This work improves and extends a general and robust workflow for the computation of anharmonic vibrational frequencies to thermodynamic functions, paving the way toward the study of large flexible molecules. The key new feature is the extension of closed-form expressions for both zero-point vibrational energies and partition functions to second-order vibrational perturbation theory based on curvilinear internal coordinates. The use of curvilinear coordinates enables the reduction of couplings between different degrees of freedom, enriching the arsenal of existing vibrational approaches, and can lead to effective, low-dimensional linear-scaling models. The accuracy of the results obtained for some prototypical systems paves the way toward the systematic use of this new implementation in the study of molecules containing a few dozen atoms, as exemplified by the test cases of a molecular motor, a nucleoside, and two hormones.
I. INTRODUCTION
The challenge of performing accurate simulations of large systems has been a fundamental issue in theoretical and computational chemistry since its inception and is likely to remain a central concern in the future.1 On the one hand, state-of-the-art quantum chemical (QC) methods are producing results with an accuracy rivaling that of experimental data for small semirigid systems.2–4 For molecules not plagued by strong static correlation, the coupled cluster model, including single, double, and (perturbatively) triple excitations [CCSD(T)],5 is considered the golden standard of contemporary quantum chemistry, provided that core-valence correlation and complete basis set extrapolation are properly handled.1 Unfortunately, the unfavorable scaling of this model with the size of the system significantly limits its application to large, flexible systems, which are of great fundamental and technological interest. This limitation is particularly pronounced in scenarios involving rugged potential energy surfaces (PESs), characterized by numerous low-energy minima separated by energy barriers of different heights. At the opposite end of the spectrum, while simplified QC models can accommodate much larger systems, they do so at the expense of substantial reductions in computational accuracy.
The latest developments of hardware and software are producing accurate low-scaling methods for the computation of electronic energies.6–9 However, thermochemical studies also require, in addition to accurate electronic energies, partition functions that depend on precise molecular structures, zero-point vibrational energies (ZPVEs), and finite-temperature (FT) contributions. These contributions benefit only marginally from the developments mentioned above, which have not yet produced analytical energy derivatives. Therefore, methods rooted in density functional theory (DFT) offer at present the only viable route, with the latest double hybrid functionals yielding remarkably accurate molecular structures and harmonic force fields.10,11 From another point of view, the standard harmonic oscillator (HO) model produces significant errors, mainly in the high-frequency regime (overestimating ZPVE contributions) and low-frequency regime (overestimating entropy contributions). The first issue can be effectively solved by the second-order vibrational perturbation theory (VPT2),12–16 provided that the presence of resonances is properly taken into account.17–20 The second problem requires the separation between different degrees of freedom, which can be performed safely only if the couplings between different modes are sufficiently small. Although Cartesian coordinates are usually employed in this connection, nonreactive potential energy surfaces can also be described in terms of a unique underlying bond pattern retained by all considered structures. These topological features allow for the automatic definition of the basic variables (bond lengths, valence, and dihedral angles),21 leading to a good description of both small- and large-amplitude internal motions (SAM and LAM, respectively), while minimizing intermode couplings at the same time.
This work is placed in the framework of a general computational strategy utilizing curvilinear internal coordinates,22 which we believe provides a robust, accurate, and effective solution to a variety of challenges in molecular modeling. The main steps of the strategy can be summarized as follows:
Black-box generation of a complete set of curvilinear internal coordinates, distinguishing between hard and soft degrees of freedom within the molecular system.23
Exploration of soft degrees of freedom by means of constrained geometry optimizations. This is achieved through a fast semi-empirical method,24 guided by suitable algorithms,25–27 aimed at identifying low-lying energy minima.
Refinement of the initial results by full geometry optimizations with DFT methods.28
Identification of inter-conversion paths between adjacent minima and removal of energy minima that can effectively relax into more stable structures under the relevant experimental conditions.22
Final refinement of structures, electronic energies and, possibly, equilibrium values of physical–chemical properties by different variants of the new Pisa composite schemes (PCS).29–31
Computation of ZPVEs and thermodynamic functions by the simple perturbation theory (SPT) model32 employing VPT2 data for SAMs and different effective one-dimensional models for individual LAMs.33
Although these steps have recently been validated and incorporated into a general computational tool,23,34 the calculation of ZPVEs and thermodynamic functions beyond the harmonic oscillator model requires further refinement. In fact, current implementations of VPT2 predominantly rely on rectilinear coordinates, which are not well-suited for accurate description of soft degrees of freedom such as torsions or inversions, resulting in significant couplings between SAMs and LAMs. The recent development of a generalized VPT2 engine that employs curvilinear internal coordinates35–37 can address these challenges. The main aim of the present work is to extend this implementation to the computation of ZPVEs and thermodynamic functions. We will demonstrate that many couplings between different vibrational modes can be effectively neglected when using curvilinear internal coordinates, allowing for a clear-cut separation of LAMs from their small-amplitude counterparts, which can then be treated as effective one-dimensional modes.
In particular, the one-dimensional hindered rotor model proposed by Schlegel and co-workers,38 when combined with the VPT2 framework for other vibrational modes, yields remarkably accurate vibrational entropies for both semi-rigid and flexible molecules.39 At the same time, uncoupled non-periodic large-amplitude motions can be described accurately by one-dimensional variational approaches employing the same local information underlying the VPT2 model.40
This paper is organized as follows: in Sec. II, after a short recap of VPT2 employing different kinds of coordinates, we generalize the well-known expressions of ZPVE and partition functions to curvilinear coordinates and introduce a general framework for the variational treatment of LAMs employing the same information needed for performing VPT2 computations. Then, after validating the proposed procedure with several small molecules for which accurate experimental results are available, we will analyze larger molecules, including a molecular motor, a nucleoside, and two hormones. The main conclusions and most promising perspectives are summarized in Sec. VI.
II. THEORETICAL BACKGROUND
In this section, the theoretical background underlying a general second-order perturbative treatment of vibro-rotational motions of isolated molecules encompassing both rectilinear (VPT2-R) and curvilinear (VPT2-C) coordinates will be summarized. Next, a resonance-free expression of ZPVE within the VPT2-C framework will be derived. The attention will then be focused on the development of a black-box algorithm for the approximate treatment of Fermi resonances (FRs) aimed at the calculation of thermochemical parameters without any threshold selection or variational treatment. Finally, the theoretical framework will be extended to the so-called reduced-dimensionality approach,41,42 in which the number of modes investigated at the VPT2 level can be lower than that of normal modes. More in detail, the case of a single degree of freedom separated from the rest of modes will be explicitly considered, paving the route toward the extension of the VPT2-C framework to the treatment of first-order transition states (TSs).
A. Vibro-rotational problem in curvilinear coordinates
The vibro-rotational motions of isolated molecules have been treated by different methods,43–65 ranging from perturbation theory12–19 to variational and quasi-variational models,66–73 not to speak of time-dependent approaches. In this work, we will employ the VPT2 model based on curvilinear coordinates (VPT2-C), whose main aspects are summarized in this section, with the aim of providing the theoretical background for the new developments presented in the present work.
As already mentioned for the matrix χ, the VPT2-R results are recovered, neglecting the derivatives of the matrix g in Eqs. (21) and (22). Then, the eigenvalues arising from the diagonalization of the different blocks of the variational matrix provide the GVPT2 energies, while the eigenvectors collect the coefficients of the harmonic wave functions necessary to build the variational states.
Although the GVPT2 model produces remarkably accurate vibrational states for semi-rigid molecules, thermochemical (and kinetic) studies require lower accuracy but general procedures providing a continuous treatment of resonances for different stationary points and/or quantum chemical models. Therefore, we extended to curvilinear coordinates an approximate model with these characteristics92,93 and further validated its accuracy.
B. Zero-point vibrational energy
C. Soft degrees of freedom
As is well known, accurate descriptions of highly anharmonic degrees of freedom cannot be obtained by the VPT2-R model. This limitation stems from its reliance on a fourth-degree polynomial representation of the PES in terms of rectilinear coordinates. Previous studies have shown that a more effective approach involves leveraging the interplay between different methodologies (e.g., perturbative and variational techniques) to address each vibrational mode with the most suitable method.
From this standpoint, a curvilinear description of internal motions can significantly reduce the magnitude of anharmonic terms involving different degrees of freedom, a feat that is challenging when using rectilinear displacements. To establish this approach, two preliminary steps are essential. First, the VPT2-C formalism must be fully developed for obtaining an accurate description of SAMs. Second, it is crucial to examine how the removal of the “special” coordinates affects the other frequencies and molecular properties.
It is important to point out that when the diagonal cubic terms ( and ) vanish (e.g., for symmetry reasons), this approach becomes equivalent to a full perturbative calculation for the SAMs subspace.
D. Black-box treatment of Fermi resonances for thermochemistry
Although this model provides an expression of the χ matrix devoid of singularities, its accuracy is strongly dependent on the magnitude of k and ɛ. For example, when ɛ is small, the conventional VPT2 model becomes unreliable, and its replacement with DCPT2 is recommended. Otherwise, large values of ɛ imply that the VPT2 model yields correct results. A classification of all possible combinations of k and ɛ is reported in Table I.
Applicability of the VPT2 and DCPT2 models in the function of the values of ɛ and k.
