Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

A time-independent, variational method for studying the photodissociation of triatomic molecules

Marco Pezzella a, Georgi Mitev a, Sergei N. Yurchenko a, Jonathan Tennyson *a and Alexander O. Mitrushchenkov b
aDepartment of Physics & Astronomy, University College London, London WC1E 6BT, UK. E-mail: j.tennyson@ucl.ac.uk
bMSME, Université Gustave Eiffel, CNRS UMR 8208, Univ Paris Est Creteil, F-77474 Marne-la-Vallée, France

Received 12th July 2024 , Accepted 14th October 2024

First published on 18th October 2024


Abstract

The photodissociation of molecules is becoming an increasingly important factor to consider in the evolution of exoplanets' atmospheres orbiting around UV-rich stars, as it leads to the enrichment of atmospheric complexity. A new method is developed for computing the rotationally and vibrationally resolved photodissociation spectrum of triatomic molecules. The time-independent Schrödinger equation is solved using the variational nuclear motion program EVEREST; a new code EXOCSMOOTH is employed to compute the cross-sections by applying Gaussian smoothing to a set of discrete transitions into the continuum. HCN is chosen as the test molecule, as it has been widely studied in the literature. Results are compared with the available experiments. Temperature dependence is explored for temperatures up to 2000 K.


1 Introduction

Understanding the photochemistry of planetary atmospheres is becoming extremely important in this new revolutionary era of astronomy. The recent discovery of photochemically produced SO2 in WASP-39b is an example where photochemical processes, including photodissociation, play a crucial role in explaining the complexity of atmosphere composition and dynamics, as well as non-LTE effects.1,2 Exoplanets orbiting near M-dwarfs, small oxygen-rich stars with temperatures between 2300 and 3800 K which are the commonest stars in our galaxy, experience high UV fluxes of photons from their host stars.3–6 Due to the planets’ proximity to the star, they have temperatures on the order of magnitude of thousands Kelvin, generating vibrationally and rotationally excited molecules. Challenges arise in collecting photodissociation cross-sections at these high temperatures,7 due to the lack of experimental and computational data. Experiments struggle to reach the temperatures required to model the top of hot atmospheres, and calculations often assume a harmonic ground state wave function.8 As a reflection of these two factors, standard databases contain data for molecules at interstellar temperatures (close to 0 K).9,10

In recent years, cross-section calculations have begun to include rovibrational excitation of molecules as a function of temperature.11 The ExoMol database12,13 was designed to produce comprehensive line lists of hot bound–bound transitions for molecules present in exoplanets’ atmospheres and is expanding to include photodissociation data. Previous studies14 have demonstrated that the photodissociation of diatomic molecules can be simulated using the program Duo,15 which was originally developed to solve the bound–bound nuclear motion problem for diatomics allowing for different types of couplings and crossings between potential energy curves. Duo uses a grid-based variational solution of the Schrödinger equation based on the use of a sinc-DVR basis which has been shown to be highly accurate16 and which results in a discretised representation of the continuum. ExoCross17 is used to generate the raw cross-section for different grid sizes. The results obtained using different radial grids and smoothing the results with an appropriate Gaussian function generate spectra comparable with those of experiments. This methodology was successfully applied to produce photodissociation line lists for HCl and HF;18 these results were subsequently corroborated by Qin et al.19

The transition from two to three atoms adds complexity to the computations involved in generating photodissociation cross-sections, not least because the continuum now becomes multi-dimensional. Determining the best set of internal coordinates for the given potential energy surfaces and defining the transition dipole moments are non-trivial tasks that depend on the chemical species under examination. HCN was selected as the initial molecule to assess the generation of photodissociation cross-sections for triatomic molecules as at lower energies it has a clear set of photodissociation products: H + CN. HCN has an ubiquitous presence in space; it is detected in the interstellar medium,20–24 comets,25–27 planets,28–30 exoplanets31 and, finally, in the atmosphere of Titan,32 where it is supposed to be a main reservoir of carbon.33,34 The HCN/HNC ratio is very sensitive to stellar UV flux35,36 and can be used as a tracer for the heating of the nebular molecular gas by UV photons.37–39

The photochemistry of HCN has undergone extensive examination, encompassing both experimental and computational approaches. Experiments explored dissociation from the predissociative 2 1A′ ← 1 1A′ transitions at various wavelengths, successfully reconstructing the potential dissociation mechanism.40–43 Dissociation events from 1 1A′′, 2 1A′′, and 2 1A′ states involve the cleavage of the H–CN bond through bending progression.44 Analysis of the electronic state distribution of the CN fragment, originating from the bending of the 3 A′ ← 1 A′ transition, indicates potential intersystem crossing and/or internal conversion between various excited states.45 This hypothesis has been corroborated by measurements from 1A′(vCH = 3) to the 1 1A′′ electronic states: the lack of rotational population suggests that the bond breaks when HCN is in the linear arrangement.46

The ground state potential energy surface (PES) has been computed by various research groups,47–52 whereas surfaces from the excited states are relatively scarce. The first study of excited states was published by Perić et al.53 in 1987. This work was followed by studies performed at the University of New Mexico between 2001 and 2003 on predissociation from the 1 1A′′ (ref. 54–56) and 2 1A′ (ref. 57) electronic states. Vibrational predissociation resonances in the latter state accelerate the dissociation due to efficient energy flow facilitated by an accidental coincidence between H–C and C–N vibrational normal modes. Nayak and collaborators calculated a variety of excited singlet states,58 followed by the work from Chenel et al.,59 which was used to explore the rates in the interstellar medium.60 The triplets states have been recently observed by Priyadarshini and coworkers.61

An alternative method to generate photodissociation cross-sections is through the use of the time-dependent Schrödinger equation. In this approach, the cross-section is proportional to the Fourier transform of the time-dependent autocorrelation function of the time evolution of the wave packet in an excited electronic state. Among all time-dependent methodologies, the multi-configurational time-dependent Hartree (MCTDH) approach, developed by Meyer et al.,62 stands out as one of the most versatile methods, as demonstrated by multiple improvements to the original code and its various implementations.63,64 The first photodissociation application appeared, for NOCl65 and NO2,66 two years after its introduction. These studies were quickly followed by investigations of the photodissociation of larger molecules, such as methyl iodide, which involved five nuclear degrees of freedom across three excited electronic states.67,68 The recent perspective from Han et al.69 discusses the most up-to-date applications of the method.

Although acknowledging the advantages of time-dependent methods for studying photodissociation, we opt for a time-independent approach for three reasons: (1) the ability to generate state-to-state photodissociation spectra, which are useful for modelling non-LTE effects, (2) the convenience of applying it to indirect photodissociation processes, as it is less time-intensive than the time-dependent approach70 and (3) it allows the use of a single, often complicated, model for all photon absorption processes,71 which includes the calculation of photoabsorption cross-sections across an extended wavelength range.72,73

In this work, we explore the ability to extend the approach used for diatomics to study continuum processes in larger systems. The DVR3D74 and EVEREST75 codes were chosen because they have been successfully used to describe bound–bound electronic transitions in previous studies, as demonstrated by Zak et al.76 and Owens et al.,77 respectively.

