Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2022 Nov 2;18(12):7603–7619. doi: 10.1021/acs.jctc.2c00773

Perturb-Then-Diagonalize Vibrational Engine Exploiting Curvilinear Internal Coordinates

Marco Mendolicchio †,*, Julien Bloino , Vincenzo Barone ‡,*
PMCID: PMC9753597  PMID: 36322968

Abstract

graphic file with name ct2c00773_0008.jpg

The present paper is devoted to the implementation and validation of a second-order perturbative approach to anharmonic vibrations, followed by variational treatment of strong couplings (GVPT2) based on curvilinear internal coordinates. The main difference with respect to the customary Cartesian-based formulation is that the kinetic energy operator is no longer diagonal, and has to be expanded as well, leading to additional terms which have to be taken into proper account. It is, however, possible to recast all the equations as well-defined generalizations of the corresponding Cartesian-based counterparts, thus achieving a remarkable simplification of the new implementation. Particular attention is paid to the treatment of Fermi resonances with significant number of test cases analyzed fully, validating the new implementation. The results obtained in this work confirm that curvilinear coordinates strongly reduce the strength of inter-mode couplings compared to their Cartesian counterparts. This increases the reliability of low-order perturbative treatments for semi-rigid molecules and paves the way toward the reliable representation of more flexible molecules where small- and large-amplitude motions can be safely decoupled and treated at different levels of theory.

1. Introduction

Thanks to significant developments in both software and hardware in the last few decades, computational spectroscopy has become an invaluable tool for both experimentally and theoretically oriented research works.1,2 In the specific case of vibrational and ro-vibrational spectroscopy, the basic rigid rotor/harmonic oscillator (RRHO) model is implemented in all major quantum chemistry programs. However, more sophisticate models are needed in several circumstances [e.g., high-resolution spectroscopy, large-amplitude motions (LAM), and so forth], which should possibly couple accuracy and feasibility for medium- to large-size systems.3

Among the different approaches available for going beyond the RRHO approximation,428 those based on low-order perturbation theory applied to the Watson Hamiltonian (i.e., a fourth-order polynomial expansion of the potential energy expressed in Cartesian normal modes) are particularly appealing for their remarkable cost/performance balance, at least for semi-rigid molecular systems.2940

Moreover, a very general and robust model (referred to as GVPT232) can be built, which involves the diagonalization of relatively small Hamiltonians coupling a reduced number of strongly interacting states and including the second-order perturbative contributions of all the other ones.41,42 A number of other models have been introduced (e.g., the so-called VPT2 + F43 and VPT2 + K44 methods), which can be seen as particular cases of the GVPT2 approach and allow, in principle, the inclusion of any type of coupling, irrespective of resonance conditions. Although the conventional implementations of these models employ different equations for spherical, linear, symmetric, and asymmetric tops,45 it has been recently shown that the canonical representation used for the development of VPT2 equations of asymmetric tops can be extended to linear and symmetric tops, provided that a series of a posteriori transformations are performed.46

Further improvements can be obtained resorting to higher-order (usually sextic) anharmonic force fields coupled with variational [e.g., vibrational configuration interaction (VCI)6,8,14] or more accurate (e.g., VPT447) perturbative developments. Unfortunately, this kind of approaches converges slowly, and their cost becomes rapidly prohibitive as the dimension of the molecular system increases. An alternative route is based on reduced-dimensionality Hamiltonians tailored to describe a limited number of LAMs. Approaches belonging to this category are the internal coordinate path Hamiltonian (ICPH)48 and the reaction path Hamiltonian (RPH),4952 aimed at describing single LAMs or the reaction surface Hamiltonian (RSH)53 and reaction volume Hamiltonian (RVH),54 for the case of two or three LAMs, respectively.

Whenever the couplings between small-amplitude motions (SAMs) and LAMs are small, the SAMs (and, possibly, the SAMs–LAMs couplings) can be treated by the GVPT2 model, whereas the sub-problem of LAMs can be solved, for instance, by the so-called discrete variable representation (DVR), which is a quasi-variational, numerical method, introduced by Light and co-workers55 and later re-derived by Colbert and Miller.56 Unfortunately, normal modes based on Cartesian coordinates often give rise to non-negligible couplings, whereas internal (curvilinear) coordinates can strongly reduce the couplings between different classes of vibrations.57 One major drawback of internal coordinates is that their definition is not unique, and their construction can be quite involved, especially when targeting medium-to-large systems. This problem is solved by the redundant set of internal coordinates composed of all the bond lengths, valence, and dihedral angles, which is uniquely defined by the molecular topology.58,59 Thus, the route is paved toward the development of a general and robust GVPT2 platform employing curvilinear coordinates.

The basic equations of VPT2 in curvilinear coordinates have been derived by Quade60 and reworked more recently by Isaacson.57 The main difference between rectilinear (Cartesian) and curvilinear (internal) coordinates stems from the expansion of the kinetic energy operator, which introduces additional, possibly resonant, terms. However, full re-derivation of the equations allowed us to recast them in terms of quite straightforward generalizations of those based on Cartesian coordinates, so that it has been possible to extend the already available GVPT2 engine to curvilinear coordinates. Of course, kinetic energy contributions must be computed, but this does not involve additional quantum chemical computations, so that GVPT2 remains extremely effective in this context as well. Since the new formulation incorporates the recent generalization of the asymmetric-top equations to non-Abelian groups,46 all kinds of molecules can be treated with the same formalism.

This paper is organized as follows. We start with a discussion of the main features of the new GVPT2 engine, emphasizing the differences and similarities with the well-known equations for Cartesian coordinates. A robust strategy for the identification and treatment of Fermi resonances is also presented, followed by some technical aspects of the general implementation. After sketching the essential computational details, a number of test cases are analyzed to validate the new engine for semi-rigid molecules and to define the most suitable routes for coupling accuracy with effectiveness. As expected, inter-mode couplings are significantly smaller for curvilinear internal coordinates than for their Cartesian counterparts, paving the way toward achieving effective models enforcing the separation between SAMs and LAMs. The main conclusions and most promising perspectives are given in the last section.

2. Theory

2.1. Framework

The simplest set of internal coordinates is represented by the so-called primitive internal coordinates (PICs), which are composed of all bond lengths, valences, and dihedral angles and are uniquely defined by the molecular topology.61,62 While this set is generally redundant, this does not cause any problem (in analogy with translation and rotations when employing Cartesian coordinates) since all eigenvectors with vanishing eigenvalues can be removed after the harmonic problem is solved (vide infra). Next, PICs can be expressed in terms of their Cartesian counterparts by means of a Taylor series expansion

2.1. 1

where Na is the number of atoms, s is the vector containing the internal coordinates, whose values at the equilibrium geometry are collected in the vector seq, and x contains the atomic Cartesian coordinates. When the interest is focused on relatively low-vibrational excitations (i.e., close to the bottom of the potential energy surface (PES) well), eq 1 can be safely truncated at the second order and rewritten in a more compact form

2.1. 2

The elements of the so-called Wilson B matrix63 and its first derivative, B′, are

2.1. 3

and have well-known analytical expressions.61 By construction, only the indices j and k of the B′ tensor commute, whereas the B matrix is not symmetric or necessarily square, since the number of internal coordinates can be different from that of Cartesian coordinates.

2.2. Vibrational Hamiltonian in Curvilinear Coordinates

The starting point of the derivation is the definition of the expression of the kinetic-energy operator T in terms of the so-called Wilson G matrix

2.2. 4

where M is the diagonal matrix of the nuclear masses, while B is defined in eq 3. As a result, the vibrational kinetic energy Inline graphic is given by6466

2.2. 5

where Inline graphic. A more convenient form of eq 5 has been proposed by Podolosky,64 further discussed by Lauvergnat,66 and re-derived in this work (see Section S1 of the Supporting Information), leading to the following expression

2.2. 6

where the last term corresponds to a purely quantum-mechanical contribution to the kinetic energy, usually referred to as the extra-potential term.67Equation 6 represents the kinetic energy operator in terms of curvilinear coordinates s. However, application of perturbation theory to solve the vibrational problem requires a set of suitable reference wave functions too. In analogy with the treatment based on Cartesian coordinates, the harmonic oscillator model is employed to this end, by means of the so-called Wilson GF method (see Section S2 of the Supporting Information),68

2.2. 7

where F is the Hessian matrix of the potential energy with respect to the internal coordinates, L is the matrix containing the normal coordinates, and Λ is the diagonal matrix of squared harmonic frequencies (ω).

One of the advantages of a polynomial expansion in the normal-mode basis is that it leads to analytic integrals for both coordinate and momentum operators, together with a particularly simple second-quantization formulation, with these features strongly simplifying the identification of non-vanishing contributions in the perturbative expansion.

Equation 6 can be rewritten in terms of the dimensionless normal coordinates q and their conjugate momenta p

2.2. 8

where g, Inline graphic, and Inline graphic are the G matrix, its determinant, and the extra-potential term expressed in wavenumbers, respectively (see Section S3 of the Supporting Information)

2.2. 9

The potential energy (expressed in terms of dimensionless coordinates q) must be added to the kinetic energy in order to complete the vibrational Hamiltonian Inline graphic. Since the extra-potential term is well approximated by its value at the equilibrium configuration,57,69 it does not play any role in the calculation of transition energies. As a consequence, it will be neglected from now on, leading to the following expression of the vibrational Hamiltonian

2.2. 10

2.3. Perturbative Expansion of the Vibrational Hamiltonian

The perturbative treatment of Inline graphic is carried out by expanding both the kinetic and potential energies up to the second order. From here on, the symbol Inline graphic will be used to indicate the first term of eq 10, so that

2.3. 11

The g matrix entering the kinetic energy expression can be expanded to the second order

2.3. 12

where gijeq = ωiδij at the equilibrium configuration (see Section S3 of the Supporting Information) and δij is the Kronecker delta.