. | VPT2 . | DCPT2 . | ||
---|---|---|---|---|
k ≪ 1 . | k ≫ 1 . | k ≪ 1 . | k ≫ 1 . | |
ɛ ≪ 1 | ✗ | ✗ | ✓ | ✓ |
ɛ ≫ 1 | ✓ | ✓ | ✓ | ✗ |
. | VPT2 . | DCPT2 . | ||
---|---|---|---|---|
k ≪ 1 . | k ≫ 1 . | k ≪ 1 . | k ≫ 1 . | |
ɛ ≪ 1 | ✗ | ✗ | ✓ | ✓ |
ɛ ≫ 1 | ✓ | ✓ | ✓ | ✗ |
III. IMPLEMENTATION
The theoretical developments summarized in Sec. II have been implemented in a standalone program described in previous work,35,37 including the DCPT2 and HDCPT2 extensions of the VPT2-C workflow, the calculation of ZPVE, and the machinery allowing for one-dimensional vibrational configuration interaction (VCI) calculations including up to sixth-order terms. The program has also been enhanced to include a general thermochemical analysis, which encompasses the calculation of partition functions at the harmonic and SPT levels, together with the derived thermodynamic functions. Such improvements lead, in our opinion, to a versatile tool for spectroscopic and thermochemical studies.
For the sake of clarity, let us summarize the organization of the whole workflow, which basically consists of the following steps:
Geometry optimization and calculation of the force constant matrix through a selected QC program.
Calculation of the harmonic vibrational frequencies and normal modes in curvilinear coordinates, through the standalone program, followed by the generation of the displaced geometries necessary for the finite differences and run of the corresponding single-point calculations by the QC package.
Extraction of the output data, construction of the anharmonic force constants and derivatives of the matrix g through the standalone program.
Anharmonic vibrational analysis, with particular reference to the calculation of transition energies, ZPVE, resonance diagnostics, and thermochemical analysis.
Although interfacing the code with any QC program is straightforward, in the present work, the Gaussian 16 package96 has always been used.
IV. COMPUTATIONAL DETAILS
Different DFT methods have been used, in particular the hybrid B3LYP (hereafter B3) functional97–99 and its double hybrid counterpart revDSD-PBEP86 (hereafter rDSD) functional.100 The empirical dispersion corrections proposed by Grimme (D3 model with Becke–Johnson damping, hereafter, D3BJ101,102) have been systematically included. The B3 model has been always employed in conjunction with the split valence polarized 6-31G* and 6-31+G* (from now on referred to as SVP) basis sets.103–105 The rDSD functional has been employed in conjunction with the 3F12− basis set,29 which is obtained from its cc-pVTZ-F12 counterpart,106 upon removal of the d function on first-row atoms and replacement of the two f functions on second- and third-row atoms by a single f function taken from the cc-pVTZ basis set.107
V. RESULTS AND DISCUSSION
A panel of prototypical molecules is used to evaluate the accuracy of the proposed protocol against data issued from experiments and state-of-the-art computational methodologies. The focus will then shift to systems involving large-amplitude motions, taking formamide and its out-of-plane motion as a case study to highlight the interplay between perturbative and variational approaches. Finally, the anharmonic vibrational and thermochemical properties of larger molecules are analyzed, using possibly dual-level methods.
A. Validation
Semi-rigid molecules are the best choices for validation purposes, as in this case different coordinate choices should provide comparable results. In this context, five representative systems have been selected for a detailed study: formaldehyde, oxirane, imidazole, fluoroamine, and vinyl radical (see Fig. 1).
Molecular structures of (a) formaldehyde, (b) oxirane, (c) vinyl radical, (d) fluoroamine, and (e) imidazole.
Molecular structures of (a) formaldehyde, (b) oxirane, (c) vinyl radical, (d) fluoroamine, and (e) imidazole.
The ZPVE of these systems has been determined through both the VPT2-R and VPT2-C frameworks at the B3/SVP level of theory. The results collected in Table II offer a breakdown of the overall ZPVE into different contributions.
Analysis and comparison of the anharmonic ZPVE (in cm−1) of selected rigid systems between Cartesian and curvilinear coordinates at the B3/SVP level of theory.
. | . | VPT2-R . | VPT2-C . | ||||
---|---|---|---|---|---|---|---|
. | Harm.a . | Anh.b . | W+Cc . | Total . | Anh.d . | ExPot.e . | Total . |
Formaldehyde | 5 874 | −80 | −1 | 5 793 | −77 | −4 | 5 793 |
Oxirane | 12 637 | −172 | 1 | 12 466 | −185 | 14 | 12 466 |
Imidazole | 15 620 | −189 | 1 | 15 432 | −192 | 4 | 15 432 |
Fluoroamine | 6 058 | −101 | −1 | 5 957 | −93 | −8 | 5 957 |
Vinyl radical | 8 041 | −123 | 3 | 7 921 | −106 | −14 | 7 921 |
. | . | VPT2-R . | VPT2-C . | ||||
---|---|---|---|---|---|---|---|
. | Harm.a . | Anh.b . | W+Cc . | Total . | Anh.d . | ExPot.e . | Total . |
Formaldehyde | 5 874 | −80 | −1 | 5 793 | −77 | −4 | 5 793 |
Oxirane | 12 637 | −172 | 1 | 12 466 | −185 | 14 | 12 466 |
Imidazole | 15 620 | −189 | 1 | 15 432 | −192 | 4 | 15 432 |
Fluoroamine | 6 058 | −101 | −1 | 5 957 | −93 | −8 | 5 957 |
Vinyl radical | 8 041 | −123 | 3 | 7 921 | −106 | −14 | 7 921 |
Harmonic ZPVE.
Anharmonic correction arising from the expansion of the potential energy.
Anharmonic correction arising from both the Coriolis effect and Watson’s terms.
Anharmonic correction arising from the expansion of both the potential and kinetic energies.
Extra-potential contribution evaluated at the reference geometry.
In particular, within the VPT2-R framework, the anharmonic correction has been partitioned into contributions arising solely from the anharmonic force constants and those arising from the additional Coriolis contribution. In the VPT2-C framework, the reported values correspond to the anharmonic contributions of the potential and kinetic energies, together with the extra-potential term evaluated at the equilibrium geometry.
The overall anharmonic correction is very close for different choices of coordinates, with the anharmonic contribution consistently dominating. Further analysis revealed that within the VPT2-C framework, the anharmonic correction issued from the potential energy typically exceeds its kinetic counterpart (see the supplementary material for further details).
The next step of the analysis concerns the determination of thermochemical properties. In this connection, the accuracy of the VPT2-C ZPVEs and HDCPT2-C energies that govern fundamental transitions will be analyzed, as both contributions enter the SPT approximation of the vibrational partition functions. Formaldehyde has been selected as a case study because of its well-known spectroscopic properties (above all the presence of a strong FR), which allow an unbiased assessment of the robustness and reliability of the computational workflow. The VPT2, DCPT2, and HDCPT2 models in conjunction with either rectilinear or curvilinear coordinates have been used for the calculation of the anharmonic ZPVE, fundamental wave numbers, absolute entropy (S), and specific heat at constant pressure (cp). The results obtained at the rDSD/3F12− level are compared to accurate computed values108 and available experimental data109 in Table III.
Comparison of the experimental and theoretical fundamental wave numbers (in cm−1), absolute entropy, and specific heat [in cal (mol K)−1] of formaldehyde at the rDSD/3F12− level of theory.
. | . | Curvilinear coords. . | Rectilinear coords. . | . | . | ||||
---|---|---|---|---|---|---|---|---|---|
Rectilinear state . | Symm. . | VPT2 . | DCPT2 . | HDCPT2 . | VPT2 . | DCPT2 . | HDCPT2 . | CCa . | Expt.b . |
ZPVE . | . | 5785 . | 5785 . | 5785 . | 5785 . | 5785 . | 5785 . | ⋯ . | ⋯ . |
|11⟩ | A1 | 2785 | 2786 | 2785 | 2785 | 2786 | 2785 | 2784 | 2783 |
|12⟩ | 1755 | 1755 | 1755 | 1755 | 1755 | 1755 | 1735 | 1746 | |
|13⟩ | 1505 | 1506 | 1506 | 1505 | 1506 | 1506 | 1495 | 1500 | |
|14⟩ | B1 | 1180 | 1180 | 1180 | 1180 | 1180 | 1180 | 1163 | 1167 |
|15⟩ | B2 | 2797 | 2816 | 2817 | 2797 | 2814 | 2813 | 2844 | 2843 |
|16⟩ | 1248 | 1248 | 1248 | 1248 | 1248 | 1248 | 1241 | 1249 | |
Sc | 52.24 | 52.24 | 52.24 | 52.24 | 52.24 | 52.24 | ⋯ | 52.28d | |
e | 8.45 | 8.45 | 8.45 | 8.45 | 8.45 | 8.45 | ⋯ | 8.46d |
. | . | Curvilinear coords. . | Rectilinear coords. . | . | . | ||||
---|---|---|---|---|---|---|---|---|---|
Rectilinear state . | Symm. . | VPT2 . | DCPT2 . | HDCPT2 . | VPT2 . | DCPT2 . | HDCPT2 . | CCa . | Expt.b . |
ZPVE . | . | 5785 . | 5785 . | 5785 . | 5785 . | 5785 . | 5785 . | ⋯ . | ⋯ . |
|11⟩ | A1 | 2785 | 2786 | 2785 | 2785 | 2786 | 2785 | 2784 | 2783 |
|12⟩ | 1755 | 1755 | 1755 | 1755 | 1755 | 1755 | 1735 | 1746 | |
|13⟩ | 1505 | 1506 | 1506 | 1505 | 1506 | 1506 | 1495 | 1500 | |
|14⟩ | B1 | 1180 | 1180 | 1180 | 1180 | 1180 | 1180 | 1163 | 1167 |
|15⟩ | B2 | 2797 | 2816 | 2817 | 2797 | 2814 | 2813 | 2844 | 2843 |
|16⟩ | 1248 | 1248 | 1248 | 1248 | 1248 | 1248 | 1241 | 1249 | |
Sc | 52.24 | 52.24 | 52.24 | 52.24 | 52.24 | 52.24 | ⋯ | 52.28d | |
e | 8.45 | 8.45 | 8.45 | 8.45 | 8.45 | 8.45 | ⋯ | 8.46d |
As expected, the value of the overall anharmonic ZPVE is very similar for the different VPT2 variants, since it does not involve any resonance. Equal wave numbers of the fundamental transitions are also obtained from standard VPT2-C and VPT2-R computations. Otherwise, a slight discrepancy of the DCPT2 and HDCPT2 results is observed for the |15⟩ state, which corresponds to the antisymmetric CH stretching. General trends are comparable in both theoretical frameworks, but there is no direct correspondence of rectilinear anharmonic force constants with elements of the σ and ρ tensors. At the same time, rectilinear and curvilinear k terms in Eq. (63) are not strictly equivalent, leading to different DCPT2 results. Of course, this discrepancy also extends to the HDCPT2 scheme, where the factor Λ introduces an additional variation.