This paper is organised as follows: in the Methods section, a description and tests of our approach, the potential energy surface used, the choice of the basis, and the cross-section calculation are discussed. In the Results section, we compared our cross-section and the integrated intensities with previous experiments and calculations. Finally, the temperature dependence of photoabsorption is explored for three different temperatures.

2 Methods

2.1 Potential energy and transition dipole moment surfaces

Potential energy surfaces for this work are taken from the literature. The ground state has been taken from the study by Makhnev et al.52 and chosen for its spectroscopic accuracy when compared to experiments. The PESs and transition dipole moment surfaces (TDMS) for the excited A′ states have been provided through private communication with Chenel et al. (16ChRoAg).59 The potential energy surfaces are represented using the bond-length-bond-angle (BLBA) representation with the following the convention: r1 = rCH, r2 = rCN and θ = HĈN. The transition dipole moments are oriented such that the z-axis is aligned to rCN, the x-axis is perpendicular to the z-axis inside the molecular plane, and the y-axis component, which has a zero contribution, is perpendicular to the molecular plane. Fig. 1 shows the molecular coordinates and the reference axis of the transition dipole moment.
image file: d4cp02771j-f1.tif
Fig. 1 BLBA coordinates of HCN are used in this work. The reference axis, shown below the molecule, indicates the transition dipole moment orientation.

Transitions from the 1 1A′ ground electronic state to the 2 1A′, 3 1A′, 4 1A′, and 5 1A′ excited states are considered. 2 1A′ and 3 1A′ states are mainly dissociative states with a bound region near the linear structure due to an energy barrier along the rCH coordinate for both surfaces. No transitions between excited states are included because of their large excitation energies. The PESs are illustrated in Fig. 2. Details of these PESs are shown in Table 1.


image file: d4cp02771j-f2.tif
Fig. 2 Cuts through the potential energy surfaces of the low-lying singlet electronic states of HCN considered in this work as a function of rCH (up) and Θ (bottom).
Table 1 Linear configuration mapping, excitation energies (T0) and vertical excitation energies (ΔE) for the HCN electronic states considered in this work. T0 values in parenthesis are from 16ChRoAG59
State Linear configuration T 0 (cm−1) ΔE (cm−1)
1 1A′ X1Σ+ 3418 (3468) 0
2 1A′ 11Δ 54[thin space (1/6-em)]040 (55[thin space (1/6-em)]491) 70[thin space (1/6-em)]241
3 1A′ 11Π 65[thin space (1/6-em)]916 (65[thin space (1/6-em)]815) 75[thin space (1/6-em)]162
4 1A′ 21Π 75[thin space (1/6-em)]744 82[thin space (1/6-em)]534
5 1A′ 31Π 85[thin space (1/6-em)]389 86[thin space (1/6-em)]203


2.2 Choice of the basis

Calculations of the photoabsorption cross-section were performed using DVR3D74 and EVEREST.75 The original DVR3D program suite, based on the use of the discrete variable representation (DVR) basis, was modified by Zak et al.76 to calculate rovibronic transitions and tested for SO2. The program suite EVEREST was used to solve the Schrödinger equation using a sinc-DVR basis.78,79 The EVEREST input parameters were tuned to reproduce the energy levels published by Makhnev et al.52 for the J = 0 vibrational energies below 10[thin space (1/6-em)]000 cm−1, the two models agree with an RMSE (root-mean-squares-error) of 0.012 cm−1.

For DVR3D, the basis set parameters used for the ground state and excited states must be the same; this is a necessary requirement for calculating the overlap between the two wave functions. DVR3D offers two forms of radial basis functions which can be optimised for a given system: Morse-like oscillators80 and spherical oscillators;81 standard ground state HCN calculations such as those of Makhnev et al.52 use Morse-like oscillators for both radial coordinates. Here, to allow for dissociation, we experimented using spherical oscillators for the H–CN coordinate in Jacobi coordinates because they have better space filling properties. However, unlike the sinc-DVR basis which fill a finite, well defined region, the spherical oscillators in principle fill all space but the representation becomes increasingly sparse at large separations. Numerical experiments showed that, while it was possible to recover some of the photodissociation behaviours using spherical oscillators, this was less satisfactory than the results using sinc-DVR basis functions, which have the additional advantage that one can perform repeat calculations with shifted box sizes and obtain useful information.71,82 The sinc-DVR basis successfully reproduced the energy levels and transitions observed in HCN, see below. The reason behind the different behaviour of the two basis functions is the way the tridimensional grid is built: the DVR basis set is constructed using a non-uniform, in principle semi-infinite Gaussian quadrature, while the sinc-DVR basis uses a finite and uniform grid. Fig. 3 shows how the distribution of the vibrational levels changes between a spherical oscillator DVR and a sinc-DVR basis for the 3 1A′ dissociative state. The two basis gave different results for energies above 70[thin space (1/6-em)]000 cm−1. The divergence is due to the spherical oscillator DVR basis being unconstrained in the radial coordinates: the basis explores larger regions which means that attempting convergence by increasing the number of points simply results in the production of an increasing, potentially infinite, number of states. For this reason, all calculations in this paper were performed using EVEREST.


image file: d4cp02771j-f3.tif
Fig. 3 The first 600 vibrational energy levels of HCN (3 1A′). For energies below 70[thin space (1/6-em)]000 cm−1, spherical oscillator DVR and sinc-DVR basis results show a similar trend, but they diverge for higher energies.

2.3 Calculation of the photoabsorption cross-section

In this section, we provide a brief overview of the procedure and steps for generating cross-sections using EVEREST. A more detailed explanation of the code and underlying algorithms can be found in the original publication by Mitrushchenkov.75

The first step required to calculate cross-sections is to generate the vibrational basis set for the electronic states of interest. This is done using the evvib.e routine. For each individual state, the corresponding symmetry is specified in terms of the electronic spin (S), the molecular term symbol (A), and the projection of the orbital angular momentum along the internuclear axis (Λ). The atomic compositions, molecular masses and coordinate systems are then specified. For each coordinate, the basis representation, the grid interval and the number of points need to be specified. The bond length coordinates are described using a sinc-DVR basis using the desired number of points (Nrpoints), ranging from rmin to rmax. The bond angle is described by image file: d4cp02771j-t1.tif points using Legendre polynomials ranging from 0 to 180 degrees. This produces the vibrational levels for each k (projection of the angular momentum on the z-axis) of interest for a Hamiltonian of the desired size.83

The second step consists of generating the rotational wave function using evrot.e. In this routine, the wave function for the desired Js is generated. The rotational embedding is specified in this step.

In the third step, the dipole moment transition matrix is generated with the evdip.e routine, using the x and z-components of the TDMS. The routine evrli.e combines the wave function with the transition dipole moments. The evrs.e routine generates the molecular spectrum. The output is then transformed and stored in the ExoMol format.13,84

A significant advantage of EVEREST is that potentials, dipole moments, and possible couplings are loaded as dynamic libraries, without the need to recompile the code for each calculation.

The EVEREST parameters used in the present study are reported in Table 2. The first section of the table presents the parameters for the vibrational input. Each state considered in this work is a singlet with A′ symmetry, while Λ varies depending on the state considered. Atom names and (atomics) masses are the input with the central atom last. Each calculation was made with a Hamiltonian of size 6000, specifying the first 100 roots for the ground state and the first 1000 roots for the electronically excited states. In the rotational input, z-embedding is specified with the z-axis lying along the rCN bond. For both rCH and rCN, a radial grid spanning from 0.6 Å to 3.5 Å is used to take into account wavefunctions for highly excited vibrational states. At this step, all energy levels below 104 cm−1 for the ground state and below 105 cm−1 for the excited states are selected. The input runs through the first N J and K roots. The overall workflow is shown in Fig. 4.

