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.

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:

  1. Black-box generation of a complete set of curvilinear internal coordinates, distinguishing between hard and soft degrees of freedom within the molecular system.23 

  2. 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.

  3. Refinement of the initial results by full geometry optimizations with DFT methods.28 

  4. 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 

  5. 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 

  6. 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.

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).

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 is well known, a molecule composed of Na atoms has N normal modes of vibration (N = 3Na − 6 for a nonlinear molecule and N = 3Na − 5 for a linear one). Within the Born–Oppenheimer approximation,74 the nuclear Hamiltonian can be expressed in the following way:
(1)
where is the reduced Planck constant, a2 is the Laplacian operator with respect to the Cartesian coordinates of the nucleus a in a laboratory-fixed frame (LF), ma is the corresponding mass, and V is the potential energy. After placing the center of mass (COM) of the molecule at the origin of the axes, a body-fixed frame (BF) can be introduced, with the new set of Cartesian coordinates x={x1,,xNa} being related to their LF counterparts by a set of Nr = 3 (2 for linear molecules) Euler angles ϕ={ϕ1,,ϕNr}. The remaining N degrees of freedom represent molecular vibrations, and they can be expressed in terms of a non-redundant set of internal curvilinear coordinates s = {s1, …, sN}. Therefore, the general expression of the Hamiltonian can be derived from Eq. (1) in terms of the generalized coordinates ξ = {XCOM, ϕ, s},75,
(2)
where XCOM is the vector that collects the Cartesian coordinates of the COM in the LF, while Gij and G̃ represent an element and the determinant of the well-known matrix G, respectively. The latter can be conveniently obtained as the inverse of the covariance metric tensor Ḡ, whose elements can be expressed within the t-vector formulation,46,76
(3)
where ti,a = xa/∂ξi. Given that the overall translations can be factored out exactly, the expression of the vibro-rotational Hamiltonian Hvr can be obtained by limiting the summations of Eq. (2) to N′ + Nr, with N′ ≤ N being the number of active vibrations, which is different from the total number of normal modes when the calculations of reduced dimensionality are considered, while N′ = N for full calculations. Following previous work,77,78 and in particular, the formulation proposed by Podolsky79 and Lauvergnat,47 the operator Hvr can be written as follows:
(4)
where the so-called extra-potential term Vg has been introduced,
(5)
Then, the normal coordinates are obtained through Wilson’s GF method,80 
(6)
where Gv is the vibrational diagonal block of the G matrix, Λ is the diagonal matrix collecting the squared angular frequencies, and L is the matrix of eigenvectors. The dimensionless normal coordinates q = {q1, …, qN} can be obtained from their mass-weighted counterparts Q = {Q1, …, QN},
(7)
where c is the light speed in vacuum, ωi is the harmonic wave number (in cm−1) of the normal mode i, and Q = L−1s. The expansion of Hvr in terms of normal coordinates leads to the following expression:
(8)
where f and g represent the degrees of the vibrational and rotational operators, respectively. The purely vibrational terms of Eq. (8) up to the second order define the VPT2 Hamiltonian,
(9)
In Eq. (9), λ is the perturbation parameter (set to 1 at the end of the calculation), while the perturbative terms contain kinetic (g) and potential (f) contributions,
(10)
with pi being the conjugate momentum of qi (pi = −i∂/∂qi). The symbols g and Vg will be adopted hereafter to indicate the matrix G and the extra-potential term expressed in cm−1, while fijkl and gij,kl represent the derivatives of potential energy and the matrix gv with respect to the dimensionless normal coordinates,45 
(11)
The application of second-order Rayleigh–Schrödinger (RSPT)81 or canonical Van Vleck (CVPT)18,82,83 perturbation theory yields the same expression of the energy of a generic vibrational state R,
(12)
where vR,i is the quantum number of the normal mode i, and the χ0 term together with the symmetric χ matrix have been introduced. Previous work was devoted to the determination of transition properties that require only an analytical expression of the χ matrix. In fact, the transition energy from R to S vibrational states is simply given by
(13)
where Δ2vSR,ij = vS,ivS,jvR,ivR,j and ΔvSR,i = vS,ivR,i. Although the VPT2-C model84–86 and its higher-order counterparts87–89 have been analyzed by different authors, only recently, a fully automatized construction of curvilinear coordinates accompanied by compact general expression based on effective tensors has been derived, unifying VPT2-R and VPT2-C models.35 As a result, any code implementing VPT2-R can be easily extended to its VPT2-C counterpart using the following expressions:
(14)
(15)
where
(16)
(17)
(18)
and (Cijχ) is the Coriolis contribution to the χij element, which is present only in the VPT2-R model and is given by
(19)
where Bτeq and ζij,τ are the equilibrium rotational constants and Coriolis interaction term coupling modes i and j along the τ principal axis of inertia. In the VPT2-C model, the (Cijχ) term is lacking, although it survives in the VPT2-R model, where, instead, the derivatives of the g matrix vanish (see the supplementary material for further details).
As is well-known, Eqs. (14) and (15) suffer from the presence of vanishing denominators when ωi ≈ 2ωj or ωiωj + ωk (referred to as FRs of type I and II, respectively). Most strategies aimed at determining FRs90,91 are based on a two-step procedure,the first step being the estimation of the energetic proximity characterizing the interacting states,
(20)
where Δω1−2 is a predefined threshold and j can be equal to k. In the second step, the “weight” of the potentially resonant term is calculated. In the present work, the test proposed by Martin et al.90 has been adopted, in which the term is marked as resonant whenever the deviation of the perturbative contributions to the energies resulting from an ad hoc, variational model is beyond a defined threshold K1−2,
(21)
Once the resonant terms have been identified, they can be simply removed, the resulting model being referred to as deperturbed VPT2 (DVPT2). In order to bypass the accuracy loss due to the elimination of interactions between interacting states, a more refined model, namely, the generalized VPT2 (GVPT2),16,17 can be employed. Within this scheme, a variational matrix is built, whose diagonal entries are represented by the DVPT2 energies. The off-diagonal elements correspond to the pairs of interacting states linked by FRs, whose expression is provided by the corresponding matrix element of the canonical Van Vleck transformed Hamiltonian H̃,
(22)