By inserting eq 12 in the definition of Inline graphic and introducing the following notation

2.3. 13

the kinetic energy can be written as a perturbative series

2.3. 14

where

2.3. 15
2.3. 16
2.3. 17

We recall that only the first and second indices of Inline graphic commute, while i only commutes with j and k only commutes with l in Inline graphic.

The expansion of the potential energy is analogous to its Cartesian counterpart

2.3. 18

where Inline graphic is the harmonic potential, while Inline graphic and Inline graphic contain, respectively, the cubic- and quartic-order contributions to the PES

2.3. 19
2.3. 20
2.3. 21

with

2.3. 22

The only difference with respect to Cartesian coordinates is the absence of the Coriolis term and the form of normal modes, which are now expressed in terms of internal (curvilinear) coordinates.

The full vibrational Hamiltonian can be written as follows

2.3. 23

where

2.3. 24
2.3. 25
2.3. 26

The curvilinear coordinate version of VPT2 requires not only the cubic and quartic force constants but also the first and second derivatives of the g (or G) matrix, whose calculation will be discussed in Section 3.

2.4. Vibrational Energies

In analogy with the treatment based on Cartesian coordinates, the anharmonic energies can be obtained through either canonical van Vleck (CV) or Rayleigh–Schrödinger (RS) perturbation theory (PT), which lead to the same final expressions. As already mentioned, the main difference with respect to the Cartesian-based framework is the presence, together with potential energy contributions, of additional terms arising from the kinetic energy. In order to clarify this point, let us recall the expression of the energy of the Rth vibrational state expanded up to the second order

2.4. 27

The form of the harmonic Hamiltonian Inline graphic is equivalent in Cartesian- and curvilinear-based formulations, so that both eigenvalues and eigenvectors are given by the customary expressions, and the first-order correction to the energy (eq 25) always vanishes

2.4. 28

since both Inline graphic and Inline graphic are odd operators in terms of normal coordinates and their conjugate momenta.

Finally, the second-order correction to the energy, εR(2), is given by

2.4. 29

Inspection of eq 29 shows that the anharmonic correction to each energy level is composed of three contributions, namely, potential (first line), kinetic (second line), and a cross term (third line). In the Cartesian version, only the potential term (albeit including Coriolis contributions) is present, so that the development becomes more complex when employing curvilinear internal coordinates. In order to accelerate the development stage, as well as reduce the possibility of errors, the derivation of εR(2) has been carried out by a multi-step procedure, with the help of ad hoc programs employing the second-quantization formalism followed by a manual post-processing.

The final expression of εR can be recast in the customary form

2.4. 30

where vR represents the vector of vibrational quantum numbers for the R-th state and ε0 is the zero-point vibrational energy (ZPVE), which will not be considered in the following because it does not affect energy differences between vibrational states. The χ matrix is given by the sum of three distinct contributions

2.4. 31

where the superscripts Inline graphic, Inline graphic, and × indicate the potential, kinetic, and a cross term, respectively. The form of the potential contribution is the same for Cartesian and curvilinear coordinates (see Section S4 of the Supporting Information), except for the presence of Coriolis contributions in the former case

2.4. 32
2.4. 33

while the purely kinetic contribution is

2.4. 34
2.4. 35

Finally, the cross term is

2.4. 36
2.4. 37

While the above expressions (fully equivalent to those reported in refs (57) and (60)) permit us to obtain the transition energies, a further algebraic manipulation leads to a more convenient form. By applying the partial fraction decomposition to eq 32 through eq 37 (see Section S5 of the Supporting Information for more details) and introducing the tensors ηijkl, σijk, and ρijk, we get

2.4. 38
2.4. 39
2.4. 40

Equation 31 can be rewritten as

2.4. 41
2.4. 42

Comparison of eqs 41 and 42 with their Cartesian counterparts (see eqs S33 and S34 of the Supporting Information) shows that the general form of the χ matrix does not change. More specifically, by reintroducing the Coriolis contribution and setting the derivatives of the g matrix to zero (ηijkl = fijkl and σijk = ρijk = fijk) in eqs 41 and 42, S36 and S37 are recovered. A similar procedure can be carried out to perform the inverse transformation. Note that, while the Coriolis term is absent in the internal-based VPT2 Hamiltonian, the perturbative development of the kinetic energy operator yields contributions formally equivalent to it. Therefore, the internal-based expression can be interpreted as a generalization of the Cartesian-based one. This formal equivalence presents two main advantages:

  • Implementation of eqs 41 and 42 into an existing code based on the Cartesian-based formulation is quite straightforward;

  • Analysis of Fermi resonances, which is the object of the next section, can be directly extended to curvilinear coordinates.

2.5. Fermi Resonances

Equations 41 and 42 show that the calculation of energy levels at the VPT2 level is plagued by Fermi resonances, irrespective of the use of rectilinear or curvilinear coordinates.38 Furthermore, the form of the perturbed vibrational Hamiltonian Inline graphic (eq 23) does not affect the nature of the contact transformation. As a consequence, the definition of the interaction terms of the contact-transformed Hamiltonian between interacting states can be directly generalized to the use of curvilinear coordinates. This premise is of fundamental importance for the analysis of Fermi resonances, the redefinition of suitable tests for their detection, and the variational correction at the basis of the GVPT2 approach.

2.5.1. Internal-Based Contact-Transformed Vibrational Hamiltonian

The off-diagonal elements of the contact-transformed Hamiltonian Inline graphic between two interacting states Inline graphic and Inline graphic can be written in terms of different orders of the original Hamiltonian Inline graphic.44

2.5.1. 43

In analogy with the expressions for the energy levels, the interaction element (eq 43) can also be partitioned into three contributions, which arise from the insertion of eq 23 in the above expression

2.5.1. 44
2.5.1. 45

where the term Inline graphic

2.5.1. 46

has been introduced for the sake of readability. Let us recall that one of the advantages of separating the contributions of different terms relies on the fact that the potential term is formally equivalent, except for the Coriolis couplings, to the expression derived in the Cartesian-based formulation. Equation 44 has been used to derive the interaction terms corresponding to Fermi resonances, which are discussed in the next section.

2.5.2. Diagnostic of Fermi Resonances: Extension of the Martin Test

The close correspondence between the χ matrix for different sets of coordinates allows a straightforward extension to curvilinear coordinates of the so-called Martin test for the identification of Fermi resonances.70 By switching back to Dirac’s notation, the matrix elements coupling the states |vR + 1k⟩ with |vR + 2i⟩ or |vR + 1i + 1j⟩ are

2.5.2. 47
2.5.2. 48

which are obtained from the corresponding Cartesian-based expressions replacing fiik and fijk by ρiik and ρijk, respectively. As a consequence, for all kinds of coordinates, the identification of Fermi resonances can be carried out by the same two-step procedure relying on the thresholds Δω1–2 and K1–2 with default values of 200 and 1 cm –1, respectively.71

Once the set of Fermi resonances has been identified, the corresponding terms in eqs 41 and 42 are discarded, and the resulting χ matrix is used for the calculation of the energy levels within the so-called deperturbed (DVPT2) scheme. The interaction terms corresponding to Fermi resonances can be treated in a successive variational step (leading to a model broadly referred to as GVPT2) by diagonalizing small matrices, whose diagonal elements are the deperturbed energies, while off-diagonal elements can be obtained from eqs 47 and 48 (Table 1).

Table 1. Formulation of the Test for the Identification of Fermi Resonances in Both Cartesian- and Internal-Based Formulations of VPT2.
  type I type II
step 1a |2ωi – ωk| ≤ Δω1–2 i + ωj – ωk| ≤ Δω1–2
step 2b ρiik4/256|2ωi – ωk|3K1–2 ρiik4/64|ωi + ωj – ωk|3K1–2
a

Step 1 is the same regardless of the formulation of VPT2.

b

ρijk = fijk in the Cartesian-based VPT2 framework.

3. Implementation

The implementation of the new engine can be split into three main steps. In the first one, the set of internal coordinates is defined starting from the reference geometry and used to build the B matrix and the B′ tensor, with the former being also used to calculate the G matrix. To this end, we have implemented a new code for the analytical computation of B, B′, G, and G′ matrices for bond lengths, valences (linear and non-linear), and dihedral angles. It is also possible to use different curvilinear coordinates by reading the B and B′ matrices generated by other programs. In both cases, the first derivative gij,k can be simply computed from Gij,k (see Section S3 of the Supporting Information), with the latter being given by

3. 49

where G′ is the tensor collecting the first Cartesian derivatives of the Wilson G matrix and can be further expanded by introducing the B′ tensor,

3. 50

which, in matrix form, becomes

3. 51

The terms gij,kl are obtained from their mass-weighted counterparts Gij,kl using an expression similar to eq 49

3. 52

In the above expression, G collects the second-order Cartesian derivatives of the G matrix,

3. 53

in matrix form, it becomes

3. 54

where B is the second-order Cartesian derivative of the B matrix. Consequently, the analytical calculation of the terms Gij,kl relies on the four-dimensional tensor B, which presents some difficulties. In the first place, it is composed of all third-order derivatives of the internal coordinates with respect to Cartesian coordinates, whose derivation and implementation involve a significant effort. Furthermore, the use of the four-dimensional tensor B with one dimension equal to N and the other three equal to 3Na may imply additional concerns in terms of both computer time and memory storage. For these reasons, a more viable strategy is the analytical computation of the first derivatives Gij,k, followed by their use in the finite-difference calculation of second derivatives.