As expected, the transition from the VPT2 model to the DCPT2 and HDCPT2 variants has a non-negligible impact only on the |15⟩ state. Even if the HDCPT2 result is closer to its experimental counterpart, the error (about 30 cm−1) is still non-negligible. Formaldehyde highlights the limits of the HDCPT2 approximation. In fact, both the VPT2 and DCPT2 values of the fundamental under consideration are well below the experimental value of 2843 cm−1, implying that even changing the value of the parameter Λ (with 0 ≤ Λ ≤ 1) would not solve the problem. Of course, a methodology exploiting a variational step for the inclusion of resonance effects (GVPT2)17 is more accurate, leading in this case to a value of 2852 cm−1, which corresponds to a deviation of 9 cm−1 with respect to the experimental data. The comparison of rDSD results with their CCSD(T) counterparts (see columns 5, 8 and 9 of Table III) confirm the remarkable accuracy of this double hybrid functional10,11 that is applicable to much larger molecules. Note that the seemingly large discrepancy for the state |15⟩ (30 cm−1) is reduced to 8 cm−1 when the GVPT2 model is used in all computations.
It is worth mentioning that high-frequency modes have a limited impact on the vibrational partition function and consequently on the resulting thermochemical properties. This is well illustrated by the absolute entropy and specific heat of formaldehyde, whose computed values differ by 0.04 and 0.01 cal (mol K)−1, respectively, from the experimental values reported by Gurvich and co-workers.110
The fundamental transition energies (which rule the SPT vibrational partition functions) have been analyzed for two other prototypical molecules, namely, formic acid and methanol [see Figs. 2(a) and 2(b)].
Molecular structures of (a) formic acid, (b) methanol, and (c) acetaldehyde.
In order to put the new approach in a broader perspective, the results issued from the HDCPT2 approach employing curvilinear coordinates are compared not only with their HDCPT2-C and experimental counterparts but also with the values obtained by using the GVPT2-R method, which takes both FRs and Darling–Dennison resonances (DDRs) into full account.111
Let us start by considering formic acid, whose fundamental wave numbers are collected in Table IV.
Comparison of the experimental and theoretical fundamental wave numbers (in cm−1) of formic acid at the rDSD/3F12− level of theory.
. | . | . | Curvilinear coords. . | Rectilinear coords. . | . | |
---|---|---|---|---|---|---|
Assignment . | Symm. . | Harm. . | HDCPT2 . | HDCPT2 . | GVPT2a . | Expt.b . |
ZPVE . | . | 7443 . | 7343 . | 7343 . | 7343 . | ⋯ . |
OH stretch. | A′ | 3757 | 3568 | 3568 | 3570 | 3570 |
CH stretch. | 3093 | 2931 | 2930 | 2942 | 2943 | |
C=O stretch. | 1808 | 1773 | 1774 | 1770 | 1770 | |
CH bend. | 1412 | 1386 | 1385 | 1383 | 1387 | |
OH bend. | 1313 | 1253 | 1251 | 1230 | 1229 | |
C–O stretch. | 1135 | 1102 | 1102 | 1102 | 1105 | |
OCO bend. | 632 | 624 | 624 | 624 | 625 | |
CH bend. | A″ | 1059 | 1037 | 1037 | 1033 | 1033 |
Torsion | 676 | 652 | 652 | 652 | 638 | |
MAE | 65 | 7 | 7 | 2 | ⋯ |
. | . | . | Curvilinear coords. . | Rectilinear coords. . | . | |
---|---|---|---|---|---|---|
Assignment . | Symm. . | Harm. . | HDCPT2 . | HDCPT2 . | GVPT2a . | Expt.b . |
ZPVE . | . | 7443 . | 7343 . | 7343 . | 7343 . | ⋯ . |
OH stretch. | A′ | 3757 | 3568 | 3568 | 3570 | 3570 |
CH stretch. | 3093 | 2931 | 2930 | 2942 | 2943 | |
C=O stretch. | 1808 | 1773 | 1774 | 1770 | 1770 | |
CH bend. | 1412 | 1386 | 1385 | 1383 | 1387 | |
OH bend. | 1313 | 1253 | 1251 | 1230 | 1229 | |
C–O stretch. | 1135 | 1102 | 1102 | 1102 | 1105 | |
OCO bend. | 632 | 624 | 624 | 624 | 625 | |
CH bend. | A″ | 1059 | 1037 | 1037 | 1033 | 1033 |
Torsion | 676 | 652 | 652 | 652 | 638 | |
MAE | 65 | 7 | 7 | 2 | ⋯ |
Anharmonic calculation based on the GVPT2 scheme accounting for both FRs and DDRs effects.
All the methods tested strongly reduce the mean absolute deviation (MAE) of the harmonic approximation (from 65 to less than 10 cm−1). This result underscores the non-negligible role of anharmonic contributions for a quantitative comparison with the experiment. While HDCPT2-C and HDCPT2-R results are very close, a non-negligible difference is observed with the GVPT2-R approach for CH stretching (11 cm−1) and OH bending (23 cm−1). These discrepancies can be traced back to the approximate treatment of FRs by using the HDCPT2 model, as a detailed analysis indicates that both vibrations are only marginally affected by DDRs. Although this aspect can be relevant for high-resolution spectroscopy, an average error below 10 cm−1 on fundamental transition energies is more than acceptable for the computation of thermochemical properties.
A similar trend is shown by methanol, whose fundamental wave numbers are collected in Table V.
Comparison of the experimental and theoretical fundamental wave numbers (in cm−1) of methanol at the rDSD/3F12− level of theory.
. | . | . | Curvilinear coords. . | Rectilinear coords. . | . | |
---|---|---|---|---|---|---|
Ass. . | Symm. . | Harm. . | HDCPT2 . | HDCPT2 . | GVPT2a . | Expt.b . |
ZPVE . | . | 11 324 . | 11 152 . | 11 152 . | 11 152 . | ⋯ . |
OH stretch. | A′ | 3865 | 3683 | 3683 | 3684 | 3681 |
CH3 d-stretch. | 3145 | 3007 | 3007 | 3016 | 3000 | |
CH3 s-stretch. | 3022 | 2888 | 2888 | 2849 | 2844 | |
CH3 d-deform. | 1527 | 1490 | 1490 | 1484 | 1477 | |
CH3 s-deform. | 1491 | 1447 | 1445 | 1457 | 1455 | |
OH bend. | 1381 | 1332 | 1331 | 1325 | 1345 | |
CH3 rock. | 1089 | 1071 | 1075 | 1074 | 1060 | |
CO stretch. | 1059 | 1033 | 1034 | 1032 | 1033 | |
CH3 d-stretch. | A″ | 3081 | 2953 | 2953 | 2974 | 2960 |
CH3 d-deform. | 1517 | 1484 | 1483 | 1476 | 1477 | |
CH3 rock. | 1183 | 1156 | 1156 | 1156 | 1165 | |
Torsion | 291 | 232 | 232 | 232 | 200 | |
MAE | 67 | 9 | 9 | 7 | ⋯ |
. | . | . | Curvilinear coords. . | Rectilinear coords. . | . | |
---|---|---|---|---|---|---|
Ass. . | Symm. . | Harm. . | HDCPT2 . | HDCPT2 . | GVPT2a . | Expt.b . |
ZPVE . | . | 11 324 . | 11 152 . | 11 152 . | 11 152 . | ⋯ . |
OH stretch. | A′ | 3865 | 3683 | 3683 | 3684 | 3681 |
CH3 d-stretch. | 3145 | 3007 | 3007 | 3016 | 3000 | |
CH3 s-stretch. | 3022 | 2888 | 2888 | 2849 | 2844 | |
CH3 d-deform. | 1527 | 1490 | 1490 | 1484 | 1477 | |
CH3 s-deform. | 1491 | 1447 | 1445 | 1457 | 1455 | |
OH bend. | 1381 | 1332 | 1331 | 1325 | 1345 | |
CH3 rock. | 1089 | 1071 | 1075 | 1074 | 1060 | |
CO stretch. | 1059 | 1033 | 1034 | 1032 | 1033 | |
CH3 d-stretch. | A″ | 3081 | 2953 | 2953 | 2974 | 2960 |
CH3 d-deform. | 1517 | 1484 | 1483 | 1476 | 1477 | |
CH3 rock. | 1183 | 1156 | 1156 | 1156 | 1165 | |
Torsion | 291 | 232 | 232 | 232 | 200 | |
MAE | 67 | 9 | 9 | 7 | ⋯ |
Anharmonic calculation based on the GVPT2 scheme accounting for both FRs and DDRs effects.
