: Structure and reactivity. J. Am. Chem. Soc. 1984, 106, 5824–5837. [Google Scholar] [CrossRef]
  6. Ceulemans, A.; Vanquickenborne, L.G. The Epikernel Principle. Struct. Bonding 1989, 71, 125–159. [Google Scholar]
  7. Breza, M. Group-Theoretical Treatment of Pseudo-Jahn-Teller Systems. In Vibronic Interactions and the Jahn-Teller Effect: Theory and Application; Atanasov, M., Daul, C., Tregenna-Piggott, P.L.W., Eds.; Springer: Dordrecht, The Netherlands; Berlin/Heidelberg, Germany; London, UK; New York, NY, USA, 2012; pp. 59–82. ISBN 1567-7354. [Google Scholar]
  8. Pelikán, P.; Breza, M. Classification of the possible symmetries of the Jahn—Teller systems. Chem. Papers 1984, 39, 255–270. [Google Scholar]
  9. Breza, M. Group-Theoretical Analysis of Jahn-Teller Systems. In The Jahn-Teller Effect. Fundamentals and Implications for Physics and Chemistry; Köppel, H., Yarkony, D.R., Barentzen, H., Eds.; Springer: Berlin/Heidelberg, Germany, 2009; pp. 51–76. ISBN 978-3642034312. [Google Scholar]
  10. Kaplan, I.G. Problems in DFT with the total spin and degenerate states. Int. J. Quantum Chem. 2007, 107, 2595–2603. [Google Scholar] [CrossRef]
  11. Schurkus, H.F.; Chan, G.K.-L.; Chen, G.-T.; Cheng, H.-P.; Stanton, J.F. Theoretical prediction of magnetic exchange coupling constants from broken-symmetry coupled cluster calculations. J. Chem. Phys. 2020, 152, 234115. [Google Scholar] [CrossRef] [PubMed]
  12. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  13. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parameterization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  14. Pritchard, B.P.; Altarawy, D.; Didier, B.; Gibson, T.D.; Windus, T.L. New Basis Set Exchange: An Open, Up-to-Date Resource for the Molecular Sciences Community. J. Chem. Inf. Model. 2019, 59, 4814–4820. [Google Scholar] [CrossRef] [PubMed]
  15. Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  16. Bauernschmitt, R.; Ahlrichs, R. Treatment of electronic excitations within the adiabatic approximation of time dependent density functional theory. Chem. Phys. Lett. 1996, 256, 454–464. [Google Scholar] [CrossRef]
  17. Scalmani, G.; Frisch, M.J.; Mennucci, B.; Tomasi, J.; Cammi, R.; Barone, V. Geometries and properties of excited states in the gas phase and in solution: Theory and application of a time-dependent density functional theory polarizable continuum model. J. Chem. Phys. 2006, 124, 94107. [Google Scholar] [CrossRef] [PubMed]
  18. Štellerová, D.; Lukeš, V.; Breza, M. How does pseudo-Jahn-Teller effect induce the photoprotective potential of curcumin? Molecules 2023, 28, 2946. [Google Scholar] [CrossRef] [PubMed]
  19. Štellerová, D.; Lukeš, V.; Breza, M. On the Potential Role of (Pseudo-) Jahn-Teller Effect in Membrane Transport Processes: Enniatin B and Beauvericin. Molecules 2023, 28, 6264. [Google Scholar] [CrossRef] [PubMed]
  20. Frisch, G.W.; Trucks, M.J.; 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, Revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  21. Ugliengo, P. MOLDRAW: A Program to Display and Manipulate Molecular and Crystal Structures (Release 2.0), University Torino, Torino. 2012. Available online: https://www.moldraw.software.informer.com (accessed on 9 September 2019).
Figure 1. Structure of 2[dmaphPcAg]0 [1].
Figure 1. Structure of 2[dmaphPcAg]0 [1].
Molecules 28 07019 g001
Figure 2. Optimized D4h structure of 2[dmaphPcAg]0 (C—black, N—blue, H—white, Ag—grey).
Figure 2. Optimized D4h structure of 2[dmaphPcAg]0 (C—black, N—blue, H—white, Ag—grey).
Molecules 28 07019 g002
Figure 3. Optimized D4 structure of 2[dmaphPcAg]0 (see Figure 2 for atom notation).
Figure 3. Optimized D4 structure of 2[dmaphPcAg]0 (see Figure 2 for atom notation).