The second step is the definition of the displacements along the curvilinear normal modes, the computation of Hessians at these geometries, and the assembly of potential and kinetic contributions to cubic and quartic force constants. This task is performed by a script, which calls an external quantum chemical package to compute the gradients and Hessians in Cartesian coordinates at suitable geometries. An external implementation has the advantage that the most computer-intensive (but embarassingly parallel) step can be performed in the most effective way, namely, distributed among different computing nodes. The calculation of the Hessian matrix F in internal coordinates can be carried out using the following expression72

3. 55

where the internal-based gradient gs can be easily obtained starting from its Cartesian counterpart gx as

3. 56

Furthermore, the overall contribution due to translations and rotations can be factored out by replacing (HxgsB′) and gx by P(HxgsB′)P and Pgx, respectively, where P = BB represents the projection matrix.

Second derivatives of the G matrix are also obtained from finite-difference expressions

3. 57
3. 58

Equation 57 includes, of course, the terms Gii,kk and Gii,ii, while eq 58 includes Gij,ij, Gii,kl and Gij,kj.

As mentioned above, these computations have been always performed by a new script preparing the input stream and submitting harmonic computations for the different geometries needed in the finite-difference evaluation. Although different electronic structure codes could be employed in this step, all the computations reported in this work have been performed by the G16 package.73 Atomic units are used systematically together with angles in radians. On the basis of previous experience and several new numerical tests, a default step (δQ) of 0.02 amu1/2 Bohr has been chosen for all kinds of coordinates.

The third step involves the implementation of the GVPT2 equations for curvilinear coordinates discussed in Section 2. This has been accomplished by extending the general platform for Cartesian coordinates already available in the Gaussian code.

A flowchart describing the whole workflow is sketched in Figure 1.

Figure 1.

Figure 1

Flowchart describing the new workflow for the anharmonic calculations in curvilinear coordinates, where the tasks performed by a generic quantum chemical code, the Gaussian package, and the novel external program are highlighted. M represents the number of active modes.

4. Computational Details

In light of previous experience, hybrid density functionals B3PW9174 and PW6B9575 were used in conjunction with the jul-cc-pVDZ (hereafter julDZ) basis set,76 whereas double-hybrid functionals B2PLYP77,78 and revDSD-PBEP8679 together with second-order Møller–Plesset PT (MP2)80 were employed in conjunction with the jun-cc-pVTZ (hereafter junTZ) basis set.76 Furthermore, empirical dispersion contributions were systematically added in DFT computations by means of Grimme’s D3 model with Becke–Johnson damping.81,82 The above computational levels will be denoted in the following as B3, PW6, B2, rDSD, and MP2, respectively.

5. Results and Discussion

In this section, we will present a number of results obtained by the new VPT2 engine with the objective of validating its implementation and highlighting the advantages of curvilinear over Cartesian coordinates concerning both effectiveness and accuracy. After considering semi-rigid systems, where different sets of coordinates provide comparable results (but the number and strength of inter-mode couplings are very different), we will consider some prototypical flexible systems, where the advantages of curvilinear coordinates become more apparent. The structures of all the studied molecules are sketched in Figure 2.

Figure 2.

Figure 2

Structures of all the studied molecules. (a) Formaldehyde, (b) acetylene, (c) cyclopropane, (d) methane, (e) ethylene, (f) oxirane, (g) acetic acid, (h) uracil, and (i) Ip conformer of glycine.

5.1. Validation of VPT2 in Curvilinear Coordinates

The new VPT2 implementation has been validated for the asymmetric (formaldehyde), linear (acetylene), symmetric (cyclopropane), and spherical (methane) tops shown in Figure 2a–d. Comparison between VPT2 results in Cartesian and curvilinear coordinates permits us to test the correctness of both the VPT2 equations (also in the presence of Fermi resonances) and the elements of the G matrix and its derivatives. All the computations have been performed at the MP2/junTZ level, which couples semi-quantitative accuracy with the lack of any numerical noise, as it would be the case for DFT methods. Note that, in the absence of numerical errors, harmonic frequencies are identical for any set of coordinates.

The results collected in Tables 25 show that for small semi-rigid molecules Cartesian and curvilinear coordinates provide equivalent results, irrespective of the symmetry (Abelian or non-Abelian point group) of the system. Furthermore, in the case of formaldehyde, Coriolis couplings are not negligible, especially for the wagging and CH2 asymmetric stretching, and the curvilinear results are much closer to the Cartesian counterparts including Coriolis couplings than to those neglecting them (see Table 2). This shows that some terms in the development of kinetic energy in curvilinear coordinates are equivalent to Coriolis couplings in Cartesian coordinates.

Table 2. Comparison of the Cartesian and Curvilinear VPT2, DVPT2, and GVPT2 Wavenumbers (in cm–1) of Formaldehyde at the MP2/junTZ Level.

        Cartesian
curvilinear
  assignment symm. ω νVPT2a νDVPT2 νGVPT2 νVPT2 νDVPT2 νGVPT2
1 CH2 s str. A1 2975 2829 (2827) 2820 2829 2829 2829 2829
2 C=O str   1756 1724 (1723) 1724 1724 1724 1724 1724
3 HCH s bend   1545 1512 (1510) 1512 1512 1512 1512 1512
4 HCH op wag B1 1203 1183 (1169) 1183 1183 1183 1183 1183
5 CH2 a str. B2 3051 3029 (3017) 2897 2866 3029 2902 2869
6 HCH a bend   1271 1250 (1246) 1250 1250 1250 1250 1250
2 + 6 comb. band   2975 2835 2967 2999 2835 2962 2995
a

In parenthesis, the VPT2 frequencies obtained without including Coriolis couplings have been reported.

Table 5. Comparison of the Cartesian and Curvilinear VPT2 Fundamental Wavenumbers (in cm–1) of Methane at the MP2/junTZ Level.

      Cartesian curvilinear
assignment symmetry ω νVPT2 νVPT2
CH str. A1 3073 2953 2953
bend. E 1586 1549 1549
CH str. T2 3209 3074 3074
bend   1352 1318 1318

Table 3. Comparison of Cartesian and Curvilinear VPT2 Fundamental Wavenumbers (in cm–1) of Acetylene at the MP2/junTZ Level.

      Cartesian curvilinear
assignment symmetry ω νVPT2 νVPT2
CH s str. Σg 3525 3397 3397
CC str   1969 1931 1931
CH a str Σu 3437 3317 3317
HCC s bend Πg 592 609 609
HCC a bend Πu 748 739 739

Table 4. Comparison of the Cartesian and Curvilinear VPT2, DVPT2, and GVPT2 Fundamental Wavenumbers (in cm–1) of Cyclopropane at the MP2/junTZ Level.

      Cartesian
curvilinear
assignment symmetry ω νVPT2 νDVPT2 νGVPT2 νVPT2 νDVPT2 νGVPT2
CH2 s str. A1 3196 3075 3075 3075 3074 3074 3074
CH2 sciss.   1533 1504 1485 1515 1501 1484 1506
ring str.   1231 1203 1203 1203 1201 1201 1201
CH2 twist A1 1166 1131 1131 1131 1128 1128 1128
CH2 wagg. A2 1085 1057 1057 1057 1054 1054 1054
CH2 a str. A2 3298 3154 3154 3154 3154 3154 3154
CH2 rock.   869 863 863 863 857 857 857
CH2 s str. E′ 3187 3067 3067 3068 3067 3067 3067
CH2 def.   1485 1440 1444 1443 1436 1439 1440
CH2 wagg.   1050 1022 1022 1022 1017 1017 1017
ring def.   905 878 878 878 876 876 876
CH2 a str. E 3279 3135 3135 3135 3135 3135 3135
CH2 twist + rock   1220 1192 1192 1192 1190 1190 1190
twist + rock.   747 741 741 741 734 734 734

5.2. Reconciling Accuracy and Feasibility

For small semi-rigid molecules, the accuracy of state-of-the-art quantum-chemical methodologies can rival that of experimental techniques.8385 However, their extension to large (possibly flexible) systems faces a number of difficulties ranging from the very unfavorable scaling of such methods with the number of basis functions to the proper description of flat PESs.3,86 A viable route to obtain accurate results, even for relatively large molecular systems (a few dozens of atoms), is provided by dual-level models, which combine accurate calculations of molecular structures and harmonic force fields to cheaper yet reliable approaches for taking into account anharmonic contributions resulting from SAMs and, possibly, a small number of LAMs. The role of curvilinear coordinates in improving these aspects is analyzed in the next subsections.