Table 2 Input parameters used in the EVEREST code for HCN
Vibrational input
N pots 2
State 1 symmetry 1 A′ Λ1
State 2 symmetry 1 A′ Λ2
Atoms H N C
Masses [Da] 1.007825
14.003074
12.000000
Geometry BLBA
r 1 Grid type Sinc-DVR
r 1 N points 50
r 1 Interval [Å] 0.6
3.5
r 2 Grid type Sinc-DVR
r 2 N points 50
r 2 Interval [Å] 0.6
3.5
Θ Grid type Legendre
Θ N points 50
Θ Interval [degrees] 0
180
K min 0
K max N
Hamiltonian size 6000
Number of roots 100
1000
Rotational input
J interval 0-N
Embedding z-Embedding
Embedding axis r x
Diagonalisation Jaco
Hamiltonian size 6000
E maxstate [ cm−1] 10[thin space (1/6-em)]000
100[thin space (1/6-em)]000
K min 0
k max N
Dipole input
Number of dipole files 2
Alx 1
Embending z-Embedding
Embending axis r x



image file: d4cp02771j-f4.tif
Fig. 4 Graphical representation of the workflow used in this paper. For each step, the program used in EVEREST is written in parenthesis in an italic character.

All calculations are performed in the UCL Department of Physics and Astronomy theory cluster using an Intel compiler and an OpenMP parallelisation scheme on 16 nodes.

2.4 Cross-section smoothing

With a sinc-DVR basis, one performs a calculation over a finite region (box) defined by internuclear distances rmin, and rmax which effectively places infinite potential barriers at the boundaries. For diatomic molecules,14 it is computationally cheap to repeat the calculation using a sinc-DVR basis generated with an increasing box size. This can form the basis of a stabilization method85–87 and is being used by the ExoMol group to study predissociation.71,73,82 The choice of the box size sets boundary conditions and controls the density resulting in a discretized continuum spectrum. By adjusting the size of the box by varying rmax, one can sample different areas of the continuum and produce cross-sections with a dependence on the calculation of the box size. For diatomics, it is computationally possible to perform hundreds or thousands of nuclear motion calculations71 as individual calculations are relatively quick, requiring, typically, less than an hour of computing time per iteration. This is not the case with triatomic molecules or larger polyatomics, where calculations can take days, weeks, or months per iteration, depending on the size and complexity of the model. As such, performing stabilization-style calculations is currently computationally prohibitive.

A practical alternative to the stabilization method is to compute a discretized continuum spectrum from a single-box calculation and smooth the corresponding temperature-dependent cross-sections, σ(ν), by applying a uniform Gaussian line profile with an appropriate half-width at half-maximum (HWHM, α). This is, at the moment, the only viable method for producing these cross-sections for polyatomics using EVEREST. The choice of α can be constrained by the structure of the discretized spectrum and the shape of the continuum cross-section.

Following Pezzella et al.,14 the initial cross-sections of 1 1A′→ U electronic transitions are produced with α = 10 cm−1. Here, U represents any given unbound electronic state. These are the cross-sections which undergo smoothing.

To quantify optimum smoothness, we require that the resulting smoothed cross-section must be continuous and that the minimum value of α is used to prevent the over-distribution of intensity to the wings.

To assess the continuity of the cross-sections, one notes that, at their intensity maxima, the cross-sections have isolated turning points. In this region, the intensity distribution is bell shaped and, in general, direct photodissociation to purely repulsive states appears to have this shape as found by Pezzella et al.14,18

As such, the smoothing of the cross-section can be constrained by the number of expected maxima, NM. The continuity condition is so defined to be that the resulting, smoothed cross-section should have a number of roots in its first and second differential equal to NM and NM + 1, respectively, hence, should be twice differentiable.

It is found that there are usually a range of α values that meet this condition, and so the minimum value is taken to satisfy the condition on minimum smoothing. Eqn (1) expresses these conditions formally:

 
image file: d4cp02771j-t2.tif(1)
where N(f) is the number of roots as a function f. Due to the discrete nature of this problem, one cannot implement this procedure as a standard minimization problem through treatments like gradient descent and instead it must be solved iteratively.

The numerical method behind this is to establish a range of α values over which to test. At each iteration, the first and second forward derivatives are computed as follows:

 
image file: d4cp02771j-t3.tif(2)
where s is the sampling window which is chosen to deal with the numerical noise of producing these differentials. In general, a value of image file: d4cp02771j-t4.tif, where image file: d4cp02771j-t5.tif is the size of the wavenumber grid, is found to be appropriate for both first and second derivatives. Roots are counted by scanning the available smoothed cross-section and finding where the sign of the derivatives change. In areas of low intensity in the initial cross-section, small changes in intensity result in large changes of the gradient. While roots in these regions are not false positives, it is impractical to count them. To counteract this, a threshold value, K, is imposed such that a root will only be counted if the smoothed cross-section is [small sigma, Greek, circumflex](ν;α) >K.

Each iteration checks in the number of counted roots which meet the threshold conditions, upon the continuity condition being met or all test values are used. An example showing how the increasing α parameter affects the derivatives and shape of a toy model's cross-section is shown in Fig. 5. As can be seen, both α = 1000 cm−1 and α = 10[thin space (1/6-em)]000 cm−1 meet the continuity condition imposed by eqn (1); however, the shape of the cross-section in the α = 10[thin space (1/6-em)]000 cm−1 case does not faithfully represent the shape of the raw cross-section in the top left panel; therefore, one must include the requirement for minimal smoothing. In this dataset, the threshold value was estimated as K = max(σ)/10.


image file: d4cp02771j-f5.tif
Fig. 5 Toy model: Dependence of the shape of a photodissociation cross-section and its first and second derivatives on α in the smoothing procedure. Left panels show the smoothed cross-section, middle panels are the first derivatives, and right panels are the second derivatives.

The benefits of this procedure are that it is simple to implement, requires minimal computational expense, and provides good results in terms of smoothing. A limitation of this method is that upon meeting the aforementioned criteria, smoothing does not tend to be sufficient at higher ν. As ν increases, the gaps between successive vibrational bands also increases and so a minimal smoothing parameter as defined above would still allow for discrete structures to show through at high ν. One could increase α to account for this; however, this results in a slight over-smoothing in the lower ν region and is more challenging to automate reliably. This is a minor limitation, however, as the temperature dependent effects of these cross-sections will manifest themselves primarily in the lower wavenumber regions.

A Python3 code, ExoSmooth, has been written to automate the procedure of optimizing the HWHM. The program applies Gaussian profiles to the unsmoothed cross-section, starting at a given HWHM and increasing the HWHM by some user-defined amounts at each iteration until the continuity conditions described above are met. This ensures that the minimum amount of smoothing possible is applied to the cross-section. Indeed, as the Gaussian profile is normalised, the integral of the smoothed and unsmoothed cross-sections is equivalent; hence, photodissociation rates calculated with either cross-section are also equivalent. The documented code is made available from https://github.com/exomol. We note that the test performed as part of our initial study on diatomics14 showed that photodissociation rates obtained with the smoothed and unsmoothed data were the same, and that comparisons with (rotation-free) full photodissociation calculations also gave essentially the same photodissociation rate at low temperatures.