Reference 114.
Once again, the inclusion of anharmonic corrections results in a remarkable reduction of the error with respect to experiment. Although the MAEs of the different anharmonic calculations are comparable, the most significant deviations of the HDCPT2 results from their GVPT2-R counterparts are observed in the case of CH3 symmetric stretching (39 cm−1), CH3 degenerate stretching (21 cm−1), and CH3 degenerate deformation (12 cm−1). The discrepancies are always due to the different treatment of FRs, while the impact of DDRs is rather limited.
Then, the analysis has been extended to the calculation of entropies. In this case, we have also considered acetaldehyde [see Fig. 2(c)] that is characterized by a low-frequency rotation of the methyl group. In this context, the simulations have been performed using a curvilinear framework for the description of internal nuclear motions. In particular, the results obtained through harmonic, VPT2-C, and HDCPT2-C computations have been compared to their experimental counterparts to evaluate the effect of FRs on the thermochemical properties of interest (see Table VI).
Comparison between computed and experimental entropies [in cal (mol K)−1] of selected systems at 298.15 and 1 atm. The rDSD/3F12− level of calculation has been systematically adopted.
Molecule . | Harm.a . | VPT2b . | HDCPT2c . | Expt.d . |
---|---|---|---|---|
Formic acid | 59.33 | 59.41 | 59.48 | 59.48 |
Methanol | 57.01 | 57.33 | 57.33 | 57.29 |
Acetaldehydee | 62.66 | 62.67 | 62.67 | 63.06 |
(63.10) | (63.10) |
Molecule . | Harm.a . | VPT2b . | HDCPT2c . | Expt.d . |
---|---|---|---|---|
Formic acid | 59.33 | 59.41 | 59.48 | 59.48 |
Methanol | 57.01 | 57.33 | 57.33 | 57.29 |
Acetaldehydee | 62.66 | 62.67 | 62.67 | 63.06 |
(63.10) | (63.10) |
Harmonic value of the thermochemical property, with the vibrational contribution being evaluated within the Wilson GF method.
Curvilinear-based formulation of VPT2.
Curvilinear-based formulation of HDCPT2.
The experimental data have been reduced by 0.03 cal (mol K)−1 to account for the pressure change from 1 bar = 0.1 MPa to 1 atm = 0.101 325 MPa.
The values in parenthesis are obtained using the HR model for methyl torsion.
Among the investigated systems, formic acid is characterized by a difference between VPT2-C and HDCPT2-C schemes comparable with the overall anharmonic correction (0.07 vs 0.08 cal (mol K)−1), with the HDCPT2-C result in excellent agreement with the experiment. Otherwise, the two perturbative variants provide the same anharmonic correction [0.32 cal (mol K)−1)] in the case of methanol, leading once again to a final value very close to its experimental counterpart. Finally, as expected, the methyl torsion of acetaldehyde is poorly described by any perturbative treatment. In fact, the computed entropy of this system differs from the experimental value by about 0.39 cal (mol K)−1. A viable strategy for addressing hindered rotations (HRs) leverages robust interpolation equations between harmonic oscillator and free-rotor models.33,38 When this approach is used for hindered rotations, all other modes can be treated by the different VPT2 variants. The automatic procedure implemented in the Gaussian16 package detects without problems the torsional mode of acetaldehyde as a hindered rotation, also providing the corresponding fundamental transition energy and contribution to ZPVE. This integrated approach gives a standard entropy of acetaldehyde [63.10 cal (mol K)−1] remarkably close to the experimental value [63.06 cal (mol K)−1].
Now, a different problem is tackled, namely, the presence of soft degrees of freedom different from hindered rotations, which, as is well known, are poorly described by low-order perturbative methods. In such circumstances, an integrated strategy combining perturbative and variational methods can be effectively used, provided that negligible couplings are observed between the degrees of freedom treated at different levels. The approach is illustrated by the case of formamide (see Fig. 3), whose lowest frequency mode, namely, the out-of-plane motion of the NH2 group, has a very large quartic force constant.
To understand the extent of the problem, the wave numbers of the fundamental transitions have been evaluated with the variants VPT2-C or VPT2-R and compared to the corresponding experimental values. The different results are reported in Table VII.
Comparison of the experimental and theoretical fundamental wave numbers (in cm−1) of formamide at the rDSD/3F12− level of theory.
. | . | . | Curvilinear coords. . | Rectilinear coords. . | . | ||
---|---|---|---|---|---|---|---|
Assignment . | Symm. . | Harm. . | VPT2 . | HDCPT2 . | HDCPT2 . | GVPT2a . | Expt.b . |
ZPVE . | . | 9986 . | 9936 . | 9936 . | 9936 . | 9936 . | ⋯ . |
Asym. NH2 stretch. | A′ | 3756 | 3554 | 3554 | 3554 | 3554 | 3564 |
Sym. NH2 stretch. | 3612 | 3436 | 3436 | 3435 | 3435 | 3440 | |
CH stretch. | 2991 | 2840 | 2840 | 2838 | 2866 | 2854 | |
CO stretch. | 1789 | 1756 | 1756 | 1756 | 1756 | 1754 | |
In plane NH2 scis. | 1631 | 1680 | 1607 | 1606 | 1585 | 1579 | |
In plane CH scis. | 1424 | 1395 | 1393 | 1393 | 1396 | 1391 | |
CN str. | 1280 | 1487 | 1267 | 1267 | 1256 | 1258 | |
i-NHO/NH2 bend. | 1059 | 1054 | 1054 | 1055 | 1056 | 1046 | |
o-NHO/NH2 bend. | 568 | 567 | 567 | 569 | 569 | 566 | |
Out of pl. CH def. | A″ | 1048 | 1021 | 1021 | 1023 | 1024 | 1033 |
Torsion | 638 | 587 | 570 | 570 | 590 | 602 | |
Out of pl. NH2 | 175 | 576 | 576 | 576 | 568 | 289 | |
(342)c | (342)c | (608)c | (608)c | ||||
MAEd | 64 | 36 | 11 | 7 | 7 | ||
MAEe | 68 | 57 | 34 | 29 | 29 | ||
MAEf | (29) | (14) | (14) | (14) | (57) |
. | . | . | Curvilinear coords. . | Rectilinear coords. . | . | ||
---|---|---|---|---|---|---|---|
Assignment . | Symm. . | Harm. . | VPT2 . | HDCPT2 . | HDCPT2 . | GVPT2a . | Expt.b . |
ZPVE . | . | 9986 . | 9936 . | 9936 . | 9936 . | 9936 . | ⋯ . |
Asym. NH2 stretch. | A′ | 3756 | 3554 | 3554 | 3554 | 3554 | 3564 |
Sym. NH2 stretch. | 3612 | 3436 | 3436 | 3435 | 3435 | 3440 | |
CH stretch. | 2991 | 2840 | 2840 | 2838 | 2866 | 2854 | |
CO stretch. | 1789 | 1756 | 1756 | 1756 | 1756 | 1754 | |
In plane NH2 scis. | 1631 | 1680 | 1607 | 1606 | 1585 | 1579 | |
In plane CH scis. | 1424 | 1395 | 1393 | 1393 | 1396 | 1391 | |
CN str. | 1280 | 1487 | 1267 | 1267 | 1256 | 1258 | |
i-NHO/NH2 bend. | 1059 | 1054 | 1054 | 1055 | 1056 | 1046 | |
o-NHO/NH2 bend. | 568 | 567 | 567 | 569 | 569 | 566 | |
Out of pl. CH def. | A″ | 1048 | 1021 | 1021 | 1023 | 1024 | 1033 |
Torsion | 638 | 587 | 570 | 570 | 590 | 602 | |
Out of pl. NH2 | 175 | 576 | 576 | 576 | 568 | 289 | |
(342)c | (342)c | (608)c | (608)c | ||||
MAEd | 64 | 36 | 11 | 7 | 7 | ||
MAEe | 68 | 57 | 34 | 29 | 29 | ||
MAEf | (29) | (14) | (14) | (14) | (57) |
Anharmonic calculation based on the GVPT2 scheme accounting for both FRs and DDRs effects.
Reference 115.
Fundamental wave number determined through a one-dimensional VCI calculation along the normal mode associated with the out-of-plane motion of the NH2 group.
Mean absolute error of the theoretical wave numbers, with the fundamental transition of the out-of-plane motion of the NH2 group being neglected.
Mean absolute error of the theoretical wave numbers, with the fundamental transition of the out-of-plane motion of the NH2 group being computed through perturbation theory.
