ACS Publications. Most Trusted. Most Cited. Most Read
Topological Effects in Vibronically Coupled Degenerate Electronic States: A Case Study on Nitrate and Benzene Radical Cation
My Activity
  • Open Access
Article

Topological Effects in Vibronically Coupled Degenerate Electronic States: A Case Study on Nitrate and Benzene Radical Cation
Click to copy article linkArticle link copied!

  • Soumya Mukherjee
    Soumya Mukherjee
    Department of Physical Chemistry, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700032, India
  • Bijit Mukherjee
    Bijit Mukherjee
    Department of Physical Chemistry, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700032, India
  • Joy Dutta
    Joy Dutta
    Department of Physical Chemistry, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700032, India
    More by Joy Dutta
  • Subhankar Sardar
    Subhankar Sardar
    Department of Chemistry, Bhatter College, Dantan, Paschim Medinipur, 721426, India
  • Satrajit Adhikari*
    Satrajit Adhikari
    Department of Physical Chemistry, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700032, India
    *E-mail: [email protected] (S.A.).
Open PDF

ACS Omega

Cite this: ACS Omega 2018, 3, 10, 12465–12475
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.8b01648
Published October 3, 2018

Copyright © 2018 American Chemical Society. This publication is licensed under these Terms of Use.

Abstract

Click to copy section linkSection link copied!

We carry out detailed investigation for topological effects of two molecular systems, NO3 radical and C6H6+ (Bz+) radical cation, where the dressed adiabatic, dressed diabatic, and adiabatic-via-dressed diabatic potential energy curves (PECs) are generated employing ab initio calculated adiabatic and diabatic potential energy surfaces (PESs). We have implemented beyond Born–Oppenheimer (BBO) theory for constructing smooth, single-valued, and continuous diabatic PESs for five coupled electronic states [J. Phys. Chem. A2017,121, 6314–6326]. In the case of NO3 radical, the nonadiabatic coupling terms (NACTs) among the low-lying five electronic states, namely, 2A2 (12B2), Ã2E″ (12A2 and 12B1), and B̃2E′ (12A1 and 22B2), bear the signature of Jahn–Teller (JT) interactions, pseudo JT (PJT) interactions, and accidental conical intersections (CIs). Similarly, Bz+ radical cation also exhibits JT, PJT, and accidental CIs in the interested domain of nuclear configuration space. In order to generate dressed PECs, two components of degenerate in-plane asymmetric stretching modes are selectively chosen for both the molecular species (Q3xQ3y pair for NO3 radical and Q16xQ16y pair for Bz+ radical cation). The JT coupling between the electronic states is essentially originated through the asymmetric stretching normal mode pair, where the coupling elements exhibit symmetric and nonlinear functional behavior along Q3x and Q16x normal modes.

Copyright © 2018 American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

The Born–Oppenheimer (BO) approximation (1,2) completely decouples the mechanics of fast-moving electrons and slow-moving nuclei, leading to satisfactory explanation of several experimental findings. On the contrary, complete or partial failure of this theoretical framework could be encountered for excited-state molecular processes such as charge-transfer reactions, measurements of scattering cross sections, and photochemical reactions. (3,4) In such phenomena, the non-negligible electron–nuclear coupling designated as nonadiabatic coupling terms (NACTs) not only plays a significant role for this discrepancy but also attains singular characteristics at degenerate points in the nuclear configuration space (CS). Therefore, overlooking of NACTs, (5,6) one of the major outcomes of BO treatment often leads to erroneous dynamical results and inaccurate numerical data regarding the molecular processes.
In the early sixties, Longuet-Higgins noticed multivaluedness of the eigenfunctions (7) while traversing along a closed path encircling a point of conical intersection (CI). Though this unusual behavior was circumvented by incorporating a phase factor (8) in an ad hoc manner, Mead and Truhlar (9) eliminated this multivaluedness behavior of wavefunctions by encompassing a vector potential into the nuclear Hamiltonian. On the other hand, the Hellmann–Feynman theorem predicts the possibility of singularity in NACTs (10,11) over the nuclear CS whenever electronic states appear to be degenerate or near degenerate. This singular characteristics can be removed by transforming to a different representation or basis, known as the diabatic representation where the singular kinetic coupling terms are converted into smoothly varying off-diagonal terms in the diabatic potential energy matrix. Such unitary transformation is known as the adiabatic-to-diabatic transformation (ADT). In this context, it may be noted that vector field created by the NACTs can be decomposed into longitudinal and transverse components, where the former is expressed as the derivative of a scalar and the latter by the curl of a vector. By performing ADT, the longitudinal component (the removable part) can be eliminated, (5,12) whereas the transverse or the nonremovable part may be neglected at the close vicinity of a CI.
Hobey and Mclachlan devised the first technique of ADT only for one degree of freedom and latter on, F. T. Smith implemented this method for a diatomic molecule. M. Baer proposed a general formalism for the diabatization of two coupled adiabatic potential energy surfaces (PESs) involved on the process of triatomic collision. In that work, the ADT condition is depicted which is the key mechanistic element for the determination of the ADT matrix elements by integrating the differential equations along a two-dimensional contour over the nuclear CS. (5,12) On the other hand, only the fulfillment of “curl condition” (12,13) guarantees the existence and the uniqueness of the solution to those equations. Generally, ADT angles can be calculated by integrating the NACTs around the CI point(s) and because of the quantized nature of the closed contour integrals, resulting angles will attain magnitudes in multiples of π (pi). (14) The satisfaction of aforementioned curl conditions also validates the presence of a set of electronic states forming a complete space. Though the above said condition is invalid at the degenerate CI point(s), nevertheless, the singular NACTs can be taken out from the SE if the line integrals exhibit quantization within the sub-Hilbert space constituted with all coupled electronic states. (15) The beyond BO (BBO) treatment for three, four, and five coupled electronic states in terms of electronic basis functions or the ADT matrix elements has been generalized by Adhikari et al., (16−19) and the explicit expressions of the NACTs, curl–divergence equations, curl conditions, and the diabatic PESs have also been formulated in terms of the ADT angles. Moreover, this theoretical development has been successfully employed on model (16−18,20) as well as realistic (19,21−24) systems for constructing smooth, continuous, and single-valued diabatic PESs. Molecular systems having strong vibronic interactions between five electronic states are omnipresent in nature. NO3 radical and C6H6+ (Bz+) radical cation are two highly intricate examples of five electronic state sub-Hilbert spaces. Recently, we have pursued formulation of BBO equations for five state sub-Hilbert spaces, devised the ADT equations and explicit expressions of NACTs in terms of ADT angles, and successfully implemented the formalism for NO3 radical and Bz+ radical cation.
In this article, our aim is to generate dressed adiabatic potential energy curves (PECs), dressed diabatic PECs, and adiabatic PECs via dressed diabatic potentials in a pair of normal mode coordinates for both the molecular/ionic species. The PECs carry the signature of topological effects in their functional variation along a particular normal mode coordinate by taking average over the other normal mode coordinate. Symmetric variation of dressed diabatic couplings indicates the absence of linear and higher odd order functions in electron–nuclear couplings, whereas asymmetric diabatic couplings incorporate both odd and even order polynomials in their functional forms. Detailed topological studies clearly reveal effects of NACTs in stabilization or destabilization of PESs.
NO3 radical is one of the most challenging molecular species in the realm of nonadiabatic dynamics. It exhibits highly complex spectral features due to intense vibronic coupling between its low-lying five electronic states 2A2 (12B2), Ã2E″ (12A2 and 12B1), and B̃2E′ (12A1 and 22B2) in the Franck–Condon region of nuclear CS. Several theoretical predictions were already proposed to elucidate the underlying principle of the complex spectral envelop, obtained from various experimental tools such as dispersed fluorescence spectroscopy, (25,26) Fourier transform infrared spectroscopy, (27,28) photoelectron spectroscopy, (29) and also from cavity ringdown spectroscopy. (30,31) In spite of various controversies about the ground state equilibrium geometry, Eisfeld and Morokuma (32) finally confirmed the radical to be of D3h symmetry. Mayer et al., (33) Faraji et al., (34) and Stanton et al. (35,36) also proposed some theoretical development to reveal the key features of the experimentally obtained spectral profile. It is noteworthy to mention that strong Jahn–Teller (JT) interactions prevail within the Ã2E″ and B̃2E′ states at the equilibrium D3h point along with several accidental CIs at C2v geometries. (19,37) In addition, the ground state, 2A2 experiences strong pseudo JT (PJT) interaction due to the B̃2E′ state. (19,37) These vibronic phenomena strongly influence the nuclear dynamics and the associated features of the photoelectron spectra (29,38) of the NO3– anion.
The photophysics of Bz+ radical cation is one of the center of attraction in the arena of BBO theory. In all theoretical developments, (39−47) the electronic states, namely, 2E1g, B̃2E2g, 2A2u, 2E1u, and Ẽ2B2u, along with the CIs among the PESs were investigated for its dynamical studies. Goode et al. (48) pointed out the complex vibrational pattern of B̃2E2g state using the combined effect of Herzberg–Teller and PJT coupling between this state and close-lying 2A2u state. JT coupling parameters for e2g vibrational modes of this radical cation which are used to illustrate the photoinduced Rydberg ionization spectra of the B̃2E2g state (49) were extensively measured by Johnson. (50) The low-lying doublet states of Bz+, B̃2E2g, and 2A2u show intense nonadiabatic interactions (39) among them. The JT effects within the ground, 2E1g as well as the excited state, B̃2E2g are quite predominant particularly lowering the symmetry of ground state equilibrium structure of Bz+ to D2h. (42,50) Apart from the JT effects, there are accidental CIs and PJT interactions prevailing between B̃2E2g and 2A2u states. (39) We have investigated the JT interactions of the ground (2E1g) and excited states (B̃2E2g) as well as accidental CIs within the higher excited states (B̃2E2g2A2u) using the ADT equations by employing the BBO formalism. (19)
The article is arranged as follows: Section 2 presents detailed theoretical framework of BBO theory for a five electronic state sub-Hilbert spaces. This section also gives a brief description of the underlying equations for calculating dressed potentials. In Section 3, we explore topological effects prevailing in both of the molecular systems. This detailed investigation will be highly useful to predict the possible functional forms of the inherent nonadiabatic couplings. Finally, the summary of this topological study is laid out in Section 4.