5.3. Coupling Issue

The accuracy of low-level perturbative treatments is, of course, related to the number and strength of couplings between different modes and, especially, to the relative role played by two- and three-mode interactions. We will use ethylene (see Figure 2e) to analyze this aspect. As a matter of fact, GVPT2 results obtained employing Cartesian or curvilinear coordinates are virtually indistinguishable (as expected for semi-rigid molecules), but the number and strength of couplings determining the final result are different in the two cases. Furthermore, the terms neglected in VPT2 energies but actually computed in the numerical differentiation of analytical Hessians (i.e., three-mode quartic force constants) are significantly different in the two implementations. This is well evidenced in Figure 3, which shows that both the number and strength of all three-mode couplings are strongly reduced when using curvilinear internal coordinates.

Figure 3.

Figure 3

Comparison of the number of cubic (fijk (ijk)) and quartic (fijkk (ij)) force constants of ethylene above a given threshold (in cm–1) computed at the MP2/junTZ level with Cartesian or curvilinear coordinates.

Another example is offered by oxirane (see Figure 2f), whose computed vibrational energies are collected in Table 6. While more accurate results can be obtained increasing the computational level, all the experimental trends are correctly reproduced and, once again, the use of curvilinear coordinates strongly reduces the role of three-mode couplings (see Figure 4), which can thus be safely neglected with a few exceptions.

Table 6. Comparison of the Cartesian and Curvilinear VPT2, DVPT2, and GVPT2 Wavenumbers (in cm–1) of Oxirane at the MP2/junTZ Level with the Experimental Data.

        Cartesian
curvilinear
 
  assign. symm. ω νVPT2 νDVPT2 νGVPT2 νVPT2 νDVPT2 νGVPT2 exp.a
1 (CH2 s-str) A1 3160 3057 3030 3058 3057 3034 3057 3006
2 (CH2 scis)   1552 1503 1503 1503 1504 1504 1504 1498
3 (ring str)   1310 1279 1279 1279 1279 1279 1279 1271
4 (CH2 wag)   1155 1125 1125 1125 1125 1125 1125 1120
5 (ring def.)   902 880 880 880 880 880 880 877
6 (CH2 a-str) A2 3264 3119 3119 3119 3119 3119 3119 3065
7 (CH2 twist)   1175 1151 1151 1151 1151 1151 1151 1142
8 (CH2 rock)   828 815 815 815 816 816 816 822
9 (CH2 s-str) B1 3153 3045 3014 3050 3045 3021 3048 3006
10 (CH2 scis)   1519 1480 1480 1480 1480 1480 1480 1472
11 (CH2 wag)   1171 1143 1143 1143 1143 1143 1143 1151
12 (ring def)   851 822 822 822 822 822 822 892
13 (CH2 a-str) B2 3250 3106 3106 3106 3106 3106 3106 3063
14 (CH2 twist)   1186 1163 1163 1163 1164 1164 1164 1142
15 (CH2 rock)   1059 1032 1032 1032 1033 1033 1033 822
2 + 2 overtone A1 3103 2979 3005 2977 2980 3003 2979  
2 + 10 comb. band   3070 2951 2981 2945 2952 2976 2949  
a

Ref (87).

Figure 4.

Figure 4

Comparison of the number of cubic (fijk (ijk)) and quartic (fijkk (ij)) force constants of oxirane above a given threshold (in cm–1) computed at the MP2/junTZ level with Cartesian or curvilinear coordinates.

While only potential couplings involve an increased computational cost of the underlying electronic computations, a fully unbiased comparison between Cartesian and curvilinear implementations requires the evaluation of the role of kinetic couplings. Figure 5 shows that, as expected, three-mode kinetic contributions are essentially negligible.

Figure 5.

Figure 5

Number of three-mode first- and second-order g matrix derivatives of oxirane above a given threshold (in cm–1) computed at the MP2/junTZ level with curvilinear coordinates.

An even more vexing problem is related to the presence of LAMs like, for example, methyl rotations. The situation is illustrated in Table 7 for the specific example of acetic acid (see Figure 2g). Although VPT2 results are very close for different sets of coordinates, the Cartesian description shows comparable contributions from the one-dimensional anharmonicity of the CH3 rotation and its coupling with other modes. As a consequence, any separation between LAMs and SAMs faces against severe difficulties. For example, neglecting inter-mode couplings, the computed frequency of CH3 torsion becomes completely unrealistic when employing Cartesian coordinates (−5213 cm–1), whereas the value issuing from curvilinear coordinates (68 cm–1) remains reasonable.

Table 7. Comparison of the Cartesian and Curvilinear GVPT2 Fundamental Wavenumbers (in cm–1) of Acetic Acid at the MP2/junTZ Level with the Experimental Data.

assignment symmetry ω Cartesian curvilinear exp.a
OH str. A′ 3760 3575 3575 3583
CH3 a str.   3227 3083 3084 3051
CH3 s str.   3101 2992 2992 2944
C=O str.   1812 1782 1782 1788
CH3 a def.   1490 1450 1448 1430
CH3 s def.   1421 1380 1377 1383
OH bend   1342 1324 1322 1264
C–O str.   1206 1161 1159 1182
CH3 rock.   1007 988 986 989
CC str.   875 856 856 847
OCO bend   583 576 577 657
CCO bend   423 424 422 581
CH3 a str. A 3184 3044 3044 2996
CH3 a def.   1498 1440 1437 1430
CH3 rock.   1074 1049 1045 1048
C=O op bend   663 644 643 642
C–O torsion   552 538 537 534
CH3 torsion   75 85 85 93
      (−5213) (68)  
a

Ref (87).

5.4. Dual-Level Methods

It is well known that harmonic frequencies are more sensitive to the level of the underlying electronic Hamiltonian than higher-order force constants. The most important reason for this is the increased importance of the nuclear repulsion contribution for higher-order derivatives, with this term being always treated exactly.88 Furthermore, the computational cost of a full quartic force field is much higher than that of the harmonic part at the same level of theory. Finally, the whole foundation of any perturbative treatment is that the final results are more sensitive to the quality of the zero-order (harmonic) contribution than to that of the first- and second-order corrections. These considerations lead to the development of the so-called dual-level (or hybrid) methods, with the simplest one (referred to as additive approach, Add)89 solving the VPT2 equations employing the low-level harmonic frequencies and higher-order derivatives. Then, the results are corrected for the difference between high- and low-level harmonic frequencies. This approach is not recommended because the denominators of the perturbative contributions are evaluated by low-level harmonic frequencies, which can lead to non-negligible distortions of the results. A simple recipe for solving this problem is offered by the so-called substitution (Sub) approach32 in which the VPT2 equations are solved employing low-level anharmonic couplings, but high-level harmonic frequencies are used to compute the denominators.

The quality of the results obtainable by dual-level methods is analyzed in some detail for the case of uracil (see Figure 2h). Inspection of Table 8 confirms that among hybrid density functionals, the B3PW91/julDZ model represents the best compromise between accuracy and feasibility for molecules too large to be treated by state-of-the-art post-Hartree–Fock methods.71

Table 8. Comparison of Harmonic Frequencies and Curvilinear GVPT2 Fundamental Wavenumbers (in cm–1) of Uracil with Experimental Data.

    B3
rDSD//B3
best//B3
 
assignment symm. ωa νa ωb addc subd ωe addf subg exp.h
N1–H str A′ 3654 3485 3661 3492 3493 3653 3484 3483 3485
N3–H str   3612 3442 3612 3442 3442 3602 3432 3428 3435
C5–H str   3265 3125 3265 3125 3125 3253 3113 3103  
C6–H str   3219 3074 3223 3078 3089 3218 3073 3069  
C=O str   1815 1785 1807 1777 1775 1790 1760 1762 1764
C4=O str   1781 1767 1774 1760 1741 1762 1748 1728 1706
C5=C6 str   1688 1652 1684 1648 1650 1678 1642 1642 1646
N1–H bend   1510 1464 1512 1466 1468 1505 1459 1460 1472
C6–H bend   1420 1387 1429 1396 1393 1427 1394 1397 1400
N3–H bend   1404 1370 1418 1384 1386 1414 1380 1381 1389
C5–H bend   1382 1347 1395 1360 1360 1394 1359 1362 1359
ring str def   1236 1204 1243 1211 1211 1248 1216 1214 1217
ring str def   1203 1177 1212 1186 1185 1205 1179 1178 1185
ring str def   1091 1073 1093 1075 1074 1084 1066 1063 1075
ring str def   992 977 997 982 980 995 980 978 980
ring str def   973 951 975 953 931 968 946 954 958
ring str def   779 755 774 750 749 773 749 766 759
ring bend def   558 555 560 557 551 545 542 541 562
ring bend def   541 532 542 533 533 541 532 536 537
ring bend def   519 512 519 512 512 517 510 510 516
C=O bend   385 384 388 387 385 387 386 374 391
C6–H op bend A 970 925 979 934 963 973 928 954 987
C5–H op bend   815 798 822 805 806 814 797 796 804
C2=O op bend   766 751 767 752 752 765 750 750 757
C4=O op bend   728 714 735 721 721 728 714 713 718
N3–H op bend   695 657 683 645 645 670 632 630 662
N1–H op bend   578 538 556 516 514 559 519 517 551
ring op def   397 384 395 382 381 388 375 385 411
ring op def   168 159 163 154 154 159 150 150 185
ring op def   151 143 146 138 138 140 132 132  
MAE     13   12 11   13 11  
a

B3/julDZ.

b

rDSD/junTZ.

c

Hybrid model based on the additive approach employing harmonic frequencies at the rDSD/junTZ level in conjunction with anharmonic corrections at the B3/julDZ level.

d

Hybrid model based on the substitution approach employing harmonic frequencies at the rDSD/junTZ level in conjunction with anharmonic corrections at the B3/julDZ level.

e

Best estimate (ref (92)).

f

Hybrid model based on the additive approach employing best-estimate harmonic frequencies in conjunction with anharmonic corrections at the B3/julDZ level.

g

Hybrid model based on the substitution approach employing best-estimate harmonic frequencies in conjunction with anharmonic corrections at the B3/julDZ level.

h

Refs (9395).