Molecules 28 07019 g003
Figure 4. Optimized D2 structure of 1[dmaphPcAg] (see Figure 2 for atom notation).
Figure 4. Optimized D2 structure of 1[dmaphPcAg] (see Figure 2 for atom notation).
Molecules 28 07019 g004
Figure 5. Possible symmetry descent paths of D4h structures in double degenerate electron states [9]. The top and bottom lines of individual rectangles denote symmetry groups and the corresponding irreducible representations, respectively.
Figure 5. Possible symmetry descent paths of D4h structures in double degenerate electron states [9]. The top and bottom lines of individual rectangles denote symmetry groups and the corresponding irreducible representations, respectively.
Molecules 28 07019 g005
Table 1. Charge q, spin multiplicity m, symmetry group G, representation of the ground electron state Γ0, DFT energy EDFT, JT stabilization energy EJT, representations Λim and wavenumbers νim of imaginary vibrations, kernel K(D4h, Λim) and epikernel K(D4h, Λim) subgroups of D4h, and representations of relevant PJT excited states Γexc of m[dmaphPc]q complexes under study (the preserved symmetry elements in the kernel and epikernel subgroups are in parentheses). The most stable structure is shown in bold.
Table 1. Charge q, spin multiplicity m, symmetry group G, representation of the ground electron state Γ0, DFT energy EDFT, JT stabilization energy EJT, representations Λim and wavenumbers νim of imaginary vibrations, kernel K(D4h, Λim) and epikernel K(D4h, Λim) subgroups of D4h, and representations of relevant PJT excited states Γexc of m[dmaphPc]q complexes under study (the preserved symmetry elements in the kernel and epikernel subgroups are in parentheses). The most stable structure is shown in bold.
qmGΓ0EDFT [Hartree]EJT [eV]Λimνim [cm−1]K(D4h, Λim)E(D4h, Λim)Γexc
+11D4h1A1g−4734.55067-b1u−48D2d(C2′) B1u
2eg−47, −31C1C2h(C2′), C2h(C2″)Eg
a1u−46D4 A1u
a2u−31C4v A2u
b2u−31D2d(C2″) B2u
+11D41A1−4734.583520.894--
+13D4h3B1u−4734.54698-b1u−18D2d(C2′) A1g
eg−18C1C2h(C2′), C2h(C2″)Eu
a1u−18D4 B1g
+13D43B1−4734.585501.048--
02D4h2B1g−4734.74708-b1u−42D2d(C2′) A1u
eg−42C1C2h(C2′), C2h(C2″)Eg
a1u−41D4 B1u
a2u−23C4v B2u
3b2u−23(3×)D2d(C2″) A2u
02D42B1−4734.772830.701--
04D24B2−4734.73113unknown--
−11D21A−4734.79846unknown--
−13D23B2−4734.83355unknown--
−22D22B1−4734.80520unknown--
−24D4h4B1u−4734.77355-b1u−47D2d(C2′) A1g
2eg−46, −27C1C2h(C2′), C2h(C2″)Eu
a1u−45D4 B1g
a2u−27C4v B2g
b2u−27D2d(C2″) A2g
−24D44B1−4734.805810.878--
Table 2. Charge q, spin multiplicity m, electron configuration, and ground electron state representation Γ0 of studied m[dmaphPc]q complexes of D4h symmetry group.
Table 2. Charge q, spin multiplicity m, electron configuration, and ground electron state representation Γ0 of studied m[dmaphPc]q complexes of D4h symmetry group.
qmElectron ConfigurationΓ0
+11…(b2g)2(eu)4(a2g)2(b1g)0(eg)0(b1u)01A1g
+13α: …(b1g)1(a2g)1(eu)2(b2g)1(eg)0(b1u)0
β: …(eu)2(b2g)1(a1u)0(eg)0(b1g)0(b1u)0
3B1u
02α:…(eu)2(b2g)1(b1g)1(a1u)1(eg)0(b1u)0
β:…(eu)2(b2g)1(a1u)1(eg)0(b1g)0(b1u)0
2B1g
04unknown4Eg or 4Eu
−11unknown1Eg or 1Eu
−13unknown3Eg or 3Eu
−22unknown2Eg or 2Eu
−24α:…(eg)2(b2g)1(a1u)1(b1g)1(eg)2(b2u)0
β:…(eg)2(a2u)1(a1u)1(eg)0(a2u)0(b2u)0
4B1u
Table 3. Charge q, spin multiplicity m, symmetry group G, ground electron state representation Γ0, representations Γexc, excitation energies Eexc, and oscillator strengths f of the low excited electron states of the studied m[dmaphPc]q complexes in D4h and D4 symmetry groups. The excited states that interact with the ground states are in bold.