2. BBO Theory and Diabatic Hamiltonian

Click to copy section linkSection link copied!

As per the BO treatment, total molecular wavefunction can be expanded as a linear combination of the electronic wavefunctions where the combining coefficients represent the nuclear counterpart. While considering a sub-Hilbert space of finite dimension, a total electron–nuclei wavefunction takes the following form
(1)
This wavefunction is the eigenfunction of molecular SE, and the corresponding eigenvalue represents the total energy of the system
(2)
The molecular Hamiltonian can be partitioned into a nuclear kinetic energy operator (n) and the electronic Hamiltonian (Ĥe(se|sn))
(3)
where the nuclear kinetic energy operator is defined as
(4)
On the other hand, electronic wavefunctions (|ξi(se|sn)⟩), that is, the basis of the BO expansion, are eigenfunctions of Ĥe(se|sn) operator with a nuclear coordinate-dependent eigenvalue ui(sn). Therefore, the electronic eigenvalue equation is presented as below
(5)
When the BO expansion of molecular wavefunction is substituted in total Hamiltonian [eq 2] and it is projected with various electronic wavefunctions, the compact adiabatic (kinetically coupled) SE takes the following matrix form
(6)
where Uij = uiδij and τ⃗ is the nonadiabatic coupling matrix (NACM) defined as
(7)
The adiabatic representation of the SE is difficult to solve because of the numerical instability of the NACTs at the points of degeneracy. Therefore, we encounter the limitations of the adiabatic representations which exhibit the inevitable necessity of a different representation where the nonadiabatic interactions appear as continuous, smooth, and single-valued diabatic coupling terms. This transformation could be achieved by using an orthogonal rotation matrix (A)
(8)
where ψ and ϕd symbolize the adiabatic and diabatic nuclear wavefunction, respectively.
While incorporating the aforementioned form of wavefunction, diabatic representation of the nuclear SE takes the following expression
(9)
where
(10)
under the condition
(11)
known as the ADT condition.
The ADT condition [eq 11] can be used as the precursor of devising the ADT equations. For five (5) electronic state sub-Hilbert spaces, the model ADT matrix can be constructed by multiplying 5C2 rotation matrices in a particular way. Those could be arranged in 5C2! different ways, but we carry out the formulation for a particular order of multiplication. On the other hand, anyone of the possible arrangements produces same numerical value of the diabatic potential energy matrices. For any one of the rotation matrices, Aijij), where ii and jj elements become cosθij and ij and ji elements take the form of sinθij and −sinθij, respectively. Among the remaining elements, diagonal terms are one, whereas the off-diagonal terms acquire the magnitude of zero.
For five state sub-Hilbert space, the order of multiplication is taken as
(12)
When this model rotation matrix and NACM are incorporated in eq 11, a set of differential equations (ADT equations) can be formulated by solving the matrix form of ADT condition. Because of the absence of any analytic solution of these differential equations, they must be integrated numerically to obtain the ADT angles which are used to construct the ADT matrix at each and every grid points in the interested domain of the nuclear CS. During the transformation of adiabatic PESs to the diabatic one, a similarity transformation [eq 10] is carried out such that the NACTs take the form of diabatic coupling elements which appear in diagonal and off-diagonal elements of the diabatic potential energy matrix.
ADT equation for “1–2” coupling of five coupled electronic states can be formulated as (19)
(13)
If the adiabatic PESs are used for the simulation of the photoelectron spectrum of any molecule, the nonadiabatic interactions are overlooked and the inconsistency between the theoretically predicted and experimentally probed spectrum is maximum at the points of JT as well as accidental CIs. Diabatization not only removes the singularity of the NACTs but also can be used for carrying out nuclear dynamics to reproduce the photodetachment spectra (21) and to study the reactive scattering dynamics (51) to obtain integral cross sections.

3. Dressed Adiabatic, Dressed Diabatic, and Adiabatic-via-Dressed Diabatic PECs: Topological Effects

Click to copy section linkSection link copied!

In a polyatomic molecule, nonadiabatic couplings could exhibit multidimensional functional behavior over the nuclear CS due to the interactions of multiple electronic states through several vibrational degrees of freedom. In order to understand the contribution of each vibrational mode on the overall nonadiabatic interactions in a molecule, detailed investigation of topological effects is the most suitable approach. Moreover, it is readily realizable that functional features of PESs and NACTs defined over a multidimensional vector space cannot be displayed in the Euclidean space because of restriction in the number of independent axes.
In our present ab initio calculation, PESs and NACTs are evaluated as functions of two normal modes (Qi and Qj), keeping the other ones at their equilibrium. As a consequence, while constructing dressed potential curves (adiabatic as well as diabatic) and dressed diabatic couplings along a specific coordinate, the average effect of other mode can be embedded by integrating adiabatic and diabatic potential matrix elements over the second one. In this topological investigation, we have employed ab initio-based calculation to evaluate adiabatic PESs and NACTs and use first-principles-based transformation equations to construct diabatic PESs as a function of different pairs of normal modes, where the other modes are at equilibrium for each case. Such adiabatic and diabatic PESs are averaged over ground vibrational eigenfunctions of respective electronic states for a specific mode, and PECs are obtained to realize the topological effect originating from the other mode. The dimensionless mass-weighted normal mode coordinates can be defined as
(14)
where ωi and mk signify the vibrational frequency of ith vibrational mode and atomic masses, respectively. On the other hand, xki and xk0 represent Cartesian coordinates at distorted and equilibrium nuclear geometry.
The dressed adiabatic potentials () for ith mode and pth electronic state can be defined as
(15)
where up(Qj|Qi)s represent five adiabatic PESs and ζp,v=0a(Qj|Qi) signifies ground vibrational eigenfunction (v = 0) of the corresponding adiabatic PESs (p). Here, are the dressed adiabatic PESs along Qi where the other effects of other normal mode (Qj) are averaged out.
Similarly, dressed diabatic potentials () for ith normal mode and m,n electronic state can be expressed as
(16)
When this dressed diabatic energy matrix is diagonalized, adiabatic-via-dressed diabatic potential () is obtained.
Dressed adiabatic, diabatic PECs, and dressed diabatic couplings depict 1D functional form of potentials as well as couplings, where the effect of another mode is averaged out with respect to ground vibrational wavefunctions of different electronic states. It is needless to say that dressed PECs and diabatic couplings have nothing to do while studying photoelectron spectrum or performing reactive scattering calculation, but those functions can predict a lot of interesting facts about the structural features of the molecule and analytic expressions of diabatic couplings as a function of a specific coordinate. Indeed, such curves are highly useful to speculate nonadiabatic effects arising from a specific mode on the photoelectron spectrum or reactive cross sections of a molecular process. It is noteworthy to mention that not only the normal mode coordinates but also their associated polar counterparts (ρ and ϕ) may be employed to construct dressed PECs. We can definitely perform topological study along these polar coordinates to explore their contribution on nonadiabatic interactions, but the origination of JT couplings (either sole contribution of even order polynomials, or incorporation of both linear and higher order interactions) cannot be illustrated because of the absence of any symmetric/asymmetric variations of PECs along the ρ coordinate. Dressed adiabatic PECs cannot incorporate any effects of nonadiabatic coupling, whereas dressed diabatic potentials and diabatic coupling elements encompass diagonal and off-diagonal electron–nuclear couplings, respectively. While diagonalizing dressed diabatic matrix, effects of off-diagonal couplings are plugged into the newly generated adiabatic-via-dressed diabatic potentials which include both diagonal and off-diagonal nonadiabatic interactions.