The double hybrid rDSD functional in conjunction with a partially augmented triple-zeta basis set can be generally used to improve the harmonic part of the force field.90,91 In the case of uracil, B3, rDSD, and, even, coupled cluster harmonic frequencies are quite close, so that the dual-level approach does not improve the results in a significant way, but the situation is different in several other cases (e.g., glycine discussed below). From a more general perspective, the results show that the GVPT2 model is capable of providing remarkably accurate results for semi-rigid molecules plagued by significant resonances, as is the case for uracil.

As a second example, we consider the Ip conformer of glycine (see Figure 2i).96 The computed vibrational frequencies are compared in Table 9 with their experimental counterparts. The agreement is remarkable for all the tested computational models and, in particular, dual-level rDSD//B3 approaches reduce the average error getting closer to the results obtained at the much more expensive rDSD level. Although the results are similar for Cartesian and curvilinear coordinates, they show a very different pattern when one tries to disentangle the contribution of LAMs (ϕ and ψ torsions, modes 23 and 24 at about 200 and 90 cm–1). As a matter of fact, Figure 6 shows that both diagonal and off-diagonal potential anharmonic contributions are very large in Cartesian coordinates, whereas this is not the case when employing curvilinear coordinates. As a consequence, separation between SAMs and LAMs should be safer in the context of curvilinear coordinates.

Table 9. Comparison of the Cartesian and Curvilinear Harmonic and GVPT2 Fundamental Wavenumbers (in cm–1) of the Ip Conformer of Glycine with the Experimental Data.

    ω
Cartesian curvilinear
 
assignment symm. B3a rDSDb rDSDb B3a addc subd rDSDb exp.
OH str A′ 3767 3766 3581 3572 3570 3571 3579 3585e
NH2 s str   3514 3521 3377 3356 3373 3366 3370 3359f
CH2 s str   3057 3068 2953 2920 2949 2938 2947 2943f
C=O str   1825 1817 1786 1790 1775 1783 1788 1779g
NH2 bend   1668 1682 1627 1574 1630 1616 1603 1608f
CH2 bend   1439 1472 1435 1404 1470 1437 1436 1429f,g
CH2 bend   1400 1417 1387 1362 1396 1379 1407 1405f
(OH + CH2) bend   1301 1317 1295 1271 1301 1285 1299 1297f
CN str + OH bend   1177 1176 1134 1140 1147 1148 1135 1136f,g
C=O str + OH bend   1143 1137 1102 1100 1091 1097 1103 1101f,g
CC str + NH2 bend   927 937 888 868 893 883 892 883f,g
CC str   832 834 808 798 803 801 811 801f,g
(NH2 + OCO) bend   634 637 633 624 630 627 636 619f,g
CCO(H) bend   464 467 462 451 457 454 464 458h
CCN bend   255 259 255 239 248 244 261 250h
NH2 a str A 3590 3599 3428 3425 3423 3414 3428 3410f,g
CH2 a str   3100 3109 2965 2957 2981 2972 2965 2969
CH2 bend   1376 1397 1357 1333 1377 1356 1360 1340
CH2 NH2 twist   1174 1194 1164 1145 1176 1156 1167 1166f
CH2 NH2 twist   913 923 911 899 918 908 913 907f,g
OH op bend   653 649 619 602 594 598 623 615f
OH op bend   509 511 495 478 483 481 499 500g
CN tors (ϕ)   210 217 203 151 168 161 232 204h
CC tors (ψ)   67 68 64 20 21 21 90  
MAE       8 16 13 11 8  
a

julDZ basis set.

b

junTZ basis set.

c

Hybrid model based on the additive approach, employing harmonic frequencies at the rDSD/junTZ level in conjunction with anharmonic corrections at the B3/julDZ level.

d

Hybrid model based on the substitution approach, employing harmonic frequencies at the rDSD/junTZ level in conjunction with anharmonic corrections at the B3/julDZ level.

e

Ref (97).

f

Ref (98).

g

Ref (99).

h

Ref (100).

Figure 6.

Figure 6

Comparison of the Cartesian (top panel) and curvilinear (bottom panel) quartic force constants of the Ip conformer of glycine involving modes 23 and 24 at the rDSD/junTZ level of theory.

Deeper insights on the role of different couplings can be obtained by comparing the results of a series of computations in which one- two- and three-mode anharmonic contributions (both potential and kinetic) are progressively added to the starting harmonic model. As already mentioned, the computational effort of electronic structure computations increases sharply with the number of different modes taken into account at the same time for potential couplings. The results collected in Table 10 show that, when employing curvilinear internal coordinates, the HCAM (harmonic coupled anharmonic modes) model has already performed a remarkable job, and inclusion of two-mode anharmonic couplings provides semi-quantitative results. These findings pave the way toward the implementation of very effective reduced-dimensionality approaches, in which only a few key anharmonic contributions are taken into account.

Table 10. Comparison of the Cartesian and Curvilinear Harmonic and GVPT2 Fundamental Wavenumbers (in cm–1) of the Ip Conformer of Glycine at the rDSD/junTZ Level of Theory Starting from Diagonal Anharmonic Couplings and Then Adding Two- and Three-Mode Couplings in a Stepwise Mannera.

      Cartesian
curvilinear
assignment symm. ω diagonalb two-modec three-moded diagonalb two-modec three-moded
OH str A′ 3766 3604 3527 3581 3603 3577 3579
NH2 s str   3521 3453 3309 3377 3449 3368 3370
CH2 s str   3068 3015 2914 2953 3013 2947 2947
C=O str   1817 1806 1785 1786 1806 1788 1788
NH2 bend   1682 1681 1732 1627 1676 1621 1603
CH2 bend   1472 1471 1475 1435 1470 1440 1436
CH2 bend   1417 1421 1391 1387 1418 1386 1407
(OH + CH2) bend   1317 1322 1366 1295 1315 1307 1299
CN str + OH bend   1176 1177 1162 1134 1174 1151 1135
C=O str + OH bend   1137 1149 1123 1102 1142 1111 1103
CC str + NH2 bend   937 946 1017 888 929 902 892
CC str   834 844 874 808 836 822 811
(NH2 + OCO) bend   637 639 645 633 639 638 636
CCO(H) bend   467 468 471 462 467 470 464
CCN bend   259 264 284 255 261 273 261
NH2 a str A 3599 3683 3297 3428 3683 3423 3428
CH2 a str   3109 3181 2907 2965 3181 2958 2969
CH2 bend   1397 1401 1378 1357 1395 1377 1340
CH2 NH2 twist   1194 1201 1188 1164 1193 1175 1167
CH2 NH2 twist   923 937 961 911 926 920 913
OH op bend   649 838 712 619 633 626 623
OH op bend   511 607 616 495 511 509 499
CN tors (ϕ)   217 1302 755 203 209 238 232
CC tors (ψ)   68 839 579 64 76 94 90
MAE     138e 94e   45f 8f  
a

Kinetic and potential terms are added at the same time in the case of curvilinear coordinates.

b

Calculation performed by including only diagonal terms.

c

Calculation performed by including up to two-mode vibrational couplings.

d

Calculation performed by including up to three-mode couplings.

e

Mean absolute error computed with respect to the three-mode Cartesian-based fundamental wavenumbers.

f

Mean absolute error computed with respect to the three-mode internal-based fundamental wavenumbers.

6. Conclusions

In this work, we have shown how the VPT2 equations for Cartesian coordinates can be extended to curvilinear internal coordinates without any additional computational bottleneck.

The results for several test cases point out the generality and robustness of the new GVPT2 engine employing curvilinear coordinates, which allows the effective treatment of medium-to-large-sized molecules for all electronic structure methods for which analytical Hessians are available. Dual-level methods combining high-level harmonic terms with lower-level anharmonic contributions further widen the range of application of the general platform.

The new development offers a number of advantages with respect to previous, ad hoc procedures. The first aspect concerns the ease of implementation since the new approach does not require any heavy modification of the codes already supporting VPT2 for asymmetric tops and Cartesian coordinates. However, the most important advantage is that the intrinsic problems of a low-order perturbative treatment based on Cartesian normal modes are strongly reduced. As a matter of fact, as clearly stated by Stanton and co-workers in connection with higher-order perturbative treatments (e.g., VPT4),47 VPT based on a rectilinear Hamiltonian is simply poorly suited to the problem of floppy molecular systems, and approaches such as VPT2 in curvilinear coordinates are to be preferred. While work aimed at developing more refined models for the treatment of LAMs is underway in our laboratory, we think that already the present implementation offers a number of interesting perspectives for the study of molecular systems of current scientific and technological interest.