3 Results

3.1 Comparison with experiments

Of the four excited states available, there are no experiments available for the 2 A′ ← 1 A′ transition; the 3 A′ ← 1 A′ transition has been explored by different authors,45,88,89 thanks to its well defined structure and being isolated from other electronic states; the 4 A′ ← 1 A′ and 5 A′ ← 1 A′ transitions are localised in the same energy region, making difficult to separate the two contributions in the absorption spectrum.45,89,90

The 3 A′ is the only excited state which shows a clear vibrational progression. These levels make it possible to evaluate the quality of the potential used and the overall validity of the model. The quality of our calculations is tested against the computational results of 16ChRoAg and the experiments of Nuth et al. (82NuGl).89 The first two models are computed including only the J = 1 ← 0 transitions, which effectively models T = 0 K. The results are shown in Fig. 6. The peak positions agree, in terms of relative intensities between the bending progression, with the measurements of 82NuGl, while the calculations of 16ChRoAg tend to be blue shifted with respect to the other sources. The cross-section magnitudes agree with 16ChRoAg up to v2 = 6, while, for v2 = 7,8 we observe a major splitting between the K components compared to the other two methods, leading to a lower cross-section peaks than 16ChRoAg. The splitting becomes evident for v2 ≥ 9 in all sources. A clear drop of the cross-section is observed for energies higher than the 3 A′ dissociation barrier at 138.84 nm (72[thin space (1/6-em)]025 cm−1). The absorption spectrum presents both bound–bound features, due to the bending progression, together with an increasing dissociative character raising from dissociative threshold at wavelengths shorter than 138.84 nm.


image file: d4cp02771j-f6.tif
Fig. 6 Comparison between our photoabsorption spectrum for the 3 A′ ← 1 A′ transition at T = 0 K and J = 0,1, with the computed model from 16ChRoAg59 and the experiments from 82NuGl.89 The blue dashed line separates the bound–bound spectrum at longer wavelengths from the photodissociation region at shorter wavelengths.

Table 3 reports measured vibrational band centers for the 3 A′(0,v2,0) ← 1 A′(0,0,0) bands from MacPherson & Simons (78MaSi),88 Lee (80Lee),45 82NuGl and our values at T = 0 K. The experimental vibrational values vary by 10 to 70 cm−1 between the different measurements. In our calculations, the ground state transition is shifted by 46 cm−1 from 80Lee with a RMSE of 68 cm−1. The RMSE is 64 cm−1 compared to that of 78Masi, while a lower RMSE of 47 cm−1 is found when comparing the measurements of 82NuGl.

Table 3 Position for the peaks of the HCN (31A′ (0, v2, 0) ← [X with combining tilde]′Σ+ (0,0,0)). Experimental values are from 78MaSi,88 80Lee45 and 82NuGl.89 Transition frequencies and the RMSE of the observed calculated results are in cm−1
v 2 ν 78MaSi ν 80Lee ν 82NuGl ν This work
0 65[thin space (1/6-em)]647 65[thin space (1/6-em)]693
1 66[thin space (1/6-em)]520 66[thin space (1/6-em)]691
2 67[thin space (1/6-em)]376 67[thin space (1/6-em)]434
3 68[thin space (1/6-em)]227 68[thin space (1/6-em)]217 68[thin space (1/6-em)]194
4 69[thin space (1/6-em)]085 69[thin space (1/6-em)]085 69[thin space (1/6-em)]080 68[thin space (1/6-em)]997
5 69[thin space (1/6-em)]950 69[thin space (1/6-em)]940 69[thin space (1/6-em)]935 69[thin space (1/6-em)]864
6 70[thin space (1/6-em)]750 70[thin space (1/6-em)]746 70[thin space (1/6-em)]746 70[thin space (1/6-em)]701
7 71[thin space (1/6-em)]515 71[thin space (1/6-em)]515 71[thin space (1/6-em)]515 71[thin space (1/6-em)]516
8 72[thin space (1/6-em)]280 72[thin space (1/6-em)]290 72[thin space (1/6-em)]296 72[thin space (1/6-em)]276
9 73[thin space (1/6-em)]035 73[thin space (1/6-em)]041 73[thin space (1/6-em)]046 73[thin space (1/6-em)]047
10 73[thin space (1/6-em)]770 73[thin space (1/6-em)]746 73[thin space (1/6-em)]741 73[thin space (1/6-em)]738
11 74[thin space (1/6-em)]570 74[thin space (1/6-em)]555 74[thin space (1/6-em)]566 74[thin space (1/6-em)]600
12 75[thin space (1/6-em)]370 75[thin space (1/6-em)]432 75[thin space (1/6-em)]438 75[thin space (1/6-em)]501
13 76[thin space (1/6-em)]214 76[thin space (1/6-em)]220 76[thin space (1/6-em)]294
14 76[thin space (1/6-em)]938 76[thin space (1/6-em)]966
RMSE 64 68 47


Fig. 7 compares a computed spectrum of the 140 nm band, calculated at 300 K with 0 < J ≤ 50, with the experiments. A partition function of 899.7 calculated using this rotational interval is in agreement within 99% with the values reported in the ExoMol database as taken from the Harris line list for HCN91 at this temperature, suggesting a near complete solution. The ExoMol partition function is based on rovibrational energies with Jmax = 60. The magnitudes of the cross-sections are lower than those measured by 78MaSi and 80Lee suggesting that the 16ChRoAg transition dipoles are too small, while the cross-sections from 82NuGl are consistently higher than those in all experiments. For higher vibrational levels, we observe that the computed spectra show a higher level of rotational splitting than experiments. The vibrational progression of v2 = 8 at 137.74 nm is spread along a 1 nm interval with a large splitting between the rotational components.


image file: d4cp02771j-f7.tif
Fig. 7 Photoabsorption spectra of the 3 A′ ← 1 A′ transition. Our computed spectrum (0 < J < 25) is in black, experimental spectra are in blue for 78MaSi,88 green for 80Lee,45 and gold for 82NuGl.89 The photodissociation region is found at wavelengths shorter than 138.84 nm.

The 2 A′ state corresponds to one component of the 1Δ state at linear configurations, making transitions to this state forbidden at linear geometries and weak for bent structures. The intensities are of the order of 10−20 cm−1, being between 2 and 3 order of magnitudes lower than the other transitions. Fig. 8 shows our computed spectrum between 145 and 175 nm when using 1000 roots. A small vibrational progression is observed for wavelengths longer than 155 nm, followed by a bound-free region below this limit.


image file: d4cp02771j-f8.tif
Fig. 8 Photoabsorption spectra of the 2A′ ← 1 A′ transition. The spectrum is computed at T = 0 K. The photodissociation regime is found for wavelengths below 155 nm.