Mean absolute error of the theoretical wave numbers, with the fundamental transition of the out-of-plane motion of the NH2 group being computed through a one-dimensional VCI.
First, the MAE was computed for the different sets of fundamentals without including the LAM. As expected, the corresponding values are in line with those obtained for formic acid and methanol, with the quite large MAE obtained at the harmonic level (68 cm−1) being strongly reduced (to 11 and 7 cm−1) when anharmonic corrections are taken into account. Inclusion of the NH2 inversion in the error analysis leads to significant increases in MAEs related to the huge anharmonic correction associated with this LAM. In order to bypass this problem, a stepwise procedure has been adopted. First, the third- and fourth-order vibrational couplings between this vibration and the other ones have been grouped in different wave-number ranges. The corresponding histograms issued from rectilinear and curvilinear coordinates are reported in Fig. 4.
Graphical representation of the distribution of cubic (left panel) and quartic (right panel) couplings for the out-of-plane motion of the NH2 group in formamide. The results are presented as functions of the wave number range, based on calculations at the rDSD/3F12− level of theory.
Graphical representation of the distribution of cubic (left panel) and quartic (right panel) couplings for the out-of-plane motion of the NH2 group in formamide. The results are presented as functions of the wave number range, based on calculations at the rDSD/3F12− level of theory.
As expected, only the rectilinear anharmonic force constants show values larger than 1000 cm−1 and, in general, at wave numbers higher than those of their curvilinear counterparts. This confirms that the LAM is much less coupled to other modes when employing curvilinear internal coordinates. Next, a VPT2 calculation is carried out in the space of SAMs, retaining the couplings with the LAM, but removing the diagonal anharmonic force constants of this mode, as well as the corresponding derivatives of the g matrix. Note that, in general, this approximation also affects the transition frequencies of the SAMs, due to the neglect of the contribution given by Eq. (41), with F being the LAM. However, in the present case, this term vanishes since the diagonal cubic force constant, and the first-order g derivative along the LAM are null by symmetry, so that the fundamental absorption bands of the SAMs do not change when going from the full perturbative treatment to this reduced-dimensionality variant.
Finally, one-dimensional calculations involving the LAM have been carried out. In previous work, relaxed one-dimensional scans along selected coordinates were performed and the corresponding one-dimensional vibrational Hamiltonian were treated by the discrete variable representation (DVR).62,116,117 Here, we employ a different approach, which does not require any additional computation with respect to those needed by the VPT2 approach. One-dimensional vibrational configuration interaction (VCI) computations have been performed with the same basis functions and anharmonic terms entering the VPT2 treatment in conjunction with force fields expressed in terms of either rectilinear or curvilinear coordinates (and g matrix in the latter case). The results of those computations (summarized in Table VII) show that the use of rectilinear coordinates leads to a fundamental transition energy (608 cm−1) even worse than the corresponding perturbative value (568 cm−1), with this behavior being related to the too large diagonal quartic force constant (more than 30 000 cm−1). In fact, finite displacements along the rectilinear approximation of the out-of-plane distortion involve necessarily non-negligible contributions from NH stretchings, with the consequent large (spurious) energy increase. Otherwise, the out-of-plane motion is well-described by the corresponding normal mode expressed in terms of curvilinear coordinates, and the one-dimensional VCI treatment leads to a value of 342 cm−1, lowering the error with respect to the experimental data from 287 cm−1 to 53 cm−1. For the sake of completeness, the convergence of the VCI procedure with respect to the size of the harmonic basis set has been analyzed. The results sketched in Fig. 5 evidence the robustness of the implemented protocol since convergence within 0.1 cm−1 is achieved with a basis set of 19 Hermite polynomials, and convergence within 0.0001 cm−1 is reached with 50 basis functions.
VCI wave number associated with the out-of-plane motion of the NH2 group of formamide at the rDSD/3F12− level of theory. The results are presented as functions of the basis set size.
VCI wave number associated with the out-of-plane motion of the NH2 group of formamide at the rDSD/3F12− level of theory. The results are presented as functions of the basis set size.
In summary, the use of curvilinear internal coordinates paves the way toward the study of flexible molecules by means of a VPT2 treatment of SAMs in conjunction with effective hindered rotor and VCI one-dimensional treatments for periodic and non-periodic LAMs, respectively. Furthermore a recent benchmark study for 21 closed-shell and 10 open-shell small molecules for which accurate experimental data are available showed that a comparable procedure based on rectilinear coordinates and rDSD/3F12− force fields achieved mean absolute errors of 0.05 kcal mol−1 for ZPVEs and 0.05 cal (mol K)−1 for entropies.39 Therefore, the rDSD/3F12− model will be our reference for larger systems.
B. Toward large molecules
The reliability of this scheme has been tested on two systems involving either hindered rotations (pyruvic acid) or crowded moieties (norbornane) for which accurate computational results and/or experimental data are available.
The Tc conformer of pyruvic acid [see Fig. 6(a)] has been analyzed by different anharmonic models, including dual-level schemes. In particular, full calculations at the B3/SVP and rDSD/3F12− levels have been performed. Furthermore, the anharmonic corrections at the B3/SVP level have been combined to harmonic frequencies at the rDSD/3F12− level and with those at the CCSD(T)/cc-pVTZ level reported in Ref. 118. In addition, in this case, the largest errors characterize the purely harmonic calculations and a substantial improvement (fourfold reduction of the MAE) is produced by anharmonic calculation at the B3/SVP level. All other anharmonic calculations have comparable MAEs (5–6 cm−1) corresponding to further twofold reduction. The most interesting aspect is that the accuracy of either the full rDSD/3F12− calculation, or the corresponding one based on B3/SVP anharmonic corrections rivals that obtained through CCSD(T) harmonic frequencies. As already mentioned, the cost of this last method becomes rapidly prohibitive as molecular size increases, making dual-level approaches entirely based on DFT particularly appealing. Let us stress that this remark is further confirmed by the comparison of the anharmonic wave numbers at the rDSD/3F12− level with those in which the anharmonic corrections are replaced by the corresponding B3/SVP values. From this perspective, only the harmonic frequencies require an enhanced level of theory, implying a significant reduction of the overall computational cost.
Molecular structures of (a) Tc conformer of pyruvic acid and (b) norbornane.
Shaw et al.121 carried out a detailed investigation of norbornane [see Fig. 6(b)] through IR spectroscopy, reporting the experimental fundamentals within the spectral range from 600 to 1600 cm−1. In this context, the computation of the harmonic wave numbers has been performed both at the B3/6-31G* and rDSD/3F12− levels of theory, while the anharmonic corrections have been computed at the B3/6-31G* level in the framework of the HDCPT2 scheme. The results obtained using Eq. (74) are reported in Table VIII and compared to the available experimental data. The MAE obtained through harmonic calculations at the B3/6-31G* level (31 cm−1) decreases to 7 cm−1 after the refinement of the harmonic frequencies and the introduction of the anharmonic corrections, thus leading to remarkable agreement with the experimental data. It should be noted that ten different states exhibit a harmonic correction with absolute value cm−1. As a matter of fact, a non-negligible reduction of the MAE is attributable to the correction of the harmonic terms, underscoring the substantial impact of this refinement.
Comparison of the experimental and theoretical fundamental wave numbers of norbornane (in cm−1) computed at the rDSD//B3 level of theory, through the HDCPT2 scheme.