Acknowledgments

Financial fundings from the Italian Ministry of University and Research (grant 2017A4XRCA) and the Italian Space Agency (ASI; “Life in Space” project, N. 2019-3-U.0) are gratefully acknowledged. We also thank the technical staff at SNS’ SMART Laboratory for managing the computational facilities.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jctc.2c00773.

  • Additional details about the kinetic energy operator in curvilinear coordinates, unit conversion, Wilson GF method, and resonant terms in the χ matrix (PDF)

The authors declare no competing financial interest.

Supplementary Material

ct2c00773_si_001.pdf (241.5KB, pdf)

References

  1. Jensen P.; Bunker P. R.. Chapter The Born-Oppenheimer Approximation. Computational Molecular Spectroscopy; Jensen P., Bunker P. R., Eds.; John Willey & Sons, Inc., 2000. [Google Scholar]
  2. Barone V. The Virtual Multifrequency Spectrometer: a New Paradigm for Spectroscopy. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2016, 6, 86–110. 10.1002/wcms.1238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Puzzarini C.; Bloino J.; Tasinato N.; Barone V. Accuracy and Interpretability: The Devil and the Holy Grail. New Routes across Old Boundaries in Computational Spectroscopy. Chem. Rev. 2019, 119, 8131–8191. 10.1021/acs.chemrev.9b00007. [DOI] [PubMed] [Google Scholar]
  4. Nielsen H. H. The Vibration-Rotation Energies of Molecules. Rev. Mod. Phys. 1951, 23, 90–136. 10.1103/revmodphys.23.90. [DOI] [Google Scholar]
  5. Mills I. M.Vibration-Rotation Structure in Asymmetric- and Symmetric-Top Molecules. Molecular Spectroscopy: Modern Research; Rao K. N., Mathews C. W., Eds.; Academic Press: New York, 1972; Chapter 3.2, pp 115–140. [Google Scholar]
  6. Whitehead R. J.; Handy N. C. Variational calculation of vibration-rotation energy levels for triatomic molecules. J. Mol. Spectrosc. 1975, 55, 356–373. 10.1016/0022-2852(75)90274-x. [DOI] [Google Scholar]
  7. Bowman J. M. Self-consistent field energies and wavefunctions for coupled oscillators. J. Chem. Phys. 1978, 68, 608–610. 10.1063/1.435782. [DOI] [Google Scholar]
  8. Dunn K. M.; Boggs J. E.; Pulay P. Vibrational energy levels of hydrogen cyanide. J. Chem. Phys. 1986, 85, 5838–5846. 10.1063/1.451545. [DOI] [Google Scholar]
  9. Bowman J. M. The self-consistent-field approach to polyatomic vibrations. Acc. Chem. Res. 1986, 19, 202–208. 10.1021/ar00127a002. [DOI] [Google Scholar]
  10. Jung J. O.; Gerber R. B. Vibrational wave functions and spectroscopy of (H2O)n, n = 2,3,4,5: Vibrational self-consistent field with correlation corrections. J. Chem. Phys. 1996, 105, 10332–10348. 10.1063/1.472960. [DOI] [Google Scholar]
  11. Carter S.; Culik S. J.; Bowman J. M. Vibrational self-consistent field method for many-mode systems: A new approach and application to the vibrations of CO adsorbed on Cu(100). J. Chem. Phys. 1997, 107, 10458. 10.1063/1.474210. [DOI] [Google Scholar]
  12. Christiansen O. Møller-Plesset perturbation theory for vibrational wave functions. J. Chem. Phys. 2003, 119, 5773–5781. 10.1063/1.1601593. [DOI] [Google Scholar]
  13. Christiansen O. Vibrational coupled cluster theory. J. Chem. Phys. 2004, 120, 2149–2159. 10.1063/1.1637579. [DOI] [PubMed] [Google Scholar]
  14. Cassam-Chenaï P.; Liévin J. The VMFCI method: A flexible tool for solving the molecular vibration problem. J. Comput. Chem. 2006, 27, 627–640. 10.1002/jcc.20374. [DOI] [PubMed] [Google Scholar]
  15. Christiansen O. Vibrational structure theory: new vibrational wave function methods for calculation of anharmonic vibrational energies and vibrational contributions to molecular properties. Phys. Chem. Chem. Phys. 2007, 9, 2942–2953. 10.1039/b618764a. [DOI] [PubMed] [Google Scholar]
  16. Christiansen O. Selected new developments in vibrational structure theory: potential construction and vibrational wave function calculations. Phys. Chem. Chem. Phys. 2012, 14, 6672–6687. 10.1039/c2cp40090a. [DOI] [PubMed] [Google Scholar]
  17. Papousek D.; Aliev M. R.. Molecular Vibrational-Rotational Spectra; Elsevier Scientific Publishing Company, 1982. [Google Scholar]
  18. Califano S.Vibrational States; John Wiley & Sons, 1976. [Google Scholar]
  19. Truhlar D. G.; Olson R. W.; Jeannotte A. C.; Overend J. Anharmonic force constants of polyatomic molecules. Test of the procedure for deducing a force field from the vibration-rotation spectrum. J. Am. Chem. Soc. 1976, 98, 2373–2379. 10.1021/ja00425a001. [DOI] [Google Scholar]
  20. Isaacson A. D.; Truhlar D. G.; Scanlon K.; Overend J. Tests of approximation schemes for vibrational energy levels and partition functions for triatomics: H2O and SO2. J. Chem. Phys. 1981, 75, 3017–3024. 10.1063/1.442394. [DOI] [Google Scholar]
  21. Isaacson A. D.; Hung S.-C. Use of second-order perturbation theory for the vibrational energy levels and partition functions at a saddle point. J. Chem. Phys. 1994, 101, 3928–3935. 10.1063/1.467511. [DOI] [Google Scholar]
  22. Clabo D. A. Jr.; Allen W. D.; Remington R. B.; Yamaguchi Y.; Schaefer H. F. III A systematic study of molecular vibrational anharmonicity and vibration-rotation interaction by self-consistent-field higher-derivative methods. Asymmetric top molecules. Chem. Phys. 1988, 123, 187–239. 10.1016/0301-0104(88)87271-9. [DOI] [Google Scholar]
  23. Maslen P. E.; Handy N. C.; Amos R. D.; Jayatilaka D. Higher analytic derivatives. IV. Anharmonic effects in the benzene spectrum. J. Chem. Phys. 1992, 97, 4233–4254. 10.1063/1.463926. [DOI] [Google Scholar]
  24. Zhang Q.; Day P. N.; Truhlar D. G. The accuracy of second order perturbation theory for multiply excited vibrational energy levels and partition functions for a symmetric top molecular ion. J. Chem. Phys. 1993, 98, 4948–4958. 10.1063/1.464950. [DOI] [Google Scholar]
  25. Barone V. Vibrational zero-point energies and thermodynamic functions beyond the harmonic approximation. J. Chem. Phys. 2004, 120, 3059–3065. 10.1063/1.1637580. [DOI] [PubMed] [Google Scholar]
  26. Bloino J.; Biczysko M.; Crescenzi O.; Barone V. Integrated computational approach to vibrationally resolved electronic spectra: Anisole as a test case. J. Chem. Phys. 2008, 128, 244105. 10.1063/1.2943140. [DOI] [PubMed] [Google Scholar]
  27. Vázquez J.; Stanton J. F. Simple(r) algebraic equation for transition moments of fundamental transitions in vibrational second-order perturbation theory. Mol. Phys. 2006, 104, 377–388. 10.1080/00268970500290367. [DOI] [Google Scholar]
  28. Vázquez J.; Stanton J. F. Treatment of Fermi resonance effects on transition moments in vibrational perturbation theory. Mol. Phys. 2007, 105, 101–109. 10.1080/00268970601135784. [DOI] [Google Scholar]
  29. Jensen P.; Bunker P. R.. Computational Molecular Spectroscopy; John Wiley and Sons Ltd: Chichester, U.K., 2000. [Google Scholar]
  30. Willets A.; Handy N. C.; Green W. H. Jr.; Jayatilaka D. Anharmonic Corrections to Vibrational Transition Intensities. J. Phys. Chem. 1990, 94, 5608–5616. 10.1021/j100377a038. [DOI] [Google Scholar]
  31. Ruden T. A.; Taylor P. R.; Helgaker T. Automated calculation of fundamental frequencies: Application to AlH3 using the coupled-cluster singles-and-doubles with perturbative triples method. J. Chem. Phys. 2003, 119, 1951–1960. 10.1063/1.1583671. [DOI] [Google Scholar]
  32. Barone V. Anharmonic vibrational properties by a fully automated second-order perturbative approach. J. Chem. Phys. 2005, 122, 014108. 10.1063/1.1824881. [DOI] [PubMed] [Google Scholar]
  33. Carbonnière P.; Dargelos A.; Pouchan C. The VCI-P code: an iterative variation-perturbation scheme for efficient computations of anharmonic vibrational levels and IR intensities of polyatomic molecules. Theor. Chem. Acc. 2010, 125, 543–554. 10.1007/s00214-009-0689-7. [DOI] [Google Scholar]