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.

As already mentioned, in previous work, the attention was focused on the calculation of vibrational transition energies35,36 as well as vibrational and vibro-rotational averages of molecular properties.37 However, a proper characterization of the ZPVE provides invaluable information for different applications, ranging from chemical kinetics and thermochemistry to conformational studies. For this reason, the first objective of the present study has been the development (and application) of an analytical expression of ZPVE within the VPT2-C framework. A first derivation of χ0 was proposed by Isaacson,85,86 also addressing the treatment of doubly degenerate vibrations, needed for the study of symmetric and linear tops. In the present context, a new derivation focusing solely on asymmetric tops has been performed, since non-Abelian symmetry groups can be effectively managed using the unified VPT2 model outlined in Ref. 94,
(23)
with χ0V, χ0T, and χ0× being the potential, kinetic, and cross term, respectively,
(24)
(25)
(26)
where
(27)
Equations (24)(26) are in full agreement with those reported by Isaacson, but an additional analysis based on the partial fraction decomposition allowed us to achieve once again a very compact expression valid for both the VPT2-C and VPT2-R variants,
where the aiiii elements are given by
(28)
and (C0χ) is the Coriolis contribution to χ0, only present in the VPT2-C variant,
(29)
The last contribution to χ0 is due to the extra-potential term, whose expression in wave numbers,
(30)
can be further rewritten by making use of the following notation:
(31)
to obtain a far more manageable equation,
(32)
In general, the extra-potential term is well-approximated by its value for the reference configuration,84,85 which has the following expression in the case of stationary points:
(33)
where the equivalence gijv,eq=ωiδij has been exploited, so that only the diagonal second-order derivatives of g̃ are needed. The terms g̃ij can be directly evaluated from the derivatives of the matrix g or by numerical differentiation of analytical determinants. For the sake of readability, a dedicated section has been included in the  Appendix. The ZPVE is the energy associated with a null vector of quantum numbers (vR,i = 0 with i = 1, …, N′), which is different from χ0 and is denoted by the symbol ɛ0,19,83,95
(34)
A combination of Eqs. (14), (15), (28), and (34) leads to the first expression (to the best of our knowledge) of the resonance-free ZPVE obtained within the VPT2-C formulation. In fact, the following expression gathers both VPT2-C and VPT2-R models:
(35)
which is of course deprived of FRs and where an additional effective tensor has been introduced,
(36)
The term (C0) is the Coriolis contribution to ɛ0, only present in the VPT2-R variant,
(37)

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.

With the purpose of illustration, one can consider a system presenting a single degree of freedom F poorly described at the VPT2 level (N′ = N − 1), with qF being the corresponding normal coordinate. Within a reduced-dimensionality scheme neglecting the overall contribution of F, the terms χFF and χiF (iF) vanish, while χii and χij (with i, jF) are deprived of the contributions defining the T matrix,
(38)
(39)
(40)
Since the main problems faced by VPT2 in the treatment of this class of vibrations could be related not only to the coordinate choice but also to the fourth-order polynomial expansion of the potential energy, the overall elements of the χ matrix can be quite similar for the VPT2-C and VPT2-R variants, suggesting that the former approach does not provide any benefit. However, any shift of the offending contributions from inter- to intra-mode couplings (as expected when one goes from rectilinear to curvilinear coordinates) would permit the use of reduced-dimensionality variational treatments, which can be easily implemented and extended to large molecules. This strategy can be particularly useful in the presence of out-of-scale vibrational couplings that could seriously affect the vibrational properties of the LAM, as well as those of the remaining modes. Conversely, the information provided by those couplings can be preserved and accounted for within the perturbative calculation involving the SAMs, while only the diagonal terms linked to the LAM are removed and used in a subsequent variational framework. In the latter case, all the elements of the T matrix vanish, with the exception of χiF,
(41)

It is important to point out that when the diagonal cubic terms (fFFF and gFF,Fv) vanish (e.g., for symmetry reasons), this approach becomes equivalent to a full perturbative calculation for the SAMs subspace.