Table 3. Charge q, spin multiplicity m, symmetry group G, ground electron state representation Γ0, representations Γexc, excitation energies Eexc, and oscillator strengths f of the low excited electron states of the studied m[dmaphPc]q complexes in D4h and D4 symmetry groups. The excited states that interact with the ground states are in bold.
qmGΓ0ΓexcEexc [eV]fGΓ0ΓexcEexc [eV]f
+11D4h1A1g11A2g0.1560.000D41A111B10.1670.000
11Eu0.1580.002 11B20.2550.000
11B2g0.1590.000 11E0.2590.000
11B1g0.3490.000 11A20.2670.000
21Eu0.3530.081 21E0.6310.014
11B1u0.3770.000 11A10.6420.000
11A1g0.4870.000 21B10.8580.000
11Eg1.0100.000 31E1.1870.050
11A1u1.0100.000 21A11.1920.001
11B1u1.0100.000 21A21.2000.000
+13D4h3B1u13B2g0.0110.000D43B113E0.1880.007
13A2g0.0150.001 13B20.1900.000
13Eu0.0160.001 13A20.1950.000
13A1g0.2350.000 13A10.6000.000
23Eu0.2350.014 23E0.6170.298
13B2u0.2390.000 13B10.7790.000
23Eu1.2550.000 33E1.2440.001
33Eu1.3270.003 43E1.3910.033
13Eg1.4930.008 23B11.4000.000
02D4h2B1g12Eu1.1950.000D42B112E1.1570.000
12Eg1.3060.000 22E1.3050.000
12B1u2.0140.000 32E1.9010.624
22Eu2.0500.710 12B21.9420.000
22Eg2.140 0.000 12A21.9440.000
12B2u2.1400.000 42E1.9560.004
12A2u2.1400.000 22A21.9690.000
32Eg2.1630.000 12A12.0550.000
22B2u2.1640.000 32A22.0560.000
22A2u2.1640.000 12B12.0590.001
−24D4h4B1u14Eg0.9570.000D44B114E0.8880.191
14A2u0.9760.000 14B11.3560.000
14B2u0.9780.000 24E1.3600.010
24Eg1.0250.000 14A11.3600.000
14A1u1.0420.000 34E1.3960.010
14B1u1.0480.000 14A21.4150.000
34Eg1.1180.110 14B21.4170.000
14A1g1.4950.000 44E1.4360.012
14B1g1.4950.000 24B21.4720.001
14Eu1.4960.000 24A21.4810.000
24A1u1.5010.000 34B11.5230.000
Table 4. Charge q, spin multiplicity m, representations of the ground electron state Γ0 and of the excited states Γexc, excitation energies Eexc, and oscillator strengths f of low excited electron states of stable m[dmaphPc]q complexes under study in the D2 symmetry group. The irreducible representations obtained through splitting the two-dimensional representations of the degenerate ground state in D4h structures are in bold.
Table 4. Charge q, spin multiplicity m, representations of the ground electron state Γ0 and of the excited states Γexc, excitation energies Eexc, and oscillator strengths f of low excited electron states of stable m[dmaphPc]q complexes under study in the D2 symmetry group. The irreducible representations obtained through splitting the two-dimensional representations of the degenerate ground state in D4h structures are in bold.
qmΓ0ΓexcEexc [eV]fqmΓ0ΓexcEexc [eV]f
044B214B10.3490.000−111A11B20.5230.000
14B30.8140.012 11B10.5390.000
24B10.8290.000 11B30.8150.000
14B20.9270.006 21B21.3370.023
34B10.9480.000 21B31.4780.028
24B31.2530.000 31B21.5510.005
24B21.2540.213 41B21.7930.809
14A1.2910.000 31B31.9650.018
34B31.3560.164 11A2.0560.000
24A1.4230.000 41B32.0970.011
−133B213B10.2720.000−222B112B10.0480.000
13B31.2560.000 22B10.3770.000
13B21.3500.021 12B20.9000.007
23B31.3560.014 22B21.0950.283
23B21.7960.792 12B31.1120.306
33B31.8350.007 32B21.1680.000
43B31.9870.012 12A1.3430.000
13A1.9930.000 42B21.3480.002
23B12.0000.000 22A1.3620.000
23A2.0920.000 22B31.3670.002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Breza, M. On the Jahn–Teller Effect in Silver Complexes of Dimethyl Amino Phenyl Substituted Phthalocyanine. Molecules 2023, 28, 7019. https://doi.org/10.3390/molecules28207019

AMA Style

Breza M. On the Jahn–Teller Effect in Silver Complexes of Dimethyl Amino Phenyl Substituted Phthalocyanine. Molecules. 2023; 28(20):7019. https://doi.org/10.3390/molecules28207019

Chicago/Turabian Style

Breza, Martin. 2023. "On the Jahn–Teller Effect in Silver Complexes of Dimethyl Amino Phenyl Substituted Phthalocyanine" Molecules 28, no. 20: 7019. https://doi.org/10.3390/molecules28207019

Article Metrics

Back to TopTop