Table 4 and Fig. 9 show our results for the 4 A′ ← 1 A′ and the 5 A′ ← 1 A′ bands compared to the experiments from 81NaKoOz and 82NuGl. Our peak assignments are made by performing calculations at T = 0 K. The experiments, however, report only the vibrational levels without information on the rotational levels. A small shift between the energy levels can be expected, as at room temperature the maximum of the cross-section will be found for higher rotational states. 81NaKoOz identifies 6 vibrational states within the two electronic state, and we found that our levels are consistently higher than theirs with an RMSE of 35.1 cm−1. We get better agreement with the 17 levels measured by 82NuGl, with a RMSE of 14.0 cm−1. A larger disagreement is found for the v2 = 2 state, with a difference of 157 cm−1. Excluding this state, the RMSE decreases to 10.0 cm−1. However, we get poor agreement with the magnitudes of the cross-sections; our computed magnitudes are lower by a factor of 5–10 in the 100–130 nm region. It is likely that the computed TDMSs are too small.

Table 4 Vibrational band centers 4 A′ ← 1 A′ and 5 A′ ← 1 A′ bands: Comparison between theory (this work) and experiment, from 82NuGl89 and 81NaKoOz90
v 1 v 2 v 3 81NaKoOz 82NuGl This work
4 A′ ← 1 A′
000 82[thin space (1/6-em)]100 82[thin space (1/6-em)]217 82[thin space (1/6-em)]243
020 83[thin space (1/6-em)]222 83[thin space (1/6-em)]227
001 83[thin space (1/6-em)]960 83[thin space (1/6-em)]991 84[thin space (1/6-em)]041
002 85[thin space (1/6-em)]948 86[thin space (1/6-em)]000
5 A′ ← 1 A′
000 88[thin space (1/6-em)]980 89[thin space (1/6-em)]063 89[thin space (1/6-em)]013
010 89[thin space (1/6-em)]718 89[thin space (1/6-em)]666
020 90[thin space (1/6-em)]170 90[thin space (1/6-em)]400 90[thin space (1/6-em)]243
001 90[thin space (1/6-em)]930 90[thin space (1/6-em)]992 91[thin space (1/6-em)]025
011 91[thin space (1/6-em)]659 91[thin space (1/6-em)]706
100 92[thin space (1/6-em)]070 92[thin space (1/6-em)]039 92[thin space (1/6-em)]113
021 92[thin space (1/6-em)]396 92[thin space (1/6-em)]508
110 92[thin space (1/6-em)]653 92[thin space (1/6-em)]682
002 92[thin space (1/6-em)]963 92[thin space (1/6-em)]936
031 93[thin space (1/6-em)]101 93[thin space (1/6-em)]090
120 93[thin space (1/6-em)]396 93[thin space (1/6-em)]384
130 94[thin space (1/6-em)]171 94[thin space (1/6-em)]159
200 95[thin space (1/6-em)]057 95[thin space (1/6-em)]057



image file: d4cp02771j-f9.tif
Fig. 9 Bound–bound photoabsorption spectra of the 4 A′ ← 1 A′ (black) and 5 A′ ← 1 A′ (pink) transitions versus the experiments using a logarithmic y-scale. Experimental spectra are in green for 80Lee, orange for 81 NaKoOz and gold for 82NuGl. The computed spectra are consistently lower than that measured by the experiments.

3.2 Integrated intensities

The calculation of the integrated intensity (I) is a second method to assess the quality of our calculations. The method itself is the integration of the photodissociation cross-sections over a specific wavelength range (λ1 and λ2):
 
image file: d4cp02771j-t6.tif(3)

The MPI-Mainz UV/VIS Spectral Atlas (ref. 92) contains 4 different measured electronic spectra of HCN. The intensities of the 3 A′ ← 1 A′ are integrated from the measurement from 78MaSi88 and 80Lee.45 Upper state excitations are measured from 80Lee, 81NaKoOz90 and 82NuGl.89 The Mainz database simply reports the overall spectrum over the experimental wavelength range, making it impossible to make a one to one comparison with our results in terms of individual transitions.

The integrated intensities for the 3 A′ ← 1 A′ bands from our model (1.57 × 10−14 cm molecule−1) and the spectrum extracted from 16ChRoAg59 (1.40 × 10−14 cm molecule−1) are within 12%, showing overall agreement between the two computational models. For these calculations, only the J = 1 ← 0 transitions are considered. At room temperature (300 K), the calculated band intensity, with J varying from 0 to 25, is 3.01 × 10−14 cm molecule−1. Experiments from 78MaSi, 80Lee, and 82NuGl show higher intensities than the computed ones, by a factor of 1.5 or 6, depending on the source. Values are reported in Table 5.

Table 5 Integrated intensities from experiments and calculations for HCN electronic transitions. Lee45 and MacPherson & Simons88
Source T (K) J I (cm molecule−1)
2 A′ ← 1 A′
Computed
This work 0 0,1 5.56 × 10−17
This work 300 0,20 8.36 × 10−18
16ChRoAg59 0,1 4.17 × 10−17
3 A′ ← 1 A′
Computed
This work 0 0,1 1.57 × 10−14
This work 300 0,25 3.01 × 10−14
16ChRoAg59 0,1 1.40 × 10−14
Experiments
80Lee45 Room temperature 8.33 × 10−14
82NuGl89 Room temperature 1.86 × 10−13
78MaSixx88 Room temperature 4.41 × 10−14
4 A′ ← 1 A′
Computed
This work 0 0,1 2.18 × 10−14
This work 300 0,20 2.67 × 10−14
5 A′ ← 1 A′
Computed
This work 0 0,1 6.07 × 10−15
This work 300 0,20 8.71 × 10−15
Upper states
Experiments
80Lee45 Room temperature 2.50 × 10−12
82NuGl89 Room temperature 1.08 × 10−12
81NaKoOz90 Room temperature 8.20 × 10−13


The integrated intensity of the 2 A′ ← 1 A′ band for our model agrees with the value from 16ChRoAg. The intensity decreases with the increase of temperature and rotational states, from 5.56 × 10−17 cm molecules at 0 K to 8.36 × 10−18 at 300 K.

There is no experimental measurement for individual higher states; rather, they are measured collectively. The intensities are of the order of magnitude of 10−12 cm molecule−1. These values are higher than our computed values by a factor between 30 and 90, depending on the experimental source. The cause of the underestimation of the cross-sections and intensities with respect to the experiment was ascribed by 16ChRoAg59 to excited states possibly stealing intensity from dark states, not considered in their work or ours, through coupling to other states. It would seem more likely that these discrepancies arise from the computed PESs and TDMs: the electronic intensities are very sensitive to the position and shape of the upper state; coupled with a possible underestimation of the transition dipole moments.

The causes of the underestimation of the cross-sections and intensities with respect to the experiment was ascribed by 16ChRoAg for two possible reasons:59 these excited states might steal intensity from dark states, not considered in this work, through coupling from other states. Other possible reasons of these discrepancies are in the computed PESs and TDMs: the electronic intensities are very sensitive to the position and shape of the upper state; coupled with a possible underestimation of the transition dipole moments.