4. Results and Discussion

Click to copy section linkSection link copied!

4.1. NO3 Radical

In the realm of nonadiabatic chemistry, NO3 radical appears to be an excellent representative system for five coupled electronic states. Several research groups have carried out thorough investigations to dig out exact equilibrium structure and complex features of photoelectron spectra. In order to achieve more structural insight, dressed potentials can be constructed by averaging out one of the normal mode coordinates. Low-lying five electronic states, namely, 2A2 (12B2), Ã2E″(12A2 and 12B1), and B̃2E′ (12A1 and 22B2), are mainly responsible for the inherent nonadiabatic interactions. Therefore, we employ polar analogue (ρ and ϕ) of two degenerate components of in-plane asymmetric stretching vibrational mode (ν3) (symbolized as Q3xQ3y plane) to perform the ab initio calculations. This radical has been optimized employing UCCSD(T) (a spin unrestricted coupled cluster method with all single and double excitations and perturbative accounts of triple excitations) level of calculation, and the optimized parameters are presented in Table 1. (37) Adiabatic PESs for five lowest electronic states (19) have been already constructed using state-averaged complete active space self-consistent field (CASSCF) calculations employing 6-31g** basis set and (9e, 8o) configuration active space (CAS). On the other hand, analytic NACTs (19) between the five low-lying electronic states have been obtained by implementing coupled-perturbed multi configuration space self-consistent field (CP-MCSCF) theory. In this nuclear plane, strong JT interactions are present within E″ as well as E′ states, forming “2–3” and “4–5” JT CIs at the equilibrium D3h point. Three additional accidental CIs can be observed within two sheets of E′ state at three equivalent C2v points (ρ ≈ 3.0 and ϕ = 30°, 150°, and 270°). Moreover, nonzero functional forms of τ14 and τ15 validate strong PJT interactions between ground state (A2) and the two sheets of E state labeled as “1–4” and “1–5” couplings.
Table 1. Optimized Geometry Parameters of NO3 Radical (37) with UCCSD(T) Method and cc-pVDZ Basis Set
parametersoptimization result (UCCSD(T))
bond length [N–O]1.237 Å
bond angle [∠O–N–O]120.0°
normal mode frequency (ν3)1121 cm-1
equilibrium energy–279.8442 a.u.
In order to exploit the inherent symmetry of the molecule (D3h), ADT equations are solved for the chosen pair of normal modes (Q3xQ3y) along the ρ coordinate for each ϕ grid. (19) Our aim is to construct dressed adiabatic, dressed diabatic, and adiabatic-via-dressed diabatic potentials along Q3x and Q3y normal mode coordinates. Dressed adiabatic and dressed diabatic PECs are plotted as functions of Qi by taking average over the other normal mode coordinate (Qj), whereas adiabatic-via-dressed diabatic potentials can be generated by diagonalizing the dressed diabatic energy matrix (WQj(Qi)). Figure 1 depicts symmetric variation of dressed adiabatic, dressed diabatic, and adiabatic-via-dressed diabatic potential energies along Q3x mode. On the other hand, PECs plotted in Figure 2 exhibit asymmetric variation while scanning along the Q3y mode. Such a different functional behavior originates if the component of the asymmetric stretching vibration does or does not produce equivalent geometry during positive and negative distortions. It can be observed that three PECs are widely separated in Figure 1a due to strong PJT couplings within ground A2 and E′ states. On the contrary, they almost coincide in Figure 1b and 1c because of weak JT interaction within E″ state. Finally, Figure 1d and 1e depict variation of third and fourth excited electronic states originated from JT and PJT interactions within the given range of Q3x normal mode coordinates. Again, Figure 2 represents asymmetric functional features of PECs along Q3y normal modes. It is noteworthy to mention that the Q3y normal mode generates nonequivalent geometries at both sides of the equilibrium structure (Q3y = 0). Moreover, dressed diabatic coupling elements are not only smooth, continuous, and single-valued but also exhibit similar functional characteristics of the diabatic potential. Figure 3 illustrates symmetric variation of dressed diabatic coupling terms [, , , and ] along Q3x coordinate where the functional forms of the dressed diabatic coupling elements are constituted only with even power polynomials of Q3x. Now, it is needless to say that JT interactions prevailing within E″ as well as E′ states do not originate from any linear coupling. Higher order electron-nuclear coupling plays the major role behind these JT interactions. On the other hand, Figure 4 represents different behaviors of electron-nuclear couplings along positive and negative directions of the Q3y normal mode. They loose their symmetric nature, and the corresponding functional form of coupling incorporates odd power terms.

Figure 1

Figure 1. Dressed adiabatic (), diabatic (), and adiabatic-via-dressed diabatic () PECs (n = 1–5) are depicted for (a) ground, (b) first excited, (c) second excited, (d) third excited, and (e) fourth excited states of NO3 radical along Q3x normal mode taking average over Q3y. All of the PECs exhibit symmetric variation and nonlinear functional characteristics.

Figure 2

Figure 2. Dressed adiabatic (), diabatic (), and adiabatic-via-dressed diabatic () PECs (n = 1–5) are represented for (a) ground, (b) first excited, (c) second excited, (d) third excited, and (e) fourth excited electronic states of NO3 along Q3y normal mode taking average over Q3x. Here, the PECs are asymmetric and incorporate linear functionalities.

Figure 3

Figure 3. Dressed diabatic coupling terms of NO3 radical along Q3x normal mode: (a) and (b) are smooth, symmetric as well as continuous appeared due to “2–3” and “4–5” JT CIs and three “4–5” accidental CIs. On the other hand, symmetric and single-valued (c) and (d) represent PJT couplings. Therefore, JT activity is originated solely because of nonlinear couplings along the Q3x mode.

Figure 4

Figure 4. Dressed diabatic coupling terms of NO3 radical along the Q3y normal mode. Here, (a) , (b) , (c) , and (d) all are smooth and continuous but the variation is asymmetric. Along Q3y normal mode, JT coupling contains linear as well as nonlinear functionalities.