  34. Krasnoshchekov S. V.; Isayeva E. V.; Stepanov N. F. Numerical-Analytic Implementation of the Higher-Order Canonical Van Vleck Perturbation Theory for the Interpretation of Medium-Sized Molecule Vibrational Spectra. J. Phys. Chem. A 2012, 116, 3691–3709. 10.1021/jp211400w. [DOI] [PubMed] [Google Scholar]
  35. Gaw F.; Willetts A.; Handy N.; Green W.. Chapter SPECTRO—a program for derivation of spectroscopic constants from provided quartic force fields and cubic dipole fields. Advances in Molecular Vibrations and Collision Dynamics; Bowman J. M., Ed.; JAI Press, 1992; Vol. 1, pp 169–185. [Google Scholar]
  36. Dressler S.; Thiel W. Anharmonic force fields from density functional theory. Chem. Phys. Lett. 1997, 273, 71–78. 10.1016/s0009-2614(97)00552-6. [DOI] [Google Scholar]
  37. Barone V.; Bloino J.; Guido C. A.; Lipparini F. A fully automated implementation of VPT2 Infrared intensities. Chem. Phys. Lett. 2010, 496, 157–161. 10.1016/j.cplett.2010.07.012. [DOI] [Google Scholar]
  38. Bloino J.; Biczysko M.; Barone V. General Perturbative Approach for Spectroscopy, Thermodynamics, and Kinetics: Methodological Background and Benchmark Studies. J. Chem. Theory Comput. 2012, 8, 1015–1036. 10.1021/ct200814m. [DOI] [PubMed] [Google Scholar]
  39. Hermes M. R.; Hirata S. Second-order many-body perturbation expansions of vibrational Dyson self-energies. J. Chem. Phys. 2013, 139, 034111. 10.1063/1.4813123. [DOI] [PubMed] [Google Scholar]
  40. Werner H.-J.; Knowles P. J.; Knizia G.; Manby F. R.; Schütz M.; Celani P.; Korona T.; Lindh R.; Mitrushenkov A.; Rauhut G.; Shamasundar K. R.; Adler T. B.; Amos R. D.; Bernhardsson A.; Berning A.; Cooper D. L.; Deegan M. J. O.; Dobbyn F.; Eckert A. J.; Goll E.; Hampel C.; Hesselmann A.; Hetzer G.; Hrenar T.; Jansen G.; Köppl C.; Liu Y.; Lloyd A. W.; Mata R. A.; May A. J.; McNicholas S. J.; Meyer W.; Mura M. E.; Nicklass A.; O’Neill D. P.; Palmieri P.; Peng D.; Pflüger R.; Pitzer K.; Reiher M.; Shiozaki T.; Stoll H.; Stone A. J.; Tarroni R.; Thorsteinsson T.; Wang M.. MOLPRO, a Package of Ab Initio Programs, version 2012.1; Ab Initio, 2012.
  41. Kemble E. C.The Fundamental Principles of Quantum Mechanics; Dover publications: New York, 2005. [Google Scholar]
  42. Messiah A.; Schiff L. I.. Quantum Mechanics; McGraw-Hill College, 1968; Vol. 643. [Google Scholar]
  43. Franke P. R.; Brice J. T.; Moradi C. P.; Schaefer H. F.; Douberly G. E. Ethyl + O2 in Helium Nanodroplets: Infrared Spectroscopy of the Ethylperoxy Radical. J. Phys. Chem. A 2019, 123, 3558–3568. 10.1021/acs.jpca.9b01867. [DOI] [PubMed] [Google Scholar]
  44. Rosnik A. M.; Polik W. F. VPT2+K spectroscopic constants and matrix elements of the transformed vibrational Hamiltonian of a polyatomic molecule with resonances using Van Vleck perturbation theory. Mol. Phys. 2014, 112, 261–300. 10.1080/00268976.2013.808386. [DOI] [Google Scholar]
  45. Piccardo M.; Bloino J.; Barone V. Generalized vibrational perturbation theory for rotovibrational energies of linear, symmetric and asymmetric tops: Theory, approximations, and automated approaches to deal with medium-to-large molecular systems. Int. J. Quantum Chem. 2015, 115, 948–982. 10.1002/qua.24931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Mendolicchio M.; Bloino J.; Barone V. General Perturb-Then-Diagonalize Model for the Vibrational Frequencies and Intensities of Molecules Belonging to Abelian and Non-Abelian Symmetry Groups. J. Chem. Theory Comput. 2021, 17, 4332–4358. 10.1021/acs.jctc.1c00240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Gong J. Z.; Matthews D. A.; Changala P. B.; Stanton J. F. Fourth-order vibrational perturbation theory with the Watson Hamiltonian: report of working equations and preliminary results. J. Chem. Phys. 2018, 149, 114102. 10.1063/1.5040360. [DOI] [PubMed] [Google Scholar]
  48. Tew D. P.; Handy N. C.; Carter S.; Irle S.; Bowman J. The internal coordinate path Hamiltonian; application to methanol and malonaldehyde. Mol. Phys. 2003, 101, 3513–3525. 10.1080/0026897042000178079. [DOI] [Google Scholar]
  49. Bowman J. M.; Huang X.; Handy N. C.; Carter S. Vibrational levels of methanol calculated by the reaction path version of MULTIMODE, using an ab initio, full-dimensional potential. J. Phys. Chem. A 2007, 111, 7317–7321. 10.1021/jp070398m. [DOI] [PubMed] [Google Scholar]
  50. Miller W. H.; Handy N. C.; Adams J. E. Reaction path Hamiltonian for polyatomic molecules. J. Chem. Phys. 1980, 72, 99–112. 10.1063/1.438959. [DOI] [Google Scholar]
  51. Page M.; McIver J. W. On evaluating the reaction path Hamiltonian. J. Chem. Phys. 1988, 88, 922–935. 10.1063/1.454172. [DOI] [Google Scholar]
  52. Jackels C. F.; Gu Z.; Truhlar D. G. Reaction-path potential and vibrational frequencies in terms of curvilinear internal coordinates. J. Chem. Phys. 1995, 102, 3188–3201. 10.1063/1.468630. [DOI] [Google Scholar]
  53. Carrington T.; Miller W. H. Reaction surface Hamiltonian for the dynamics of reactions in polyatomic systems. J. Chem. Phys. 1984, 81, 3942–3950. 10.1063/1.448187. [DOI] [Google Scholar]
  54. Koch A.; Billing G. D. The reaction volume Hamiltonian model: Further development and application. J. Chem. Phys. 1997, 107, 7242–7251. 10.1063/1.474965. [DOI] [Google Scholar]
  55. Light J. C.; Hamilton I. P.; Lill J. V. Generalized discrete variable approximation in quantum mechanics. J. Chem. Phys. 1985, 82, 1400–1409. 10.1063/1.448462. [DOI] [Google Scholar]
  56. Colbert D. T.; Miller W. H. A novel discrete variable representation for quantum mechanical reactive scattering via the S-matrix Kohn method. J. Chem. Phys. 1992, 96, 1982–1991. 10.1063/1.462100. [DOI] [Google Scholar]
  57. Isaacson A. D. Including Anharmonicity in the Calculation of Rate Constants. 1. The HCN/HNC Isomerization Reaction. J. Phys. Chem. A 2006, 110, 379–388. 10.1021/jp058113y. [DOI] [PubMed] [Google Scholar]
  58. Pulay P.; Fogarasi G. Geometry optimization in redundant internal coordinates. J. Chem. Phys. 1992, 96, 2856–2860. 10.1063/1.462844. [DOI] [Google Scholar]
  59. Baiardi A.; Bloino J.; Barone V. Simulation of Vibronic Spectra of Flexible Systems: Hybrid DVR-Harmonic Approaches. J. Chem. Theory Comput. 2017, 13, 2804–2822. 10.1021/acs.jctc.7b00236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Quade C. R. Internal coordinate formulation for the vibration–rotation energies of polyatomic molecules. J. Chem. Phys. 1976, 64, 2783–2795. 10.1063/1.432577. [DOI] [PubMed] [Google Scholar]
  61. Bakken V.; Helgaker T. The efficient optimization of molecular geometries using redundant internal coordinates. J. Chem. Phys. 2002, 117, 9160–9174. 10.1063/1.1515483. [DOI] [Google Scholar]
  62. Baiardi A.; Bloino J.; Barone V. Accurate simulation of Resonance-Raman spectra of flexible molecules: an internal coordinates approach. J. Chem. Theory Comput. 2015, 11, 3267–3280. 10.1021/acs.jctc.5b00241. [DOI] [PubMed] [Google Scholar]
  63. Wilson E. B. A Method of Obtaining the Expanded Secular Equation for the Vibration Frequencies of a Molecule. J. Chem. Phys. 1939, 7, 1047–1052. 10.1063/1.1750363. [DOI] [Google Scholar]
  64. Podolsky B. Quantum-mechanically correct form of Hamiltonian function for conservative systems. Phys. Rev. 1928, 32, 812. 10.1103/physrev.32.812. [DOI] [Google Scholar]
  65. Schaad L. J.; Hu J. The schrödinger equation in generalized coordinates. J. Mol. Struct.: THEOCHEM 1989, 185, 203–215. 10.1016/0166-1280(89)85014-6. [DOI] [Google Scholar]
  66. Lauvergnat D.; Nauts A. Exact numerical computation of a kinetic energy operator in curvilinear coordinates. J. Chem. Phys. 2002, 116, 8560–8570. 10.1063/1.1469019. [DOI] [Google Scholar]
  67. Mátyus E.; Czakó C.; Császár A. G. Toward black-box-type full- and reduced-dimensional variational (ro)vibrational computations. J. Chem. Phys. 2009, 130, 134112. 10.1063/1.3076742. [DOI] [PubMed] [Google Scholar]