3.3 Temperature dependence of the 3 A′ ← 1 A′ band

As in the case of diatomic molecules, the increase in temperature leads to a change in the shape of the cross-sections. In order to understand this effect, the cross-sections were computed at 300, 1000 and 2000 K and are plotted in Fig. 10 in linear and logarithmic forms. The calculated partition function of 50[thin space (1/6-em)]681.8 at 2000 K is 99% of the value in the Harris line list for HCN91 from the ExoMol database. The increase of temperature corresponds to an increase of the background continuum, the well separated vibrational progression up to 2000 K. The maximum of the cross-section, corresponding to v2 = 7 peak, is red shifted by 1.79 nm (890 cm−1) from 300 K to 1000 K, and it remains constant up to 2000 K. The cross-section height decreases by a factor of 6 when the temperature reaches 2000 K from 300 K. These values are collected in Table 6. However, importantly for modelling applications, the threshold to photodissociation cross-section drops significantly with temperature, as observed previously in our study of diatomics,18 which can lead to very large increases in the effective rate of dissociation.
image file: d4cp02771j-f10.tif
Fig. 10 Photoabsorption spectra of the 3 A′ ← 1 A′ transition at 300, 1000, and 2000 K. The main plot is in a linear scale, while the inset is the same plot but in the logarithmic scale.
Table 6 Temperature dependence of various spectroscopic properties (σmax, λmax) of the 3 A′ ← 1 A′ transition
T (K) σ max (cm2 molecule−1) λ max (nm)
300 1.67 × 10−17 140.92
1000 8.22 × 10−18 142.71
2000 3.04 × 10−18 142.71


4 Conclusions

In this work, we explore the possibility of studying the direct photodissociation via dissociative potential surfaces and indirect photodissociation via quasi-bound excited states of triatomic molecules. The distinction between these two mechanisms is less clear cut in polyatomics than in diatomics. Our approach is based on solving the time-independent Schrödinger equation using variational nuclear motion calculations provided by adapting existing programs. We chose HCN as a test molecule, as it is a well-studied molecule experimentally and various computational studies. We performed the calculations using EVEREST, a program suite developed from DVR3D with some advantages when studying unbound states. This is because the latter code utilises a sinc-DVR basis, which provides better coverage of the dissociating coordinates compared to spherical oscillators, which tend to become increasingly sparse at large interatomic separations.

Our method is able to reproduce the experiments and the original work from 16ChRoAg.59 The discrepancies between experiments and our calculations can be explained by our use of an adiabatic description of the PESs and issues with the calculated transition dipole moments. Our cross-sections and intensities, like as in the previous study by 16ChRoAg, are underestimated with respect to the experiments. In their paper, 16ChRoAg mention as possible reasons for these discrepancies the lack of couplings in the model and the difficulty of computing accurate transition dipole moments involving electronically excited states.

The temperature dependence of the cross-section has been tested on the 3 A′ ← 1 A′ transitions at three different temperatures. A shift in the cross-section peak and a decrease of the cross-section height plus a lower of the photodissociation threshold are observed with increasing temperatures, as found for diatomic molecules.14,18

The next steps will be to produce a complete photodissociation model for HCN, performing electronic structure calculations for the states not included in this work and refining the available potential energy and transition dipole surfaces to recover the low-temperature experimental results. This model will then be used to provide cross-sections over an extended range of molecular temperatures. We plan to use our newly developed method to study other important molecules such as H2O and H2S. Non-adiabatic couplings between potential surfaces can provide an important mechanism for photodissociation. We note that EVEREST allows the implementation of different forms of surface couplings, including non-adiabatic effects. These features have been successfully tested on bound–bound electronic transitions of CaOH, which are characterised by both Renner–Teller and spin–orbit coupling effects, as demonstrated by Owens et al.77 This is a supplementary reason for our choice of EVEREST. Calculations on the photodissociation of H2S which include the explicit treatment of non-adiabatic couplings using EVEREST are currently in progress.

Data availability

The data for this article are contained in the article: numerical values for the figures are available from the authors. The documented codes, ExoSmooth and DVR3D, are available from https://github.com/exomol.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We thank professor Octavio Roncero for providing the potential energy surfaces and transition dipole moment surfaces. This work was funded by the ERC Advanced Investigator Project 883830 and the STFC Project ST/Y001508/1.