4.2. Bz+ Radical Cation

Detailed ab initio study has been already carried out to explore the nonadiabatic interactions prevailing within five low-lying electronic states of the Bz+ radical cation. Further investigations are necessary to illustrate the origin and functional forms of dressed potentials as well as dressed diabatic coupling elements. In order to reveal those functional features, we have constructed three types of dressed PECs (adiabatic, diabatic, and adiabatic-via-dressed diabatic) and dressed diabatic coupling terms along two components of the degenerate in-plane asymmetric stretching mode (ν16) (Q16x and Q16y). This pairwise normal mode is selectively chosen because of its immense importance while depicting major inherent nonadiabaticities present in this radical cation. Before discussing our theoretical findings, a brief summary of the ab initio details (19) should be clearly mentioned for proper understanding of the newly constructed dressed PECs. Those first-principles-based quantum calculations for neutral species reveal the D6h symmetric structure at the equilibrium geometry, whereas the cationic species attends D2h point group due to intense JT distortion. In the case of Bz+ radical cation, geometry optimization has been carried out implementing B3LYP (Becke, three-parameter, Lee–Yang–Parr) calculation and the optimized parameters are supplied in Table 2. Ab initio adiabatic PESs have been constructed in CASSCF methodology for the lowest five doublet states of Bz+, viz, X̃2E1g (12B3g and 12B2g), B̃2E2g (12Ag and 12B1g), and C̃2A2u (12B1u) employing Gaussian basis set 6-31g** with an active space of 29 electrons distributed over 15 orbitals (29e, 15o). Moreover, analytic NACTs are obtained from CP-MCSCF methodology implemented in MOLPRO (52) quantum chemistry software. Because Q16x and Q16y produce strong JT interactions within E1g and E2g states at the D6h point (ρ = 0.0), lower sheets get more stabilized and the corresponding NACTs (τ12 and τ34) tend to blow up (at ρ = 0.0), validating the presence of JT CIs.
Table 2. Optimized Geometry Parameters of Bz+ Radical Cation with B3LYP Method and 6-311g(2d,p) Basis Set
parametersoptimization result (B3LYP)
bond length [C–C]1.382 Å
bond length [C–H]1.075 Å
bond angle [∠C–C–C]120.0°
bond angle [∠H–C–C]120.0°
normal mode frequency (ν16)1777 cm-1
equilibrium energy–230.7616 a.u.
In order to carry out diabatization of PESs, dimensionless mass-weighted normal mode coordinates are first converted into their polar counterparts (ρ and ϕ) to scan the entire adiabatic PESs and the NACTs. Our ultimate objective is to construct dressed adiabatic, dressed diabatic, and adiabatic-via-dressed diabatic potentials along Q16x and Q16y normal mode coordinates. Figure 5 displays symmetric functional forms of dressed PECs for four lowest electronic states along the Q16x normal mode, whereas the PECs exhibit asymmetric behavior along the Q16y mode (Figure 6). Likewise NO3 radical, such type of interesting characteristic features are originated because of the formation of equivalent and nonequivalent geometries along Q16x and Q16y coordinates, respectively. For the Bz+ radical cation, three types of PECs (dressed adiabatic, dressed diabatic, and adiabatic-via-dressed diabatic) along Q16x are largely separated from each other (Figure 5a–d) because of strong JT couplings. On the other hand, those PECs are exactly superimposed for fifth electronic state because of their noninvolvement in any nonadiabatic interactions along the Q16xQ16y pairwise mode. It is evident from Figure 5 that adiabatic-via-dressed diabatic curves obtained from diagonalization of dressed diabatic potential matrix exhibit one shallow minimum for ground and second excited electronic states (Figure 5a and 5c) and two shallow troughs for first and third excited ones (Figure 5b and 5d). On the other hand, asymmetric functional features of dressed PECs are exhibited along the Q16y mode (Figure 6) which signify the incorporation of both odd and even power polynomials in the analytic formulation. Functional forms of dressed diabatic coupling elements also display similar trends along the two normal modes. Symmetric variation of the dressed diabatic coupling terms are represented in Figure 7a and 7b, validating the contribution of only even power polynomials for electron–nuclear couplings in the strong symmetry breaking JT interactions. On the other hand, asymmetric dressed diabatic couplings are depicted along Q16y coordinate in Figure 8a and 8b, where the nonadiabatic interactions incorporate both odd and even power polynomials for couplings. Therefore, it can be concluded that JT interaction along this component (Q16y) shows both linear and nonlinear dependence.

Figure 5

Figure 5. Dressed adiabatic (), diabatic (), and adiabatic-via-dressed diabatic () PECs (n = 1–4) are plotted for low-lying four electronic states of Bz+ radical cation in diagrams (a–d), respectively, along Q16x normal mode taking average over the other component, Q16y. All PECs undergo symmetric variation and include only the even order functional forms.

Figure 6

Figure 6. Dressed adiabatic (), diabatic (), and adiabatic-via-dressed diabatic () PECs (n = 1–4) are portrayed for lowest four electronic states of Bz+ radical cation in diagrams (a–d), respectively, along the Q16y normal mode by averaging over Q16x. Here, PECs are asymmetric and incorporate both linear and nonlinear terms.

Figure 7

Figure 7. Dressed diabatic coupling terms are plotted along Q16x normal mode of the Bz+ radical cation. Figures (a,b) depict and , respectively. Both are symmetric, indicating nonlinear “1–2” and “3–4” JT interactions.

Figure 8

Figure 8. Dressed diabatic coupling terms are plotted with the variation of Q16y normal mode coordinates of the Bz+ radical cation. Figures (a,b) describe and , respectively. Along Q16y normal mode, JT coupling contains linear functions along with nonlinear terms.

5. Conclusions

Click to copy section linkSection link copied!

For the last few decades, several investigations have been carried out for highly complex scenario in the field of nonadiabatic dynamics of NO3 radical and Bz+ radical cation. In this perspective, an accurate investigation is necessary to explore the topological effects among the pairwise normal mode coordinates. Till now, nonadiabatic couplings have been extensively studied for two to five electronic state sub-Hilbert spaces (19,21−24) to generate highly accurate diabatic PESs. On the contrary, topological studies have been carried out only for two and three coupled electronic states (22−24) which is relevant for predicting the functional forms of dressed diabatic couplings. It is needless to say that in most cases, linear functions of normal mode coordinates are only responsible for JT stabilization, but we have already explored the sole contribution of even power couplings (24) for JT and RT CIs in the case of NO2 molecule. It has been clearly revealed that selection rule for the electronic transition in NO2 molecule originates mainly because of even power coupling elements. In this article, dressed adiabatic, dressed diabatic, and adiabatic-via-dressed diabatic electronic states are constructed for two real molecular systems of five electronic state sub-Hilbert space where one can explore the importance of even power coupling modes in JT interaction apart from the traditional ones involving both odd and even power interactions.
In Sections 2 and 3, the basic theoretical developments of BBO methodology as well as detailed background on the formulation of dressed PECs are presented explicitly and in the following section, importance of those PECs for NO3 radical and Bz+ radical cation is investigated. From the previous theoretical investigations, (19,37) it has been observed that NO3 possesses several kinds of nonadiabatic interactions, namely, JT CI, accidental CI and PJT coupling, whereas the Bz+ radical cation is deliberately chosen for its strong JT interaction. NO3 exhibits JT interaction within E″ as well as E′ states, accidental CIs between two sheets of E′ states along with substantial amount of PJT couplings between “1–4” and “1–5” states. Several groups have performed detailed study on the vibronic couplings to unfold its structural details and complex spectral features, but in this work, topological study for five coupled electronic states is carried out for the first time. Calculated dressed diabatic coupling terms exhibit symmetric and asymmetric functional behaviors along Q3x and Q3y modes, respectively though both of the normal modes are responsible for JT coupling. Similar trend is also observed for Q16x and Q16y modes of Bz+ radical cation. These modes are mainly responsible for JT CIs within E1g and E2g electronic states. Q16x component produces symmetric variation of PECs as well as dressed diabatic couplings, which clearly indicates the signature of higher order JT interactions originated from even power functions of Q16x coordinate. On the contrary, Q16y shows profound asymmetry in dressed PECs and dressed diabatic couplings. For the latter one, analytic forms of JT couplings contain both linear and nonlinear polynomials. Finally, we conclude that since this study validates our previous findings not only for lower but also for higher dimensional sub-Hilbert space, the effect of nonadiabaticity on molecular processes is a topological one.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Soumya Mukherjee - Department of Physical Chemistry, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700032, India
    • Bijit Mukherjee - Department of Physical Chemistry, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700032, India
    • Joy Dutta - Department of Physical Chemistry, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700032, India
    • Subhankar Sardar - Department of Chemistry, Bhatter College, Dantan, Paschim Medinipur, 721426, India
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

S.M. (file no.: SPM-07/080(0250)/2016-EMR-I) and B.M. (file no.: 09/080(0960)/2014-EMR-I) thank CSIR India for research fellowship, and J.D. acknowledges IACS for the same. S.S. thanks Principal, Bhatter College, Dantan for supplying research facility in his institution. S.A. is thankful to DST, India, through project no. EMR/2015/001314 for research funding. S.A. also acknowledges IACS for CRAY super computing facility.

References

Click to copy section linkSection link copied!

This article references 52 other publications.