State . | Symm. . | Harm.a . | ΔHarm.b . | ΔAnharm.c . | Anharm.d . | Expt.e . |
---|---|---|---|---|---|---|
ZPVE . | . | 39 092 . | −98 . | −1321 . | 37 673 . | ⋯ . |
|11⟩ | A1 | 3113 | 2 | −148 | 2967 | ⋯ |
|12⟩ | 3108 | −1 | −149 | 2958 | ⋯ | |
|13⟩ | 3066 | −4 | −108 | 2954 | ⋯ | |
|14⟩ | 3057 | −7 | −135 | 2915 | ⋯ | |
|15⟩ | 1556 | −19 | −42 | 1495 | 1487 | |
|16⟩ | 1524 | −20 | −45 | 1459 | 1455 | |
|17⟩ | 1365 | −11 | −35 | 1319 | 1310 | |
|18⟩ | 1302 | −6 | −31 | 1265 | 1253 | |
|19⟩ | 1178 | −6 | −28 | 1144 | 1140 | |
|110⟩ | 1014 | 7 | −23 | 998 | 989 | |
|111⟩ | 935 | 9 | −19 | 925 | 916 | |
|112⟩ | 890 | 3 | −19 | 874 | 864 | |
|113⟩ | 828 | 2 | −16 | 814 | 818 | |
|114⟩ | 765 | 4 | −15 | 754 | 758 | |
|115⟩ | 406 | −3 | −9 | 394 | 401 | |
|116⟩ | A2 | 3086 | 2 | −149 | 2939 | ⋯ |
|117⟩ | 3054 | −4 | −115 | 2935 | ⋯ | |
|118⟩ | 1513 | −19 | −42 | 1452 | 1451 | |
|119⟩ | 1345 | −3 | −37 | 1305 | 1294 | |
|120⟩ | 1321 | −9 | −33 | 1279 | 1290 | |
|121⟩ | 1253 | −6 | −33 | 1214 | 1212 | |
|122⟩ | 1151 | −5 | −20 | 1126 | 1111 | |
|123⟩ | 975 | 7 | −19 | 963 | 956 | |
|124⟩ | 967 | −4 | −27 | 936 | 949 | |
|125⟩ | 548 | −3 | −11 | 534 | 546 | |
|126⟩ | 158 | −13 | −8 | 137 | 152 | |
|127⟩ | B1 | 3108 | 2 | −149 | 2961 | ⋯ |
|128⟩ | 3102 | 1 | −147 | 2956 | ⋯ | |
|129⟩ | 3064 | −2 | −125 | 2937 | ⋯ | |
|130⟩ | 1532 | −18 | −58 | 1456 | 1466 | |
|131⟩ | 1368 | −11 | −33 | 1324 | 1315 | |
|132⟩ | 1256 | −5 | −39 | 1212 | 1211 | |
|133⟩ | 1141 | −10 | −16 | 1115 | 1103 | |
|134⟩ | 1050 | 1 | −27 | 1024 | 1027 | |
|135⟩ | 972 | 5 | −23 | 954 | 957 | |
|136⟩ | 905 | 9 | −21 | 893 | 877 | |
|137⟩ | 811 | −7 | −19 | 785 | 786 | |
|138⟩ | 453 | −6 | 2 | 449 | 435 | |
|139⟩ | B2 | 3103 | −2 | −137 | 2964 | ⋯ |
|140⟩ | 3090 | 1 | −144 | 2947 | ⋯ | |
|141⟩ | 3054 | −4 | −115 | 2935 | ⋯ | |
|142⟩ | 1527 | −20 | −51 | 1456 | 1462 | |
|143⟩ | 1365 | −13 | −34 | 1318 | 1320 | |
|144⟩ | 1306 | −6 | −32 | 1268 | 1272 | |
|145⟩ | 1288 | −7 | −34 | 1247 | 1244 | |
|146⟩ | 1199 | −3 | −31 | 1165 | 1177 | |
|147⟩ | 1104 | −6 | −22 | 1076 | 1081 | |
|148⟩ | 971 | 4 | −17 | 958 | 956 | |
|149⟩ | 829 | 8 | −18 | 819 | 814 | |
|150⟩ | 769 | 5 | −19 | 755 | 760 | |
|151⟩ | 337 | −5 | −16 | 316 | 336 | |
MAEf | 31 | ⋯ | ⋯ | 7 | ⋯ |
State . | Symm. . | Harm.a . | ΔHarm.b . | ΔAnharm.c . | Anharm.d . | Expt.e . |
---|---|---|---|---|---|---|
ZPVE . | . | 39 092 . | −98 . | −1321 . | 37 673 . | ⋯ . |
|11⟩ | A1 | 3113 | 2 | −148 | 2967 | ⋯ |
|12⟩ | 3108 | −1 | −149 | 2958 | ⋯ | |
|13⟩ | 3066 | −4 | −108 | 2954 | ⋯ | |
|14⟩ | 3057 | −7 | −135 | 2915 | ⋯ | |
|15⟩ | 1556 | −19 | −42 | 1495 | 1487 | |
|16⟩ | 1524 | −20 | −45 | 1459 | 1455 | |
|17⟩ | 1365 | −11 | −35 | 1319 | 1310 | |
|18⟩ | 1302 | −6 | −31 | 1265 | 1253 | |
|19⟩ | 1178 | −6 | −28 | 1144 | 1140 | |
|110⟩ | 1014 | 7 | −23 | 998 | 989 | |
|111⟩ | 935 | 9 | −19 | 925 | 916 | |
|112⟩ | 890 | 3 | −19 | 874 | 864 | |
|113⟩ | 828 | 2 | −16 | 814 | 818 | |
|114⟩ | 765 | 4 | −15 | 754 | 758 | |
|115⟩ | 406 | −3 | −9 | 394 | 401 | |
|116⟩ | A2 | 3086 | 2 | −149 | 2939 | ⋯ |
|117⟩ | 3054 | −4 | −115 | 2935 | ⋯ | |
|118⟩ | 1513 | −19 | −42 | 1452 | 1451 | |
|119⟩ | 1345 | −3 | −37 | 1305 | 1294 | |
|120⟩ | 1321 | −9 | −33 | 1279 | 1290 | |
|121⟩ | 1253 | −6 | −33 | 1214 | 1212 | |
|122⟩ | 1151 | −5 | −20 | 1126 | 1111 | |
|123⟩ | 975 | 7 | −19 | 963 | 956 | |
|124⟩ | 967 | −4 | −27 | 936 | 949 | |
|125⟩ | 548 | −3 | −11 | 534 | 546 | |
|126⟩ | 158 | −13 | −8 | 137 | 152 | |
|127⟩ | B1 | 3108 | 2 | −149 | 2961 | ⋯ |
|128⟩ | 3102 | 1 | −147 | 2956 | ⋯ | |
|129⟩ | 3064 | −2 | −125 | 2937 | ⋯ | |
|130⟩ | 1532 | −18 | −58 | 1456 | 1466 | |
|131⟩ | 1368 | −11 | −33 | 1324 | 1315 | |
|132⟩ | 1256 | −5 | −39 | 1212 | 1211 | |
|133⟩ | 1141 | −10 | −16 | 1115 | 1103 | |
|134⟩ | 1050 | 1 | −27 | 1024 | 1027 | |
|135⟩ | 972 | 5 | −23 | 954 | 957 | |
|136⟩ | 905 | 9 | −21 | 893 | 877 | |
|137⟩ | 811 | −7 | −19 | 785 | 786 | |
|138⟩ | 453 | −6 | 2 | 449 | 435 | |
|139⟩ | B2 | 3103 | −2 | −137 | 2964 | ⋯ |
|140⟩ | 3090 | 1 | −144 | 2947 | ⋯ | |
|141⟩ | 3054 | −4 | −115 | 2935 | ⋯ | |
|142⟩ | 1527 | −20 | −51 | 1456 | 1462 | |
|143⟩ | 1365 | −13 | −34 | 1318 | 1320 | |
|144⟩ | 1306 | −6 | −32 | 1268 | 1272 | |
|145⟩ | 1288 | −7 | −34 | 1247 | 1244 | |
|146⟩ | 1199 | −3 | −31 | 1165 | 1177 | |
|147⟩ | 1104 | −6 | −22 | 1076 | 1081 | |
|148⟩ | 971 | 4 | −17 | 958 | 956 | |
|149⟩ | 829 | 8 | −18 | 819 | 814 | |
|150⟩ | 769 | 5 | −19 | 755 | 760 | |
|151⟩ | 337 | −5 | −16 | 316 | 336 | |
MAEf | 31 | ⋯ | ⋯ | 7 | ⋯ |
B3/6-31G* level of theory.
Difference between the harmonic contributions at the rDSD/3F12− level and those at the B3/6-31G* level.
Anharmonic corrections at the B3/6-31G* level.
Fully anharmonic results.
Reference 121.
Mean absolute error of the fundamental wave numbers with respect to their experimental counterparts, considering only those modes for which experimental values are available.
This kind of computation can be routinely performed for molecules containing up to about 25 atoms. For larger systems, anharmonic force fields can still be obtained at the B3 level, but harmonic computations at the rDSD level become prohibitively expensive in terms of both time and memory requirements. Therefore, one has to rely on B3 results for both harmonic and anharmonic contributions. The errors issued by this approximation have been tested by comparison of rDSD/3F12− and B3/SVP harmonic zero point energies for norbornane together with the panel of molecules shown in Fig. 7.
Molecular structures of (a) benzoquinone, (b) naphthoquinone, (c) (2CN-)naphthalene, (d) (9CN-)anthracene, (e) (9CN-)phenanthrene, and (f) (2CN-)pyrene. X = H, CN.
Molecular structures of (a) benzoquinone, (b) naphthoquinone, (c) (2CN-)naphthalene, (d) (9CN-)anthracene, (e) (9CN-)phenanthrene, and (f) (2CN-)pyrene. X = H, CN.
The results collected in Table X highlight once again not only the importance of an anharmonic description of vibrational motions but also the significant role played by an accurate description of the underlying harmonic force field. In particular, B3 harmonic frequencies are systematically larger than their rDSD counterparts and a scaling factor of 0.996 reduces the average error of ZPVE’s around 0.1 kcal mol−1 for representative hydrocarbons and quinones. The OH stretching of pyruvic acid (whose frequency is underestimated by about 100 cm−1 as shown in Table IX) is paradigmatic of pathological situations (including also CN stretchings in cyano compounds122 or some X–H stretching modes in the presence of hydrogen bridges123) affected by much larger errors. Here, internal coordinates offer significant advantages as a limited number of energies (and/or gradients) along predefined directions is sufficient to determine the required corrections. Although the full development of this approach goes beyond the scope of the present paper, its potentialities are illustrated in Table X for the specific case of cyano derivatives of polyaromatic hydrocarbons (CN-PAHs), which are of increasing importance in the field of astrochemistry.124,125 As a matter of fact, a constant correction of 0.3 kcal mol−1 (1.3 times the nearly constant difference of 77 cm−1 between the harmonic frequencies of CN stretchings computed at the B3 and rDSD level) for each CN group produces errors close to 0.1 kcal mol−1 also for those molecules. Note that an analogous procedure (scaling by 0.996 for all the modes plus 1.3 times the difference of the OH stretching frequency) also performs a very good job for the ZPVE of pyruvic acid [the harmonic B3 value is increased from 46.48 to 46.68 kcal mol−1 upon scaling, to be compared to the identical rDSD and CCSD(T) values of 46.69 kcal mol−1]. Otherwise, the effect of these corrections is negligible on partition functions (hence, entropies and thermal contributions to enthalpies).