References

  1. V. H. J. Clark and S. N. Yurchenko, Phys. Chem. Chem. Phys., 2021, 23, 11990–12004 RSC.
  2. S.-M. Tsai, et al. , Nat. Astron., 2023, 617, 483–487 CAS.
  3. O. Venot, M. Rocchetto, S. Carl, A. R. Hashim and L. Decin, Astrophys. J., 2016, 830, 77 CrossRef.
  4. M. A. Badhan, E. T. Wolf, R. K. Kopparapu, G. Arney, E. M.-R. Kempton, D. Deming and S. D. Domagal-Goldman, Astrophys. J., 2019, 887, 34 CrossRef CAS.
  5. B. Fleury, M. S. Gudipati, B. L. Henderson and M. Swain, Astrophys. J., 2019, 871, 158 CrossRef CAS.
  6. N. K. Lewis, H. R. Wakeford, R. J. MacDonald, J. M. Goyal, D. K. Sing, J. Barstow, D. Powell, T. Kataria, I. Mishra, M. S. Marley, N. E. Batalha, J. I. Moses, P. Gao, T. J. Wilson, K. L. Chubb, T. Mikal-Evans, N. Nikolov, N. Pirzkal, J. J. Spake, K. B. Stevenson, J. Valenti and X. Zhang, Astrophys. J., 2020, 902, L19 CrossRef CAS.
  7. O. Venot, N. Fray, Y. Bénilan, M.-C. Gazeau, E. Hébrard, G. Larcher, M. Schwell, M. Dobrijevic and F. Selsis, Astron. Astrophys., 2013, 551, A131 CrossRef.
  8. R. Valiev, A. Berezhnoy, I. Gritsenko, B. Merzlikin, V. Cherepanov, T. Kurten and C. Wöhler, Astron. Astrophys., 2020, 633, A39 CrossRef CAS.
  9. A. N. Heays, A. D. Bosman and E. F. van Dishoeck, Astron. Astrophys., 2017, 602, A105 CrossRef.
  10. A. Noelle, A. C. Vandaele, J. Martin-Torres, C. Yuan, B. N. Rajasekhar, A. Fahr, G. K. Hartmann, D. Lary, Y.-P. Lee, P. Limao-Vieira, R. Locht, K. McNeill, J. J. Orlando, F. Salama and R. P. Wayne, J. Quant. Spectrosc. Radiat. Transfer., 2020, 253, 107056 CrossRef CAS PubMed.
  11. S. Y. Grebenshchikov, J. CO2 Util., 2016, 15, 32–40 CrossRef CAS.
  12. J. Tennyson, S. N. Yurchenko, A. F. Al-Refaie, E. J. Barton, K. L. Chubb, P. A. Coles, S. Diamantopoulou, M. N. Gorman, C. Hill, A. Z. Lam, L. Lodi, L. K. McKemmish, Y. Na, A. Owens, O. L. Polyansky, T. Rivlin, C. Sousa-Silva, D. S. Underwood, A. Yachmenev and E. Zak, J. Mol. Spectrosc., 2016, 327, 73–94 CrossRef CAS.
  13. J. Tennyson, S. N. Yurchenko, A. F. Al-Refaie, V. H. J. Clark, K. L. Chubb, E. K. Conway, A. Dewan, M. N. Gorman, C. Hill, A. E. Lynas-Gray, T. Mellor, L. K. McKemmish, A. Owens, O. L. Polyansky, M. Semenov, W. Somogyi, G. Tinetti, A. Upadhyay, I. Waldmann, Y. Wang, S. Wright and O. P. Yurchenko, J. Quant. Spectrosc. Radiat. Transfer, 2020, 255, 107228 CrossRef CAS.
  14. M. Pezzella, S. N. Yurchenko and J. Tennyson, Phys. Chem. Chem. Phys., 2021, 23, 16390–16400 RSC.
  15. S. N. Yurchenko, L. Lodi, J. Tennyson and A. V. Stolyarov, Comput. Phys. Commun., 2016, 202, 262–275 CrossRef CAS.
  16. I. I. Mizus, L. Lodi, J. Tennyson, N. F. Zobov and O. L. Polyansky, J. Mol. Spectrosc., 2022, 368, 111621 CrossRef.
  17. S. N. Yurchenko, A. F. Al-Refaie and J. Tennyson, Astron. Astrophys., 2018, 614, A131 CrossRef.
  18. M. Pezzella, S. N. Yurchenko and J. Tennyson, Mon. Not. R. Astron. Soc., 2022, 514, 4413–4425 CrossRef CAS.
  19. Z. Qin, T. Bai and L. Liu, Mon. Not. R. Astron. Soc., 2022, 516, 550–561 CrossRef CAS.
  20. L. E. Snyder and D. Buhl, Astrophys. J., 1971, 163, L47 CrossRef CAS.
  21. T. Hirota, S. Yamamoto, H. Mikami and M. Ohishi, Astrophys. J., 1998, 503, 717–728 CrossRef CAS.
  22. J. T. van Loon, J. R. Marshall, M. Cohen, M. Matsuura, P. R. Wood, I. Yamamura and A. A. Zijlstra, Astron. Astrophys., 2006, 447, 971–989 CrossRef CAS.
  23. J. S. Carr and J. R. Najita, Astrophys. J., 2011, 733, 102 CrossRef.
  24. S. Bruderer, D. Harsono and E. F. van Dishoeck, Astron. Astrophys., 2015, 575, A94 CrossRef.
  25. T. Hirota, S. Yamamoto, K. Kawaguchi, A. Sakamoto and N. Ukita, Astrophys. J., 1999, 520, 895–900 CrossRef CAS.
  26. M. Lippi, G. L. Villanueva, M. A. DiSanti, H. Boehnhardt, M. J. Mumma, B. P. Bonev and D. Prialnik, Astron. Astrophys., 2013, 551, A51 CrossRef.
  27. E. S. Wirström, M. S. Lerner, P. Källström, A. Levinsson, A. Olivefors and E. Tegehall, Astron. Astrophys., 2016, 588, A72 CrossRef.
  28. E. Lellouch, P. N. Romani and J. Rosenqvist, Icarus, 1994, 108, 112–136 CrossRef CAS.
  29. E. Lellouch, B. Bezard, D. F. Strobel, G. L. Bjoraker, F. M. Flasar and P. N. Romani, Icarus, 2006, 184, 478–497 CrossRef.
  30. E. Lellouch, M. Gurwell, B. Butler, T. Fouchet, P. Lavvas, D. F. Strobel, B. Sicardy, A. Moullet, R. Moreno, D. Bockelee-Morvan, N. Biver, L. Young, D. Lis, J. Stansberry, A. Stern, H. Weaver, E. Young, X. Zhu and J. Boissier, Icarus, 2017, 286, 289–307 CrossRef CAS.
  31. A. Tsiaras, M. Rocchetto, I. P. Waldmann, G. Tinetti, R. Varley, G. Morello, E. J. Barton, S. N. Yurchenko and J. Tennyson, Astrophys. J., 2016, 820, 99 CrossRef.
  32. E. M. Molter, C. A. Nixon, M. A. Cordiner, J. Serigano, P. G. J. Irwin, N. A. Teanby, S. B. Charnley and J. E. Lindberg, Astrophys. J., 2016, 152, 42 Search PubMed.
  33. L. Rezac, A. A. Kutepov, A. Faure, P. Hartogh and A. G. Feofilov, Astron. Astrophys., 2013, 555, A122 CrossRef.
  34. D. Dubois, N. Carrasco, M. Petrucciani, L. Vettier, S. Tigrine and P. Pernot, Icarus, 2019, 317, 182–196 CrossRef CAS.
  35. S. E. van Terwisga, E. F. van Dishoeck, P. Cazzoletti, S. Facchini, L. Trapman, J. P. Williams, C. F. Manara, A. Miotello, N. van der Marel, M. Ansdell, M. R. Hogerheijde, M. Tazzari and L. Testi, Astron. Astrophys., 2019, 623, A150 CrossRef CAS.
  36. F. Long, A. D. Bosman, P. Cazzoletti, E. F. van Dishoeck, K. I. Oberg, S. Facchini, M. Tazzari, V. V. Guzman and L. Testi, Astron. Astrophys., 2021, 647, A118 CrossRef CAS.
  37. M. Jin, J.-E. Lee and K.-T. Kim, Astrophys. J., Suppl. Ser., 2015, 219, 2 CrossRef.
  38. J. Bublitz, J. H. Kastner, M. Santander-García, V. Bujarrabal, J. Alcolea and R. Montez, Astron. Astrophys., 2019, 625, A101 CrossRef CAS.
  39. J. Bublitz, J. H. Kastner, P. Hily-Blant, T. Forveille, M. Santander-García, J. Alcolea and V. Bujarrabal, Astron. Astrophys., 2022, 659, A197 CrossRef CAS.
  40. D. D. Davis and H. Okabe, J. Chem. Phys., 1968, 49, 5526–5531 CrossRef CAS.