  1. 1
    Born, M.; Oppenheimer, R. Zur Quantentheorie der Molekeln. Ann. Phys. 1927, 389, 457484,  DOI: 10.1002/andp.19273892002
  2. 2
    Born, M.; Huang, K. Dynamical Theory of Crystal Lattices; Oxford University Press: Oxford, 1954.
  3. 3
    Coe, J. D.; Martínez, T. J. Competitive Decay at Two- and Three-state Conical Intersections in Excited-State Intramolecular Proton Transfer. J. Am. Chem. Soc. 2005, 127, 45604561,  DOI: 10.1021/ja043093j
  4. 4
    Worth, G. A.; Robb, M. A.; Lasorne, B. Solving the time-dependent Schrödinger equation for nuclear motion in one step: direct dynamics of non-adiabatic systems. Mol. Phys. 2008, 106, 20772091,  DOI: 10.1080/00268970802172503
  5. 5
    Baer, M. Adiabatic and Diabatic Representations for Atom-Molecule Collisions: Treatment of the Collinear Arrangement. Chem. Phys. Lett. 1975, 35, 112118,  DOI: 10.1016/0009-2614(75)85599-0
  6. 6
    Top, Z. H.; Baer, M. Incorporation of Electronically Nonadiabatic Effects into Bimolecular Reactive Systems. I. Theory. J. Chem. Phys. 1977, 66, 13631371,  DOI: 10.1063/1.434032
  7. 7
    Longuet-Higgins, H. C. Some Recent Developments in the Theory of Molecular Energy Levels. Advances in Spectroscopy; H. W. Thompson, 1961; Vol. 2, p 429.
  8. 8
    Herzberg, G.; Longuet-Higgins, H. C. Intersection of Potential Energy Surfaces in Polyatomic Molecules. Discuss. Faraday Soc. 1963, 35, 7782,  DOI: 10.1039/df9633500077
  9. 9
    Mead, C. A.; Truhlar, D. G. On the Determination of Born-Oppenheimer Nuclear Motion Wave Functions Including Complications due to Conical Intersections and Identical Nuclei. J. Chem. Phys. 1979, 70, 22842296,  DOI: 10.1063/1.437734
  10. 10
    Hellmann, H. Einfuhrang in die Quantenchemie; Franz Duetiche: Leipzig, Germany, 1937.
  11. 11
    Feynman, R. P. Forces in Molecules. Phys. Rev. 1939, 56, 340343,  DOI: 10.1103/physrev.56.340
  12. 12
    Baer, M. Introduction to the Theory of Electronic Non-Adiabatic Coupling Terms in Molecular Systems. Phys. Rep. 2002, 358, 75142,  DOI: 10.1016/s0370-1573(01)00052-7
  13. 13
    Baer, M. Beyond Born–Oppenheimer: Conical Intersections and Electronic Nonadiabatic Coupling Terms; Wiley Interscience: NJ, 2006.
  14. 14
    Baer, M.; Englman, R. A Study of the Diabatic Electronic Representation within the Born-Oppenheimer Approximation. Mol. Phys. 1992, 75, 293303,  DOI: 10.1080/00268979200100231
  15. 15
    Mebel, A. M.; Halász, G. J.; Vibók, Á.; Alijah, A.; Baer, M. Quantization of the 3×3 nonadiabatic coupling matrix for three coupled states of the C2H molecule. J. Chem. Phys. 2002, 117, 9911000,  DOI: 10.1063/1.1483854
  16. 16
    Sarkar, B.; Adhikari, S. Extended Born-Oppenheimer Equation for a Three-State System. J. Chem. Phys. 2006, 124, 074101,  DOI: 10.1063/1.2170089
  17. 17
    Sarkar, B.; Adhikari, S. Curl Condition for a Four-State Born–Oppenheimer System Employing the Mathieu Equation. J. Phys. Chem. A 2008, 112, 98689885,  DOI: 10.1021/jp8029709
  18. 18
    Paul, A. K.; Sardar, S.; Sarkar, B.; Adhikari, S. Single Surface Beyond Born-Oppenheimer Equation for a Three-State Model Hamiltonian of Na3 Cluster. J. Chem. Phys. 2009, 131, 124312,  DOI: 10.1063/1.3236839
  19. 19
    Mukherjee, S.; Mukherjee, B.; Adhikari, S. Five Electronic State Beyond Born-Oppenheimer Equations and Their Applications to Nitrate and Benzene Radical Cation. J. Phys. Chem. A 2017, 121, 63146326,  DOI: 10.1021/acs.jpca.7b04592
  20. 20
    Paul, A. K.; Ray, S.; Mukhopadhyay, D.; Adhikari, S. Ab initio Calculations on the Excited States of Na3 Cluster to Explore Beyond Born-Oppenheimer Theories: Adiabatic to Diabatic Potential Energy Surfaces and Nuclear Dynamics. J. Chem. Phys. 2011, 135, 034107,  DOI: 10.1063/1.3609247
  21. 21
    Mukherjee, S.; Bandyopadhyay, S.; Paul, A. K.; Adhikari, S. Construction of Diabatic Hamiltonian Matrix from Ab Initio Calculated Molecular Symmetry Adapted Nonadiabatic Coupling Terms and Nuclear Dynamics for the Excited States of Na3 Cluster. J. Phys. Chem. A 2013, 117, 34753495,  DOI: 10.1021/jp311597c
  22. 22
    Mukherjee, S.; Adhikari, S. The Excited States of K3 Cluster: The Molecular Symmetry Adapted Non-Adiabatic Coupling Terms and Diabatic Hamiltonian Matrix. Chem. Phys. 2014, 440, 106118,  DOI: 10.1016/j.chemphys.2014.05.022
  23. 23
    Mukherjee, S.; Mukhopadhyay, D.; Adhikari, S. Conical Intersections and Diabatic Potential Energy Surfaces for the Three Lowest Electronic Singlet States of H3+. J. Chem. Phys. 2014, 141, 204306,  DOI: 10.1063/1.4901986
  24. 24
    Mukherjee, S.; Mukherjee, B.; Sardar, S.; Adhikari, S. Ab Initio Constructed Diabatic Surfaces of NO2 and the Photodetachment Spectra of its Anion. J. Chem. Phys. 2015, 143, 244307,  DOI: 10.1063/1.4938526
  25. 25
    Nelson, H. H.; Pasternack, L.; McDonald, J. R. Laser-Induced Excitation and Emission Spectra of Nitrate Radical (NO3). J. Phys. Chem. 1983, 87, 12861288,  DOI: 10.1021/j100231a003
  26. 26
    Ishiwata, T.; Fujiwara, I.; Naruge, Y.; Obi, K.; Tanaka, I. Study of Nitrate Radical by Laser-Induced Fluorescence. J. Phys. Chem. 1983, 87, 13491352,  DOI: 10.1021/j100231a016
  27. 27
    Ishiwata, T.; Tanaka, I.; Kawaguchi, K.; Hirota, E. Infrared diode laser spectroscopy of the NO3 ν3 band. J. Chem. Phys. 1985, 82, 21962205,  DOI: 10.1063/1.448362
  28. 28
    Hirota, E.; Ishiwata, T.; Kawaguchi, K.; Fujitake, M.; Ohashi, N.; Tanaka, I. Near-infrared Band of the Nitrate Radical NO3 Observed by Diode Laser Spectroscopy. J. Chem. Phys. 1997, 107, 28292838,  DOI: 10.1063/1.474641
  29. 29
    Weaver, A.; Arnold, D. W.; Bradforth, S. E.; Neumark, D. M. Examination of the 2A2′ and 2E′ states of NO3 by ultraviolet photoelectron spectroscopy of NO3. J. Chem. Phys. 1991, 94, 17401751,  DOI: 10.1063/1.459947
  30. 30
    Deev, A.; Sommar, J.; Okumura, M. Cavity ringdown spectrum of the forbidden Ã2E″ ← 2A2′ transition of NO3: Evidence for static Jahn-Teller distortion in the à state. J. Chem. Phys. 2005, 122, 224305,  DOI: 10.1063/1.1897364
  31. 31
    Codd, T.; Chen, M.-W.; Roudjane, M.; Stanton, J. F.; Miller, T. A. Jet cooled cavity ringdown spectroscopy of the Ã2E″ ← 2A2′ transition of the NO3 radical. J. Chem. Phys. 2015, 142, 184305,  DOI: 10.1063/1.4919690
  32. 32
    Eisfeld, W.; Morokuma, K. A Detailed Study on the Symmetry Breaking and its Effect on the Potential Surface of NO3. J. Chem. Phys. 2000, 113, 55875597,  DOI: 10.1063/1.1290607
  33. 33
    Mayer, M.; Cederbaum, L. S.; Köppel, H. Ground State Dynamics of NO3: Multimode Vibronic Borrowing Including Thermal Effects. J. Chem. Phys. 1994, 100, 899911,  DOI: 10.1063/1.466572
  34. 34
    Faraji, S.; Köppel, H.; Eisfeld, W.; Mahapatra, S. Towards a higher-order description of Jahn-Teller coupling effects in molecular spectroscopy: The state of NO3. Chem. Phys. 2008, 347, 110119,  DOI: 10.1016/j.chemphys.2007.10.006
  35. 35
    Okumura, M.; Stanton, J. F.; Deev, A.; Sommar, J. New insights into the Jahn-Teller effect in NO3 via the dark Ã2E″ state. Phys. Scr. 2006, 73, C64C70,  DOI: 10.1088/0031-8949/73/1/n12
  36. 36
    Stanton, J. F. On the Vibronic Level Structure in the NO3 Radical. I. The Ground Electronic State. J. Chem. Phys. 2007, 126, 134309,  DOI: 10.1063/1.2715547
  37. 37
    Mukherjee, B.; Mukherjee, S.; Sardar, S.; Shamasundar, K. R.; Adhikari, S. An Ab Initio Investigation of Non-Adiabatic Couplings and Conical Intersections Among the Lowest Five Electronic States of the NO3 Radical. Mol. Phys. 2017, 115, 28332848,  DOI: 10.1080/00268976.2017.1340680
  38. 38
    Mukherjee, B.; Mukherjee, S.; Sardar, S.; Shamasundar, K. R.; Adhikari, S. A Beyond Born-Oppenheimer Treatment of Five State Molecular System NO3 and the Photodetachment Spectra of its Anion. Chem. Phys. 2018 (manuscript in press). DOI: 10.1016/j.chemphys.2018.09.017
  39. 39
    Köppel, H.; Cederbaum, L. S.; Domcke, W. Interplay of Jahn-Teller and pseudo-Jahn-Teller vibronic dynamics in the benzene cation. J. Chem. Phys. 1988, 89, 20232040,  DOI: 10.1063/1.455100
  40. 40
    Köppel, H. New Ultrafast Nonradiative Decay Mechanism in the Benzene Radical Cation. Chem. Phys. Lett. 1993, 205, 361370,  DOI: 10.1016/0009-2614(93)87135-p
  41. 41
    Döscher, M.; Köppel, H. Multiple surface intersections and strong nonadiabatic coupling effects between the 2E1u and 2B2u states of C6H6+. Chem. Phys. 1997, 225, 93105,  DOI: 10.1016/s0301-0104(97)00228-0
  42. 42
    Döscher, M.; Köppel, H.; Szalay, P. G. Multistate Vibronic Interactions in the Benzene Radical Cation. I. Electronic Structure Calculations. J. Chem. Phys. 2002, 117, 26452656,  DOI: 10.1063/1.1491397
  43. 43
    Köppel, H.; Döscher, M.; Bâldea, I.; Meyer, H.-D.; Szalay, P. G. Multistate Vibronic Interactions in the Benzene Radical Cation. II. Quantum Dynamical Simulations. J. Chem. Phys. 2002, 117, 26572671,  DOI: 10.1063/1.1491398
  44. 44
    Köppel, H.; Bâldea, I.; Szalay, P. G. Combined Jahn-Teller and Pseudo-Jahn-Teller Effects in the Benzene Radical Cation. Advances in Quantum Chemistry; Academic Press, 2003; Vol. 44, pp 199217.
  45. 45
    Sardar, S.; Paul, A. K.; Sharma, R.; Adhikari, S. A “Classical Trajectory Driven Nuclear Dynamics by a Parallelized Quantum-Classical Approach to a Realistic Model Hamiltonian of Benzene Radical Cation. Int. J. Quantum Chem. 2009, 111, 27412759,  DOI: 10.1002/qua.22578
  46. 46
    Sardar, S.; Paul, A. K.; Sharma, R.; Adhikari, S. The Multistate Multimode Vibronic Dynamics of Benzene Radical Cation with a Realistic Model Hamiltonian Using a Parallelized Algorithm of the Quantumclassical Approach. J. Chem. Phys. 2009, 130, 144302,  DOI: 10.1063/1.3108488
  47. 47
    Sardar, S.; Adhikari, S. A quantum-classical simulation of a multi-surface multi-mode nuclear dynamics on C6H6+ incorporating degeneracy among electronic states. J. Chem. Sci. 2012, 124, 5158,  DOI: 10.1007/s12039-011-0195-z
  48. 48
    Goode, J. G.; Hofstein, J. D.; Johnson, P. M. The observation of strong pseudo-Jahn-Teller activity in the benzene cation 2E2g state. J. Chem. Phys. 1997, 107, 17031716,  DOI: 10.1063/1.474526
  49. 49
    Johnson, P. M. The Jahn-Teller effect in the lower electronic states of benzene cation. II. Vibrational analysis and coupling constants of the 2E2g state. J. Chem. Phys. 2002, 117, 1000110007,  DOI: 10.1063/1.1519007
  50. 50
    Johnson, P. M. The Jahn-Teller Effect in the Lower Electronic States of Benzene Cation. I. Calculation of Linear Parameters for the e2g Modes. J. Chem. Phys. 2002, 117, 999110000,  DOI: 10.1063/1.1519006
  51. 51
    Ghosh, S.; Mukherjee, S.; Mukherjee, B.; Mandal, S.; Sharma, R.; Chaudhury, P.; Adhikari, S. Beyond Born-Oppenheimer theory for ab initio constructed diabatic potential energy surfaces of singlet H3+ to study reaction dynamics using coupled 3D time-dependent wave-packet approach. J. Chem. Phys. 2017, 147, 074105,  DOI: 10.1063/1.4998406
  52. 52
    Werner, H.-J.; MOLPRO , version 2010.1, a Package of Ab initio Programs, 2010; see http://www.molpro.net.