Comparison of the experimental and theoretical fundamental wave numbers (in cm−1) of the Tc conformer of pyruvic acid.
. | . | ωa . | δνb . | νc . | ||||||
---|---|---|---|---|---|---|---|---|---|---|
Ass. . | Symm. . | B3d . | rDSDe . | CCf . | B3d . | B3d . | rDSD//B3g . | CC//B3h . | rDSDe . | exp.i . |
ZPVE . | . | 15 674 . | 15 743 . | 15 743 . | −191 . | 15 483 . | 15 552 . | 15 552 . | 15 553 . | ⋯ . |
OH str. | A′ | 3569 | 3665 | 3667 | −205 | 3364 | 3460 | 3462 | 3465 | 3463 |
CH3 a-str. | 3176 | 3181 | 3174 | −155 | 3021 | 3026 | 3019 | 3036 | 3025 | |
CH3 s-str. | 3061 | 3062 | 3056 | −117 | 2944 | 2945 | 2939 | 2948 | 2941 | |
CO str. | 1849 | 1835 | 1853 | −38 | 1811 | 1797 | 1815 | 1801 | 1804 | |
CO str. | 1788 | 1768 | 1773 | −28 | 1760 | 1740 | 1745 | 1733 | 1737 | |
CH3 a-bend. | 1477 | 1468 | 1464 | −34 | 1443 | 1434 | 1430 | 1433 | 1424 | |
CC a-str. | 1421 | 1423 | 1430 | −35 | 1386 | 1388 | 1395 | 1393 | 1391 | |
CH3 s-bend. | 1391 | 1394 | 1394 | −34 | 1357 | 1360 | 1360 | 1344 | 1360 | |
COH bend. | 1248 | 1263 | 1271 | −42 | 1206 | 1221 | 1229 | 1221 | 1211 | |
CO str. | 1163 | 1167 | 1166 | −28 | 1135 | 1139 | 1138 | 1138 | 1133 | |
CH3 rock. | 991 | 988 | 984 | −15 | 976 | 973 | 969 | 972 | 970 | |
CC s-str. | 766 | 777 | 775 | −17 | 749 | 760 | 758 | 761 | 761 | |
CO bend. | 606 | 611 | 608 | −7 | 599 | 604 | 601 | 606 | 604 | |
CO bend. | 525 | 530 | 532 | −1 | 524 | 529 | 531 | 531 | ⋯ | |
CCO bend. | 394 | 394 | 395 | −6 | 388 | 388 | 389 | 393 | 389 | |
CCC bend. | 254 | 253 | 253 | −1 | 253 | 252 | 252 | 267 | 258 | |
CH3 a-str. | A″ | 3118 | 3131 | 3126 | −153 | 2965 | 2978 | 2973 | 2983 | ⋯ |
CH3 a-bend. | 1478 | 1474 | 1470 | −40 | 1438 | 1434 | 1430 | 1431 | ⋯ | |
CH3 rock. | 1040 | 1044 | 1040 | −18 | 1022 | 1026 | 1022 | 1022 | 1030 | |
CO rock. | 741 | 741 | 741 | −26 | 715 | 715 | 715 | 724 | ⋯ | |
OH tors. | 687 | 701 | 701 | −26 | 661 | 675 | 675 | 681 | 668 | |
CO rock. | 395 | 396 | 394 | −3 | 392 | 393 | 391 | 396 | 394 | |
CH3 tors. | 119 | 126 | 124 | 21j | 140 | 147 | 145 | 144k | 134 | |
CC tors. | 90 | 93 | 94 | 5j | 95 | 98 | 99 | 95k | 90 | |
MAEl | 37 | 42 | 43 | ⋯ | 11 | 5 | 6 | 5 | ⋯ |
. | . | ωa . | δνb . | νc . | ||||||
---|---|---|---|---|---|---|---|---|---|---|
Ass. . | Symm. . | B3d . | rDSDe . | CCf . | B3d . | B3d . | rDSD//B3g . | CC//B3h . | rDSDe . | exp.i . |
ZPVE . | . | 15 674 . | 15 743 . | 15 743 . | −191 . | 15 483 . | 15 552 . | 15 552 . | 15 553 . | ⋯ . |
OH str. | A′ | 3569 | 3665 | 3667 | −205 | 3364 | 3460 | 3462 | 3465 | 3463 |
CH3 a-str. | 3176 | 3181 | 3174 | −155 | 3021 | 3026 | 3019 | 3036 | 3025 | |
CH3 s-str. | 3061 | 3062 | 3056 | −117 | 2944 | 2945 | 2939 | 2948 | 2941 | |
CO str. | 1849 | 1835 | 1853 | −38 | 1811 | 1797 | 1815 | 1801 | 1804 | |
CO str. | 1788 | 1768 | 1773 | −28 | 1760 | 1740 | 1745 | 1733 | 1737 | |
CH3 a-bend. | 1477 | 1468 | 1464 | −34 | 1443 | 1434 | 1430 | 1433 | 1424 | |
CC a-str. | 1421 | 1423 | 1430 | −35 | 1386 | 1388 | 1395 | 1393 | 1391 | |
CH3 s-bend. | 1391 | 1394 | 1394 | −34 | 1357 | 1360 | 1360 | 1344 | 1360 | |
COH bend. | 1248 | 1263 | 1271 | −42 | 1206 | 1221 | 1229 | 1221 | 1211 | |
CO str. | 1163 | 1167 | 1166 | −28 | 1135 | 1139 | 1138 | 1138 | 1133 | |
CH3 rock. | 991 | 988 | 984 | −15 | 976 | 973 | 969 | 972 | 970 | |
CC s-str. | 766 | 777 | 775 | −17 | 749 | 760 | 758 | 761 | 761 | |
CO bend. | 606 | 611 | 608 | −7 | 599 | 604 | 601 | 606 | 604 | |
CO bend. | 525 | 530 | 532 | −1 | 524 | 529 | 531 | 531 | ⋯ | |
CCO bend. | 394 | 394 | 395 | −6 | 388 | 388 | 389 | 393 | 389 | |
CCC bend. | 254 | 253 | 253 | −1 | 253 | 252 | 252 | 267 | 258 | |
CH3 a-str. | A″ | 3118 | 3131 | 3126 | −153 | 2965 | 2978 | 2973 | 2983 | ⋯ |
CH3 a-bend. | 1478 | 1474 | 1470 | −40 | 1438 | 1434 | 1430 | 1431 | ⋯ | |
CH3 rock. | 1040 | 1044 | 1040 | −18 | 1022 | 1026 | 1022 | 1022 | 1030 | |
CO rock. | 741 | 741 | 741 | −26 | 715 | 715 | 715 | 724 | ⋯ | |
OH tors. | 687 | 701 | 701 | −26 | 661 | 675 | 675 | 681 | 668 | |
CO rock. | 395 | 396 | 394 | −3 | 392 | 393 | 391 | 396 | 394 | |
CH3 tors. | 119 | 126 | 124 | 21j | 140 | 147 | 145 | 144k | 134 | |
CC tors. | 90 | 93 | 94 | 5j | 95 | 98 | 99 | 95k | 90 | |
MAEl | 37 | 42 | 43 | ⋯ | 11 | 5 | 6 | 5 | ⋯ |
Harmonic result.
Anharmonic correction.
Full anharmonic result.
B3/SVP level of theory.
rDSD/3F12− level of theory.
CCSD(T)/cc-pVTZ level of theory from Ref. 118.
Anharmonic fundamental wave numbers obtained by combining rDSD/3F12− harmonic frequencies with B3/SVP anharmonic corrections.
Anharmonic fundamental wave numbers obtained by combining CCSD(T)/cc-pVTZ harmonic frequencies with B3/SVP anharmonic corrections.
Correction computed through the HR model.
Fundamental wave number computed through the HR model.
Mean absolute error of the computed anharmonic wave numbers with respect to their experimental counterparts.
Comparison of harmonic and anharmonic ZPVE (in kcal mol−1) for medium-sized molecules.
Molecule . | Harm. . | ΔAnharm . | ||
---|---|---|---|---|
B3a . | rDSDb . | B3 scal.c . | B3a . | |
Norbornane | 115.94 | 115.49 | 115.48 | −3.92 |
Benzoquinone | 53.41 | 53.26 | 53.20 | −0.50 |
Naphtoquinone | 83.05 | 83.31 | 82.98 | −0.81 |
Naphtalene | 92.67 | 92.27 | 92.30 | −1.02 |
Anthracene | 121.95 | 121.31 | 121.46 | −1.33 |
Phenanthrene | 122.24 | 121.68 | 121.75 | −1.29 |
Pyrene | 130.21 | 129.56 | 129.69 | −1.14 |
2CN-naphtalene | 91.97 | 91.23 | 91.30d | −0.89 |
9CN-anthracene | 121.16 | 120.39 | 120.38d | −1.24 |
9CN-phenanthrene | 121.39 | 120.61 | 120.60d | −1.21 |
2CN-pyrene | 129.27 | 128.48 | 128.45d | −1.10 |
Molecule . | Harm. . | ΔAnharm . | ||
---|---|---|---|---|
B3a . | rDSDb . | B3 scal.c . | B3a . | |
Norbornane | 115.94 | 115.49 | 115.48 | −3.92 |
Benzoquinone | 53.41 | 53.26 | 53.20 | −0.50 |
Naphtoquinone | 83.05 | 83.31 | 82.98 | −0.81 |
Naphtalene | 92.67 | 92.27 | 92.30 | −1.02 |
Anthracene | 121.95 | 121.31 | 121.46 | −1.33 |
Phenanthrene | 122.24 | 121.68 | 121.75 | −1.29 |
Pyrene | 130.21 | 129.56 | 129.69 | −1.14 |
2CN-naphtalene | 91.97 | 91.23 | 91.30d | −0.89 |
9CN-anthracene | 121.16 | 120.39 | 120.38d | −1.24 |
9CN-phenanthrene | 121.39 | 120.61 | 120.60d | −1.21 |
2CN-pyrene | 129.27 | 128.48 | 128.45d | −1.10 |
B3/SVP level of theory.
rDSD/3F12− level of theory.