  68. Wilson E. B.; Decius J. C.; Cross P. C.. Molecular Vibrations: The Theory of Infrared and Raman Vibrational Spectra; McGraw-Hill Book Company, 1955. [Google Scholar]
  69. Quade C. R. Internal coordinate formulation for the vibration–rotation energies of polyatomic molecules. J. Chem. Phys. 1976, 64, 2783–2795. 10.1063/1.432577. [DOI] [PubMed] [Google Scholar]
  70. Martin J. M. L.; Lee T. J.; Taylor P. M.; François J.-P. The anharmonic force field of ethylene, C2H4, by means of accurate ab initio calculations. J. Chem. Phys. 1995, 103, 2589–2602. 10.1063/1.469681. [DOI] [Google Scholar]
  71. Yang Q.; Mendolicchio M.; Barone V.; Bloino J. Accuracy and Reliability in the Simulation of Vibrational Spectra: A Comprehensive Benchmark of Energies and Intensities issuing from Generalized Vibrational Perturbation Theory to Second Order (GVPT2). Front. Astron. Space Sci. 2021, 8, 665232. 10.3389/fspas.2021.665232. [DOI] [Google Scholar]
  72. Baiardi A.; Bloino J.; Barone V. General formulation of vibronic spectroscopy in internal coordinates. J. Chem. Phys. 2016, 144, 084114. 10.1063/1.4942165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-Young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian 16, Revision A.03; Gaussian Inc: Wallingford CT, 2016.
  74. Becke A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. 10.1063/1.464913. [DOI] [Google Scholar]
  75. Zhao Y.; Truhlar D. G. Design of Density Functionals That Are Broadly Accurate for Thermochemistry, Thermochemical Kinetics, and Nonbonded Interactions. J. Phys. Chem. A 2005, 109, 5656–5667. 10.1021/jp050536c. [DOI] [PubMed] [Google Scholar]
  76. Papajak E.; Zheng J.; Xu X.; Leverentz H. R.; Truhlar D. G. Perspectives on basis sets beautiful: Seasonal plantings of diffuse basis functions. J. Chem. Theory Comput. 2011, 7, 3027–3034. 10.1021/ct200106a. [DOI] [PubMed] [Google Scholar]
  77. Grimme S. Semiempirical hybrid density functional with perturbative second-order correlation. J. Chem. Phys. 2006, 124, 034108. 10.1063/1.2148954. [DOI] [PubMed] [Google Scholar]
  78. Dierksen M.; Grimme S. A theoretical study of the chiroptical properties of molecules with isotopically engendered chirality. J. Chem. Phys. 2006, 124, 174301. 10.1063/1.2191044. [DOI] [PubMed] [Google Scholar]
  79. Santra G.; Sylvetsky N.; Martin J. M. Minimally empirical double-hybrid functionals trained against the GMTKN55 database: revDSD-PBEP86-D4, revDOD-PBE-D4, and DOD-SCAN-D4. J. Phys. Chem. A 2019, 123, 5129–5143. 10.1021/acs.jpca.9b03157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Møller C.; Plesset M. S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46, 618–622. 10.1103/PHYSREV.46.618. [DOI] [Google Scholar]
  81. Grimme S.; Antony J.; Ehrlich S.; Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
  82. Grimme S.; Ehrlich S.; Goerigk L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. 10.1002/jcc.21759. [DOI] [PubMed] [Google Scholar]
  83. Martin J. M. L.; Kesharwani M. K. Assessment of CCSD(T)-F12 approximations and basis sets for harmonic vibrational frequencies. J. Chem. Theory Comput. 2014, 10, 2085–2090. 10.1021/ct500174q. [DOI] [PubMed] [Google Scholar]
  84. Agbaglo D.; Fortenberry R. C. The performance of explicitly correlated wavefunctions [CCSD(T)-F12b] in the computation of anharmonic vibrational frequencies. Int. J. Quantum Chem. 2019, 119, e25899 10.1002/qua.25899. [DOI] [Google Scholar]
  85. Heim Z. N.; Amberger B. K.; Esselman B. J.; Stanton J. F.; Woods R. C.; McMahon R. J. Molecular structure determination: Equilibrium structure of pyrimidine from rotational spectroscopy(reSE) and high-level ab initio calculation (re) agree within the uncertainty of experimental measurement. J. Chem. Phys. 2020, 152, 104303. 10.1063/1.5144914. [DOI] [PubMed] [Google Scholar]
  86. Barone V.; Puzzarini C.; Mancini G. Integration of theory, simulation, artificial intelligence and virtual reality: a four-pillar approach for reconciling accuracy and interpretability in computational spectroscopy. Phys. Chem. Chem. Phys. 2021, 23, 17079–17096. 10.1039/d1cp02507d. [DOI] [PubMed] [Google Scholar]
  87. Shimanouchi T.Tables of Molecular Vibrational Frequencies Consolidated; National Bureau of Standards: Washington, DC, 1972; Vol. 1, pp 1–160. [Google Scholar]
  88. Allen V. D.; Csaszar A. G. On the ab-initio determination of higher-order force-constants at nonstationary reference geometries. J. Chem. Phys. 1995, 98, 2983–3015. 10.1063/1.464127. [DOI] [Google Scholar]
  89. Willetts A.; Gaw J. F.; Handy N. C.; Carter S. A. A study of the ground electronic state of hydrogen peroxide. J. Mol. Spectrosc. 1989, 135, 370–388. 10.1016/0022-2852(89)90163-x. [DOI] [Google Scholar]
  90. Fornaro T.; Burini D.; Biczysko M.; Barone V. Hydrogen-bonding effects on infrared spectra from anharmonic computations: uracil-water complexes and uracil dimers. J. Phys. Chem. A 2015, 119, 4224–4236. 10.1021/acs.jpca.5b01561. [DOI] [PubMed] [Google Scholar]
  91. Barone V.; Ceselin G.; Fusè M.; Tasinato N. Accuracy Meets Interpretability for Computational Spectroscopy by Means of Hybrid and Double-Hybrid Functionals. Front. Chem. 2020, 8, 584203. 10.3389/fchem.2020.584203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Puzzarini C.; Biczysko M.; Barone V.; Peña I.; Cabezas C.; Alonso J. L. Accurate molecular structure and spectroscopic properties of nucleobases: a combined computational–microwave investigation of 2-thiouracil as a case study. Phys. Chem. Chem. Phys. 2013, 15, 16965–16975. 10.1039/c3cp52347k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Graindourze M.; Smets J.; Zeegers-Huyskens T.; Maes G. Fourier transform—infrared spectroscopic study of uracil derivatives and their hydrogen bonded complexes with proton donors: Part I. Monomer infrared absorptions of uracil and some methylated uracils in argon matrices. J. Mol. Struct. 1990, 222, 345–364. 10.1016/0022-2860(90)85045-k. [DOI] [Google Scholar]
  94. Szczesniak M.; Nowak M.; Rostkowska H.; Szczepaniak K.; Person W. B.; Shugar D. Matrix isolation studies of nucleic acid constituents. 1. Infrared spectra of uracil monomers. J. Am. Chem. Soc. 1983, 105, 5969–5976. 10.1021/ja00357a002. [DOI] [Google Scholar]
  95. Chin S.; Scott I.; Szczepani K.; Person W. B. Matrix isolation studies of nucleic acid constituents. 2. Quantitative ab initio prediction of the infrared spectrum of in-plane modes of uracil. J. Am. Chem. Soc. 1984, 106, 3415–3422. 10.1021/ja00324a006. [DOI] [Google Scholar]
  96. Barone V.; Biczysko M.; Bloino J.; Puzzarini C. Glycine conformers: a never-ending story?. Phys. Chem. Chem. Phys. 2013, 15, 1358–1363. 10.1039/c2cp43884d. [DOI] [PubMed] [Google Scholar]
  97. Huisken F.; Werhahn O.; Ivanov A. Y.; Krasnokutski S. The O–H stretching vibrations of glycine trapped in rare gas matrices and helium clusters. J. Chem. Phys. 1999, 111, 2978–2984. 10.1063/1.479580. [DOI] [Google Scholar]
  98. Bazsó G.; Magyarfalvi G.; Tarczay G. Tunneling lifetime of the ttc/VIp conformer of glycine in low-temperature matrices. J. Phys. Chem. A 2012, 116, 10539–10547. 10.1021/jp3076436. [DOI] [PubMed] [Google Scholar]
  99. Stepanian S.; Reva I.; Radchenko E.; Rosado M.; Duarte M.; Fausto R.; Adamowicz L. Matrix-isolation infrared and theoretical studies of the glycine conformers. J. Phys. Chem. A 1998, 102, 1041–1054. 10.1021/jp973397a. [DOI] [Google Scholar]
  100. Balabin R. M. Conformational equilibrium in glycine: experimental jet-cooled Raman spectrum. J. Phys. Chem. Lett. 2010, 1, 20–23. 10.1021/jz900068n. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ct2c00773_si_001.pdf (241.5KB, pdf)

Articles from Journal of Chemical Theory and Computation are provided here courtesy of American Chemical Society

RESOURCES