Cited By

Click to copy section linkSection link copied!

This article is cited by 10 publications.

  1. Koushik Naskar, Soumya Mukherjee, Bijit Mukherjee, Satyam Ravi, Saikat Mukherjee, Subhankar Sardar, Satrajit Adhikari. ADT: A Generalized Algorithm and Program for Beyond Born–Oppenheimer Equations of “N” Dimensional Sub-Hilbert Space. Journal of Chemical Theory and Computation 2020, 16 (3) , 1666-1680. https://doi.org/10.1021/acs.jctc.9b00948
  2. Soumya Mukherjee, Koushik Naskar, Saikat Hazra, Mantu Kumar Sah, Satrajit Adhikari. Beyond Born-Oppenheimer Treatment for Multi-State Photoelectron Spectra, Phase Transitions of Solids and Scattering Processes. Journal of Physics: Conference Series 2024, 2769 (1) , 012012. https://doi.org/10.1088/1742-6596/2769/1/012012
  3. Mantu Kumar Sah, Soumya Mukherjee, Swagato Saha, Koushik Naskar, Satrajit Adhikari. Photoelectron spectra of benzene: Can path dependent diabatic surfaces provide unique observables?. The Journal of Chemical Physics 2023, 159 (24) https://doi.org/10.1063/5.0177186
  4. Saikat Hazra, Soumya Mukherjee, Satyam Ravi, Subhankar Sardar, Satrajit Adhikari. Construction of Beyond Born‐Oppenheimer Based Diabatic Surfaces and Generation of Photoabsorption Spectra: The Touchstone Pyrazine (C 4 N 2 H 4 ). ChemPhysChem 2022, 23 (23) https://doi.org/10.1002/cphc.202200482
  5. Soumya Mukherjee, Satyam Ravi, Joy Dutta, Subhankar Sardar, Satrajit Adhikari. Beyond Born–Oppenheimer based diabatic surfaces of 1,3,5-C 6 H 3 F 3 + to generate the photoelectron spectra using time-dependent discrete variable representation approach. Physical Chemistry Chemical Physics 2022, 24 (4) , 2185-2202. https://doi.org/10.1039/D1CP04733G
  6. Soumya Mukherjee, Satyam Ravi, Koushik Naskar, Subhankar Sardar, Satrajit Adhikari. A beyond Born–Oppenheimer treatment of C6H6+ radical cation for diabatic surfaces: Photoelectron spectra of its neutral analog using time-dependent discrete variable representation. The Journal of Chemical Physics 2021, 154 (9) https://doi.org/10.1063/5.0040361
  7. Joy Dutta, Soumya Mukherjee, Koushik Naskar, Sandip Ghosh, Bijit Mukherjee, Satyam Ravi, Satrajit Adhikari. The role of electron–nuclear coupling on multi-state photoelectron spectra, scattering processes and phase transitions. Physical Chemistry Chemical Physics 2020, 22 (47) , 27496-27524. https://doi.org/10.1039/D0CP04052E
  8. Bijit Mukherjee, Koushik Naskar, Soumya Mukherjee, Sandip Ghosh, Tapas Sahoo, Satrajit Adhikari. Beyond Born–Oppenheimer theory for spectroscopic and scattering processes. International Reviews in Physical Chemistry 2019, 38 (3-4) , 287-341. https://doi.org/10.1080/0144235X.2019.1672987
  9. Soumya Mukherjee, Joy Dutta, Bijit Mukherjee, Subhankar Sardar, Satrajit Adhikari. Conical intersections and nonadiabatic coupling terms in 1,3,5-C6H3F3+: A six state beyond Born-Oppenheimer treatment. The Journal of Chemical Physics 2019, 150 (6) https://doi.org/10.1063/1.5064519
  10. Bijit Mukherjee, Sandip Ghosh, Satrajit Adhikari. Beyond Born-Oppenheimer treatment on spectroscopic and scattering processes. Journal of Physics: Conference Series 2018, 1148 , 012001. https://doi.org/10.1088/1742-6596/1148/1/012001

ACS Omega

Cite this: ACS Omega 2018, 3, 10, 12465–12475
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.8b01648
Published October 3, 2018

Copyright © 2018 American Chemical Society. This publication is licensed under these Terms of Use.