B3/SVP value scaled by a constant factor of 0.996.
Reduction of 0.3 kcal mol−1 for the presence of a cyano group.
The results obtained through this approach for molecules containing up to about 50 atoms are shown in Table XI for the two hormones, the molecular motor, and the nucleoside sketched in Fig. 8.
Comparison of the harmonic and anharmonic values of the ZPVE (in kcal mol−1), enthalpy difference between standard conditions and 0 K (ΔH in kcal mol−1), entropy [S in cal (mol K)−1], and specific heat at constant pressure [cp in cal (mol K)−1] at 298 K and 1 atm, for different large-sized systems, evaluated within the VPT2-C framework through the HDCPT2 scheme at the B3/SVP level of theory.
. | Property . | Harm.a . | ΔAnharm.b . | Anharm.c . |
---|---|---|---|---|
Androsterone | ZPVEd | 291.76 | −3.68 | 288.08 |
(290.59) | (286.91) | |||
ΔH | 12.66 | 0.09 | 12.75 | |
S | 137.72 | 2.70 | 140.41 | |
cp | 82.72 | 1.00 | 83.73 | |
β-estradiol | ZPVEd | 244.25 | −3.21 | 241.04 |
(243.27) | (240.06) | |||
ΔH | 11.53 | −0.28 | 11.25 | |
S | 130.79 | −3.33 | 127.46 | |
cp | 74.74 | −0.35 | 74.39 | |
Mol. Motor | ZPVEd | 242.78 | −2.64 | 240.14 |
(241.81) | (239.17) | |||
ΔH | 12.91 | 0.35 | 13.26 | |
S | 142.66 | 8.12 | 150.78 | |
cp | 85.09 | 1.41 | 86.50 | |
Uridine | ZPVEd | 142.62 | −2.02 | 140.60 |
(142.05) | (140.03) | |||
ΔH | 10.01 | −0.08 | 9.93 | |
S | 123.19 | −1.55 | 121.63 | |
cp | 59.94 | −0.47 | 59.47 |
. | Property . | Harm.a . | ΔAnharm.b . | Anharm.c . |
---|---|---|---|---|
Androsterone | ZPVEd | 291.76 | −3.68 | 288.08 |
(290.59) | (286.91) | |||
ΔH | 12.66 | 0.09 | 12.75 | |
S | 137.72 | 2.70 | 140.41 | |
cp | 82.72 | 1.00 | 83.73 | |
β-estradiol | ZPVEd | 244.25 | −3.21 | 241.04 |
(243.27) | (240.06) | |||
ΔH | 11.53 | −0.28 | 11.25 | |
S | 130.79 | −3.33 | 127.46 | |
cp | 74.74 | −0.35 | 74.39 | |
Mol. Motor | ZPVEd | 242.78 | −2.64 | 240.14 |
(241.81) | (239.17) | |||
ΔH | 12.91 | 0.35 | 13.26 | |
S | 142.66 | 8.12 | 150.78 | |
cp | 85.09 | 1.41 | 86.50 | |
Uridine | ZPVEd | 142.62 | −2.02 | 140.60 |
(142.05) | (140.03) | |||
ΔH | 10.01 | −0.08 | 9.93 | |
S | 123.19 | −1.55 | 121.63 | |
cp | 59.94 | −0.47 | 59.47 |
Harmonic value of the thermochemical property, with the vibrational contribution being evaluated within the Wilson GF method.
Anharmonic correction to the thermochemical property.
Curvilinear-based formulation of HDCPT2.
In parenthesis: harmonic values scaled by 0.996.
Molecular structure of (a) androsterone, (b) β-estradiol, (c) molecular motor, and (d) uridine.
Molecular structure of (a) androsterone, (b) β-estradiol, (c) molecular motor, and (d) uridine.
The electronic energies of molecules of this size can be computed with an error within 1 kcal mol−1 through last generation CCSD(T) methods using frozen natural orbitals (FNO) or pair natural orbitals (PNO) possibly paired with explicit correlation (F12) and other techniques to accelerate basis set convergence.6,8,9 It is quite apparent that in this case, anharmonic contributions cannot be neglected and their inclusion at a reasonable cost, together with the scaling of harmonic contributions computed at the B3 level, paves the way toward systematic investigation of large molecules.
VI. CONCLUSION
This work extends and validates a recent second-order perturbative method based on curvilinear internal coordinates for the calculation of thermodynamic functions. In this way, inter-mode couplings are generally reduced with respect to rectilinear descriptions, and above all, the remaining contributions are shifted from potential to kinetic terms. This feature determines the great computational advantage of evaluating a coordinate-dependent kinetic energy operator (from analytical derivatives of internal curvilinear coordinates with respect to Cartesian ones) instead of high-order coupling potentials from expensive electronic structure calculations. Apart from that, the interpretation of mode couplings is simplified, and finally, the resulting deeper understanding of vibrational systems may yield better problem-adapted approximations.
The computational workflow starts with an accurate evaluation of anharmonic zero-point vibrational energies for semi-rigid molecular systems. The hybrid degeneracy-corrected second-order perturbation theory is then applied for the calculation of fundamental vibrational bands, enabling reliable calculations even in pathological cases and automatically handling soft degrees of freedom through accurate one-dimensional models. The simple perturbation theory is finally used to obtain reliable yet computationally inexpensive vibrational partition functions that provide all the needed vibrational contributions to thermodynamic properties. The accuracy of the proposed model has been validated by computing fundamental vibrational transitions for a panel of molecules of different sizes. The reliability of a global hybrid functional paired to a polarized split-valence basis set for anharmonic calculations is also highlighted. Additional improvements are achieved when B3 harmonic frequencies are replaced with their counterparts obtained by a double hybrid functional in conjunction with more extended basis sets. In general, this computational strategy enables a fully unsupervised evaluation of vibrational contributions to thermochemical properties with an overall accuracy close to 0.1 kcal mol−1 in the case of enthalpies and 0.1 cal mol−1 K−1 in the case of entropies for molecules containing up to about 50 atoms.
The proposed approach can be applied alongside any quantum mechanical model for which analytical Hessians are available. This means that, in addition to isolated molecules, a black-box procedure for evaluating thermochemical properties beyond the harmonic approximation is also applicable to solution-phase systems described by the polarizable continuum model, whose analytical second derivatives are available.96 Although further improvements are certainly required, particularly in relation to coupled large-amplitude motions, we think that the present version of the computational workflow can be confidently used for the study of medium-to-large molecular systems in both the gas phase and in solution by experiment-oriented researchers.
Let us finally remark that the methodology proposed in this work has been devised to provide an automatic, black-box approach for thermochemistry, retaining the intuitive system of harmonic oscillator quantum numbers, as well as a treatment of Fermi resonances deprived of parameters defined by the user. Although this framework is, in general, well-suited for the calculation of thermochemical data, it is not necessarily the most accurate approach for the simulation of vibrational spectra. Therefore,the perturbative treatment based on curvilinear coordinates will be shortly extended to its generalized version, explicitly accounting for both Fermi and Darling–Dennison resonances.
SUPPLEMENTARY MATERIAL
The supplementary material includes the definition of the anharmonic χ matrix within the different frameworks of VPT2.
ACKNOWLEDGMENTS
The authors acknowledge the funding from Gaussian, Inc.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
M. Mendolicchio: Conceptualization (equal); Data curation (lead); Formal analysis (lead); Funding acquisition (supporting); Investigation (lead); Methodology (equal); Software (lead); Supervision (lead); Validation (lead); Writing – original draft (equal). V. Barone: Conceptualization (equal); Funding acquisition (lead); Methodology (equal); Validation (supporting); Writing – original draft (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article and its supplementary material.
APPENDIX: CALCULATION OF THE EXTRA-POTENTIAL TERM
Inspection of Eq. (30) shows that the extra-potential term (Vg) depends on first- and second-order derivatives of the determinant (or the coefficients) of the vibro-rotational matrix g. Moreover, the second-order off-diagonal derivatives of g are not required at a stationary point, as the matrix g is diagonal when the Eckart embedding is adopted . Therefore, it is possible to restrict the analysis to the coefficients and .
In summary, calculating Vg requires the complete set of first-order derivatives of the vibro-rotational matrix g, along with the diagonal derivatives of its vibrational and rotational blocks. As these terms are required for evaluating vibro-rotational spectroscopic parameters (see Ref. 37), Vg can be obtained without any additional computational effort.