  41. A. Mele and H. Okabe, J. Chem. Phys., 1968, 51, 4798–4808 CrossRef.
  42. A. Meenakshi and K. K. Innes, J. Chem. Phys., 1986, 84, 6550–6554 CrossRef CAS.
  43. J. Guo, R. Eng, T. Carrington and S. V. Filseth, J. Chem. Phys., 2000, 112, 8904–8909 CrossRef CAS.
  44. S. Tatematsu and K. Kuchitsu, Bull. Chem. Soc. Jpn., 1977, 50, 2896–2898 CrossRef CAS.
  45. L. C. Lee, J. Chem. Phys., 1980, 72, 6414–6421 CrossRef CAS.
  46. C. R. Bucher and K. K. Lehmann, Chem. Phys. Lett., 1998, 294, 173–180 CrossRef CAS.
  47. U. G. Jørgensen, J. Almlöf, B. Gustafsson, M. Larsson and P. Siegbahn, J. Chem. Phys., 1985, 83, 3034–3042 CrossRef.
  48. T. van Mourik, G. J. Harris, O. L. Polyansky, J. Tennyson, A. G. Császár and P. J. Knowles, J. Chem. Phys., 2001, 115, 3706–3718 CrossRef CAS.
  49. J. M. Bowman, S. Irle, K. Morokuma and A. Wodtke, J. Chem. Phys., 2001, 114, 7923–7934 CrossRef CAS.
  50. A. J. C. Varandas and S. P. J. Rodrigues, J. Phys. Chem. A, 2006, 110, 485–493 CrossRef CAS PubMed.
  51. R. Dawes, A. F. Wagner and D. L. Thompson, J. Phys. Chem. A, 2009, 113, 4709–4721 CrossRef CAS PubMed.
  52. V. Y. Makhnev, A. A. Kyuberis, O. L. Polyansky, I. I. Mizus, J. Tennyson and N. F. Zobov, J. Mol. Spectrosc., 2018, 353, 40–53 CrossRef CAS.
  53. M. Perić, H. Dohmann, S. Peyerimhoff and R. J. Buenker, ZPhys-e.D: At., Mol. Clusters, 1987, 5, 65–75 Search PubMed.
  54. D. Xu, D. Xie and H. Guo, Chem. Phys. Lett., 2001, 345, 517–524 CrossRef CAS.
  55. D. Xu, D. Xie and H. Guo, J. Chem. Phys., 2002, 116, 10626–10635 CrossRef CAS.
  56. D. Xu, H. Guo and D. Xie, J. Theor. Comput. Chem., 2003, 02, 639–648 CrossRef CAS.
  57. D. Xu, D. Xie and H. Guo, J. Phys. Chem. A, 2002, 106, 10174–10183 CrossRef CAS.
  58. M. K. Nayak, R. K. Chaudhuri and S. N. L. G. Krishnamachari, J. Chem. Phys., 2005, 122, 184323 CrossRef PubMed.
  59. A. Chenel, O. Roncero, A. Aguado, M. Agúndez and J. Cernicharo, J. Chem. Phys., 2016, 144, 144306 CrossRef PubMed.
  60. A. Aguado, O. Roncero, A. Zanchet, M. Agúndez and J. Cernicharo, Astrophys. J., 2017, 838, 33 CrossRef PubMed.
  61. M. S. Priyadarshini, S. M. Jo, S. Venturi, D. W. Schwenke, R. L. Jaffe and M. Panesi, J. Phys. Chem. A, 2022, 126, 8249–8265 CrossRef CAS PubMed.
  62. H.-D. Meyer, U. Manthe and L. Cederbaum, Chem. Phys. Lett., 1990, 165, 73–78 CrossRef CAS.
  63. M. H. Beck, A. Jäckle, G. A. Worth and H. D. Meyer, Phys. Rep., 2000, 324, 1–105 CrossRef CAS.
  64. G. A. Worth, H.-D. Meyer, H. Köppel, L. S. Cederbaum and I. Burghardt, Int. Rev. Phys. Chem., 2008, 27, 569–606 Search PubMed.
  65. U. Manthe, H. Meyer and L. S. Cederbaum, J. Chem. Phys., 1992, 97, 3199–3213 CrossRef CAS.
  66. U. Manthe, H. Meyer and L. S. Cederbaum, J. Chem. Phys., 1992, 97, 9062–9071 CrossRef CAS.
  67. U. Manthe and A. D. Hammerich, Chem. Phys. Lett., 1993, 211, 7–14 CrossRef CAS.
  68. A. D. Hammerich, U. Manthe, R. Kosloff, H. Meyer and L. S. Cederbaum, J. Chem. Phys., 1994, 101, 5623–5646 CrossRef.
  69. S. Han, C. Xie, X. Hu, D. R. Yarkony, H. Guo and D. Xie, J. Phys. Chem. Lett., 2023, 14, 10517–10530 CrossRef CAS PubMed.
  70. R. Schinke, Photodissociation Dynamics, Cambridge University Press, 1993 Search PubMed.
  71. G. B. Mitev, S. N. Yurchenko and J. Tennyson, J. Chem. Phys., 2024, 160, 144110 CrossRef CAS PubMed.
  72. J. Tennyson, M. Pezzella, J. Zhang and S. N. Yurchenko, RAS Tech. Instr., 2023, 2, 231–237 CrossRef.
  73. S. N. Yurchenko, W. Szajna, R. Hakalla, M. Semenov, A. Sokolov, J. Tennyson, Y. Pavlenko and M. R. Schmidt, Mon. Not. R. Astron. Soc., 2024, 527, 9736–9756 CrossRef.
  74. J. Tennyson, M. A. Kostin, P. Barletta, G. J. Harris, O. L. Polyansky, J. Ramanlal and N. F. Zobov, Comput. Phys. Commun., 2004, 163, 85–116 CrossRef.
  75. A. O. Mitrushchenkov, J. Chem. Phys., 2012, 136, 024108 CrossRef PubMed.
  76. E. J. Zak and J. Tennyson, J. Chem. Phys., 2017, 147, 094305 CrossRef PubMed.
  77. A. Owens, A. Mitrushchenkov, S. N. Yurchenko and J. Tennyson, Mon. Not. R. Astron. Soc., 2022, 516, 3995–4002 CrossRef CAS.
  78. D. T. Colbert and W. H. Miller, J. Chem. Phys., 1992, 96, 1982–1991 CrossRef CAS.
  79. G. C. Groenenboom and D. T. Colbert, J. Chem. Phys., 1993, 99, 9681–9696 CrossRef CAS.
  80. J. Tennyson and B. T. Sutcliffe, J. Chem. Phys., 1982, 77, 4061–4072 CrossRef CAS.
  81. J. Tennyson and B. T. Sutcliffe, J. Mol. Spectrosc., 1983, 101, 71–82 CrossRef CAS.
  82. G. B. Mitev, R. P. Brady, O. Smola, J. Tennyson and S. N. Yurchenko, Phys. Chem. Chem. Phys., 2024 Search PubMed In preparation.
  83. J. Tennyson and B. T. Sutcliffe, Mol. Phys., 1986, 58, 1067–1085 CrossRef CAS.
  84. J. Tennyson, C. Hill and S. N. Yurchenko, 6th international conference on atomic and molecular data and their applications ICAMDATA-2012, 2013, pp. 186–195.
  85. A. U. Hazi and H. S. Taylor, Phys. Rev. A: At., Mol., Opt. Phys., 1970, 1, 1109–1120 CrossRef.
  86. Z. Bacic and J. Simons, J. Phys. Chem., 1982, 86, 1192–1200 CrossRef CAS.
  87. V. A. Mandelshtam, T. R. Ravuri and H. S. Taylor, Phys. Rev. Lett., 1993, 70, 1932–1935 CrossRef CAS PubMed.
  88. M. T. Macpherson and J. P. Simons, J. Chem. Soc., Faraday Trans. 2, 1978, 74, 1965–1977 RSC.
  89. J. A. Nuth and S. Glicker, J. Quant. Spectrosc. Radiat. Transfer, 1982, 28, 223–231 CrossRef CAS.
  90. T. Nagata, T. Kondow, Y. Ozaki and K. Kuchitsu, Chem. Phys., 1981, 57, 45–53 CrossRef CAS.
  91. R. J. Barber, J. K. Strange, C. Hill, O. L. Polyansky, G. C. Mellau, S. N. Yurchenko and J. Tennyson, Mon. Not. R. Astron. Soc., 2014, 437, 1828–1835 CrossRef CAS.
  92. H. Keller-Rudek, G. K. Moortgat, R. Sander and R. Sörensen, Earth Syst. Sci. Data, 2013, 5, 365–373 CrossRef.

Footnotes

Present address: Dipartimento di Fisica e Geologia, Universitá di Perugia, 06123 Perugia, Italy.
https://www.uv-vis-spectral-atlas-mainz.org/uvvis

This journal is © the Owner Societies 2024
Click here to see how this site uses Cookies. View our privacy policy here.