Article Views

784

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Dressed adiabatic (), diabatic (), and adiabatic-via-dressed diabatic () PECs (n = 1–5) are depicted for (a) ground, (b) first excited, (c) second excited, (d) third excited, and (e) fourth excited states of NO3 radical along Q3x normal mode taking average over Q3y. All of the PECs exhibit symmetric variation and nonlinear functional characteristics.

    Figure 2

    Figure 2. Dressed adiabatic (), diabatic (), and adiabatic-via-dressed diabatic () PECs (n = 1–5) are represented for (a) ground, (b) first excited, (c) second excited, (d) third excited, and (e) fourth excited electronic states of NO3 along Q3y normal mode taking average over Q3x. Here, the PECs are asymmetric and incorporate linear functionalities.

    Figure 3

    Figure 3. Dressed diabatic coupling terms of NO3 radical along Q3x normal mode: (a) and (b) are smooth, symmetric as well as continuous appeared due to “2–3” and “4–5” JT CIs and three “4–5” accidental CIs. On the other hand, symmetric and single-valued (c) and (d) represent PJT couplings. Therefore, JT activity is originated solely because of nonlinear couplings along the Q3x mode.

    Figure 4

    Figure 4. Dressed diabatic coupling terms of NO3 radical along the Q3y normal mode. Here, (a) , (b) , (c) , and (d) all are smooth and continuous but the variation is asymmetric. Along Q3y normal mode, JT coupling contains linear as well as nonlinear functionalities.

    Figure 5

    Figure 5. Dressed adiabatic (), diabatic (), and adiabatic-via-dressed diabatic () PECs (n = 1–4) are plotted for low-lying four electronic states of Bz+ radical cation in diagrams (a–d), respectively, along Q16x normal mode taking average over the other component, Q16y. All PECs undergo symmetric variation and include only the even order functional forms.

    Figure 6

    Figure 6. Dressed adiabatic (), diabatic (), and adiabatic-via-dressed diabatic () PECs (n = 1–4) are portrayed for lowest four electronic states of Bz+ radical cation in diagrams (a–d), respectively, along the Q16y normal mode by averaging over Q16x. Here, PECs are asymmetric and incorporate both linear and nonlinear terms.

    Figure 7

    Figure 7. Dressed diabatic coupling terms are plotted along Q16x normal mode of the Bz+ radical cation. Figures (a,b) depict and , respectively. Both are symmetric, indicating nonlinear “1–2” and “3–4” JT interactions.

    Figure 8

    Figure 8. Dressed diabatic coupling terms are plotted with the variation of Q16y normal mode coordinates of the Bz+ radical cation. Figures (a,b) describe and , respectively. Along Q16y normal mode, JT coupling contains linear functions along with nonlinear terms.

  • References


    This article references 52 other publications.

    1. 1
      Born, M.; Oppenheimer, R. Zur Quantentheorie der Molekeln. Ann. Phys. 1927, 389, 457484,  DOI: 10.1002/andp.19273892002
    2. 2
      Born, M.; Huang, K. Dynamical Theory of Crystal Lattices; Oxford University Press: Oxford, 1954.
    3. 3
      Coe, J. D.; Martínez, T. J. Competitive Decay at Two- and Three-state Conical Intersections in Excited-State Intramolecular Proton Transfer. J. Am. Chem. Soc. 2005, 127, 45604561,  DOI: 10.1021/ja043093j
    4. 4
      Worth, G. A.; Robb, M. A.; Lasorne, B. Solving the time-dependent Schrödinger equation for nuclear motion in one step: direct dynamics of non-adiabatic systems. Mol. Phys. 2008, 106, 20772091,  DOI: 10.1080/00268970802172503
    5. 5
      Baer, M. Adiabatic and Diabatic Representations for Atom-Molecule Collisions: Treatment of the Collinear Arrangement. Chem. Phys. Lett. 1975, 35, 112118,  DOI: 10.1016/0009-2614(75)85599-0
    6. 6
      Top, Z. H.; Baer, M. Incorporation of Electronically Nonadiabatic Effects into Bimolecular Reactive Systems. I. Theory. J. Chem. Phys. 1977, 66, 13631371,  DOI: 10.1063/1.434032
    7. 7
      Longuet-Higgins, H. C. Some Recent Developments in the Theory of Molecular Energy Levels. Advances in Spectroscopy; H. W. Thompson, 1961; Vol. 2, p 429.
    8. 8
      Herzberg, G.; Longuet-Higgins, H. C. Intersection of Potential Energy Surfaces in Polyatomic Molecules. Discuss. Faraday Soc. 1963, 35, 7782,  DOI: 10.1039/df9633500077
    9. 9
      Mead, C. A.; Truhlar, D. G. On the Determination of Born-Oppenheimer Nuclear Motion Wave Functions Including Complications due to Conical Intersections and Identical Nuclei. J. Chem. Phys. 1979, 70, 22842296,  DOI: 10.1063/1.437734
    10. 10
      Hellmann, H. Einfuhrang in die Quantenchemie; Franz Duetiche: Leipzig, Germany, 1937.
    11. 11
      Feynman, R. P. Forces in Molecules. Phys. Rev. 1939, 56, 340343,  DOI: 10.1103/physrev.56.340
    12. 12
      Baer, M. Introduction to the Theory of Electronic Non-Adiabatic Coupling Terms in Molecular Systems. Phys. Rep. 2002, 358, 75142,  DOI: 10.1016/s0370-1573(01)00052-7
    13. 13
      Baer, M. Beyond Born–Oppenheimer: Conical Intersections and Electronic Nonadiabatic Coupling Terms; Wiley Interscience: NJ, 2006.
    14. 14
      Baer, M.; Englman, R. A Study of the Diabatic Electronic Representation within the Born-Oppenheimer Approximation. Mol. Phys. 1992, 75, 293303,  DOI: 10.1080/00268979200100231
    15. 15
      Mebel, A. M.; Halász, G. J.; Vibók, Á.; Alijah, A.; Baer, M. Quantization of the 3×3 nonadiabatic coupling matrix for three coupled states of the C2H molecule. J. Chem. Phys. 2002, 117, 9911000,  DOI: 10.1063/1.1483854
    16. 16
      Sarkar, B.; Adhikari, S. Extended Born-Oppenheimer Equation for a Three-State System. J. Chem. Phys. 2006, 124, 074101,  DOI: 10.1063/1.2170089
    17. 17
      Sarkar, B.; Adhikari, S. Curl Condition for a Four-State Born–Oppenheimer System Employing the Mathieu Equation. J. Phys. Chem. A 2008, 112, 98689885,  DOI: 10.1021/jp8029709
    18. 18
      Paul, A. K.; Sardar, S.; Sarkar, B.; Adhikari, S. Single Surface Beyond Born-Oppenheimer Equation for a Three-State Model Hamiltonian of Na3 Cluster. J. Chem. Phys. 2009, 131, 124312,  DOI: 10.1063/1.3236839
    19. 19
      Mukherjee, S.; Mukherjee, B.; Adhikari, S. Five Electronic State Beyond Born-Oppenheimer Equations and Their Applications to Nitrate and Benzene Radical Cation. J. Phys. Chem. A 2017, 121, 63146326,  DOI: 10.1021/acs.jpca.7b04592
    20. 20
      Paul, A. K.; Ray, S.; Mukhopadhyay, D.; Adhikari, S. Ab initio Calculations on the Excited States of Na3 Cluster to Explore Beyond Born-Oppenheimer Theories: Adiabatic to Diabatic Potential Energy Surfaces and Nuclear Dynamics. J. Chem. Phys. 2011, 135, 034107,  DOI: 10.1063/1.3609247
    21. 21
      Mukherjee, S.; Bandyopadhyay, S.; Paul, A. K.; Adhikari, S. Construction of Diabatic Hamiltonian Matrix from Ab Initio Calculated Molecular Symmetry Adapted Nonadiabatic Coupling Terms and Nuclear Dynamics for the Excited States of Na3 Cluster. J. Phys. Chem. A 2013, 117, 34753495,  DOI: 10.1021/jp311597c
    22. 22
      Mukherjee, S.; Adhikari, S. The Excited States of K3 Cluster: The Molecular Symmetry Adapted Non-Adiabatic Coupling Terms and Diabatic Hamiltonian Matrix. Chem. Phys. 2014, 440, 106118,  DOI: 10.1016/j.chemphys.2014.05.022
    23. 23
      Mukherjee, S.; Mukhopadhyay, D.; Adhikari, S. Conical Intersections and Diabatic Potential Energy Surfaces for the Three Lowest Electronic Singlet States of H3+. J. Chem. Phys. 2014, 141, 204306,  DOI: 10.1063/1.4901986
    24. 24
      Mukherjee, S.; Mukherjee, B.; Sardar, S.; Adhikari, S. Ab Initio Constructed Diabatic Surfaces of NO2 and the Photodetachment Spectra of its Anion. J. Chem. Phys. 2015, 143, 244307,  DOI: 10.1063/1.4938526
    25. 25
      Nelson, H. H.; Pasternack, L.; McDonald, J. R. Laser-Induced Excitation and Emission Spectra of Nitrate Radical (NO3). J. Phys. Chem. 1983, 87, 12861288,  DOI: 10.1021/j100231a003
    26. 26
      Ishiwata, T.; Fujiwara, I.; Naruge, Y.; Obi, K.; Tanaka, I. Study of Nitrate Radical by Laser-Induced Fluorescence. J. Phys. Chem. 1983, 87, 13491352,  DOI: 10.1021/j100231a016
    27. 27
      Ishiwata, T.; Tanaka, I.; Kawaguchi, K.; Hirota, E. Infrared diode laser spectroscopy of the NO3 ν3 band. J. Chem. Phys. 1985, 82, 21962205,  DOI: 10.1063/1.448362
    28. 28
      Hirota, E.; Ishiwata, T.; Kawaguchi, K.; Fujitake, M.; Ohashi, N.; Tanaka, I. Near-infrared Band of the Nitrate Radical NO3 Observed by Diode Laser Spectroscopy. J. Chem. Phys. 1997, 107, 28292838,  DOI: 10.1063/1.474641
    29. 29
      Weaver, A.; Arnold, D. W.; Bradforth, S. E.; Neumark, D. M. Examination of the 2A2′ and 2E′ states of NO3 by ultraviolet photoelectron spectroscopy of NO3. J. Chem. Phys. 1991, 94, 17401751,  DOI: 10.1063/1.459947
    30. 30
      Deev, A.; Sommar, J.; Okumura, M. Cavity ringdown spectrum of the forbidden Ã2E″ ← 2A2′ transition of NO3: Evidence for static Jahn-Teller distortion in the à state. J. Chem. Phys. 2005, 122, 224305,  DOI: 10.1063/1.1897364
    31. 31
      Codd, T.; Chen, M.-W.; Roudjane, M.; Stanton, J. F.; Miller, T. A. Jet cooled cavity ringdown spectroscopy of the Ã2E″ ← 2A2′ transition of the NO3 radical. J. Chem. Phys. 2015, 142, 184305,  DOI: 10.1063/1.4919690
    32. 32
      Eisfeld, W.; Morokuma, K. A Detailed Study on the Symmetry Breaking and its Effect on the Potential Surface of NO3. J. Chem. Phys. 2000, 113, 55875597,  DOI: 10.1063/1.1290607
    33. 33
      Mayer, M.; Cederbaum, L. S.; Köppel, H. Ground State Dynamics of NO3: Multimode Vibronic Borrowing Including Thermal Effects. J. Chem. Phys. 1994, 100, 899911,  DOI: 10.1063/1.466572
    34. 34
      Faraji, S.; Köppel, H.; Eisfeld, W.; Mahapatra, S. Towards a higher-order description of Jahn-Teller coupling effects in molecular spectroscopy: The state of NO3. Chem. Phys. 2008, 347, 110119,  DOI: 10.1016/j.chemphys.2007.10.006
    35. 35
      Okumura, M.; Stanton, J. F.; Deev, A.; Sommar, J. New insights into the Jahn-Teller effect in NO3 via the dark Ã2E″ state. Phys. Scr. 2006, 73, C64C70,  DOI: 10.1088/0031-8949/73/1/n12
    36. 36
      Stanton, J. F. On the Vibronic Level Structure in the NO3 Radical. I. The Ground Electronic State. J. Chem. Phys. 2007, 126, 134309,  DOI: 10.1063/1.2715547
    37. 37
      Mukherjee, B.; Mukherjee, S.; Sardar, S.; Shamasundar, K. R.; Adhikari, S. An Ab Initio Investigation of Non-Adiabatic Couplings and Conical Intersections Among the Lowest Five Electronic States of the NO3 Radical. Mol. Phys. 2017, 115, 28332848,  DOI: 10.1080/00268976.2017.1340680
    38. 38
      Mukherjee, B.; Mukherjee, S.; Sardar, S.; Shamasundar, K. R.; Adhikari, S. A Beyond Born-Oppenheimer Treatment of Five State Molecular System NO3 and the Photodetachment Spectra of its Anion. Chem. Phys. 2018 (manuscript in press). DOI: 10.1016/j.chemphys.2018.09.017
    39. 39
      Köppel, H.; Cederbaum, L. S.; Domcke, W. Interplay of Jahn-Teller and pseudo-Jahn-Teller vibronic dynamics in the benzene cation. J. Chem. Phys. 1988, 89, 20232040,  DOI: 10.1063/1.455100
    40. 40
      Köppel, H. New Ultrafast Nonradiative Decay Mechanism in the Benzene Radical Cation. Chem. Phys. Lett. 1993, 205, 361370,  DOI: 10.1016/0009-2614(93)87135-p
    41. 41
      Döscher, M.; Köppel, H. Multiple surface intersections and strong nonadiabatic coupling effects between the 2E1u and 2B2u states of C6H6+. Chem. Phys. 1997, 225, 93105,  DOI: 10.1016/s0301-0104(97)00228-0
    42. 42
      Döscher, M.; Köppel, H.; Szalay, P. G. Multistate Vibronic Interactions in the Benzene Radical Cation. I. Electronic Structure Calculations. J. Chem. Phys. 2002, 117, 26452656,  DOI: 10.1063/1.1491397
    43. 43
      Köppel, H.; Döscher, M.; Bâldea, I.; Meyer, H.-D.; Szalay, P. G. Multistate Vibronic Interactions in the Benzene Radical Cation. II. Quantum Dynamical Simulations. J. Chem. Phys. 2002, 117, 26572671,  DOI: 10.1063/1.1491398
    44. 44
      Köppel, H.; Bâldea, I.; Szalay, P. G. Combined Jahn-Teller and Pseudo-Jahn-Teller Effects in the Benzene Radical Cation. Advances in Quantum Chemistry; Academic Press, 2003; Vol. 44, pp 199217.
    45. 45
      Sardar, S.; Paul, A. K.; Sharma, R.; Adhikari, S. A “Classical Trajectory Driven Nuclear Dynamics by a Parallelized Quantum-Classical Approach to a Realistic Model Hamiltonian of Benzene Radical Cation. Int. J. Quantum Chem. 2009, 111, 27412759,  DOI: 10.1002/qua.22578
    46. 46
      Sardar, S.; Paul, A. K.; Sharma, R.; Adhikari, S. The Multistate Multimode Vibronic Dynamics of Benzene Radical Cation with a Realistic Model Hamiltonian Using a Parallelized Algorithm of the Quantumclassical Approach. J. Chem. Phys. 2009, 130, 144302,  DOI: 10.1063/1.3108488
    47. 47
      Sardar, S.; Adhikari, S. A quantum-classical simulation of a multi-surface multi-mode nuclear dynamics on C6H6+ incorporating degeneracy among electronic states. J. Chem. Sci. 2012, 124, 5158,  DOI: 10.1007/s12039-011-0195-z
    48. 48
      Goode, J. G.; Hofstein, J. D.; Johnson, P. M. The observation of strong pseudo-Jahn-Teller activity in the benzene cation 2E2g state. J. Chem. Phys. 1997, 107, 17031716,  DOI: 10.1063/1.474526
    49. 49
      Johnson, P. M. The Jahn-Teller effect in the lower electronic states of benzene cation. II. Vibrational analysis and coupling constants of the 2E2g state. J. Chem. Phys. 2002, 117, 1000110007,  DOI: 10.1063/1.1519007
    50. 50
      Johnson, P. M. The Jahn-Teller Effect in the Lower Electronic States of Benzene Cation. I. Calculation of Linear Parameters for the e2g Modes. J. Chem. Phys. 2002, 117, 999110000,  DOI: 10.1063/1.1519006
    51. 51
      Ghosh, S.; Mukherjee, S.; Mukherjee, B.; Mandal, S.; Sharma, R.; Chaudhury, P.; Adhikari, S. Beyond Born-Oppenheimer theory for ab initio constructed diabatic potential energy surfaces of singlet H3+ to study reaction dynamics using coupled 3D time-dependent wave-packet approach. J. Chem. Phys. 2017, 147, 074105,  DOI: 10.1063/1.4998406
    52. 52
      Werner, H.-J.; MOLPRO , version 2010.1, a Package of Ab initio Programs, 2010; see http://www.molpro.net.