Once the coordinate of interest has been separated, the last step is represented by the variational solution of a one-dimensional problem, where the corresponding Hamiltonian matrix elements,
(42)
can be easily obtained through the second-quantization formalism, leading to a set of compact expressions. In this context, the one-dimensional vibrational Hamiltonian HF can be written in the following way:
(43)
where the kinetic and potential energies can be expressed as
(44)
(45)
The symbol VgF denotes the contribution to the extra-potential term arising only from the mode F,
(46)
The expressions of the elements of interest, namely, HvR,FvR,F+n (with n = 1, …, 6) are
(47)
(48)
(49)
(50)
(51)
(52)
(53)
Although this work deals with local minima of the PES, the above-mentioned equations are also applicable to transition states. In particular, the symbol S denotes the sign of the harmonic force constant (i.e., S = 1 for a minimum and S = −1 for a TS), while BR,F(n) is a factor depending on the vibrational quantum numbers,40 
(54)
In the previous sections, the VPT2-C framework has been generalized to the calculation of the ZPVE, as well as to the treatment of a coordinate belonging to a different class, including the case of an imaginary frequency. However, the calculation of the energy levels still suffers from the presence of FRs, with this issue also affecting the calculation of the vibrational partition functions and, therefore, of thermochemical properties. Thrular and Isaacson proposed the so-called simple perturbation theory (SPT) in which all normal modes are assumed to be decoupled to build vibrational partition functions,32 
(55)
but the ZPVEs (ɛ0) and fundamental energy levels (ε1i) are evaluated at the VPT2 level, the latter requiring the detection and proper treatment of resonances. To this end, the matrix χ can be partitioned into three contributions,
(56)
where T is given in Eq. (38), N collects contributions of Eqs. (14) and (15) that are not affected by possible resonances;
(57)
(58)
and S includes potentially resonant contributions that deserve particular attention,
(59)
(60)
As already mentioned, the most accurate treatment of these contributions is offered by the GVPT2 model, in which these terms are neglected in the perturbative treatment and reintroduced in a successive variational step.17,20 However, this approach is not well suited for thermochemical applications, as the selection of resonant terms can be different for different methods and/or different stationary points. At the same time, the accuracy required for thermochemical applications is lower than that needed for high-resolution spectroscopy. In this framework, Kuhler et al.92 proposed a general approach based on the systematic replacement of all potentially resonant contributions with non-resonant ones. In particular, using an analogy with the simplified Hamiltonian,
(61)
whose eigenvalues are
(62)
the authors showed that it is possible to rewrite the potentially resonant terms using the following transformation:
(63)
where ɛ corresponds to half the absolute value of the frequency difference, which may cause the singularity, and k2 is the absolute value of the constant term and S is the sign (i.e., 1 or −1). Application of Eq. (63) to Eq. (59) and (60) leads to a version known as degeneracy corrected VPT2 (DCPT2) for matrix S,
(64)
where
(65)
(66)
has been introduced for the sake of readability. The similarity of the matrix χ for both formulations of VPT2 has allowed a straightforward extension of the DCPT2 scheme to the use of curvilinear coordinates replacing the anharmonic force constants with the elements of the tensors η, σ, and ρ. Within the DCPT2 scheme, the fundamental energies can be written as follows:
(67)

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.

TABLE I.

Applicability of the VPT2 and DCPT2 models in the function of the values of ɛ and k.

VPT2DCPT2
k ≪ 1k ≫ 1k ≪ 1k ≫ 1
ɛ ≪ 1     
ɛ ≫ 1     
VPT2DCPT2
k ≪ 1k ≫ 1k ≪ 1k ≫ 1
ɛ ≪ 1     
ɛ ≫ 1     
Based on the previous discussion, an improved approach can be based on a smooth transition from VPT2 to DCPT2, depending on the nature of the resonant term. This consideration led to the development of the hybrid degeneracy-corrected second-order perturbation theory (HDCPT2),93 in which each quantity f is written in the following parametric form:
(68)
where fH, f, and fD represent the term within HDCPT2, VPT2, and DCPT2 frameworks, respectively.
The parameter Λ depends not only on k and ɛ but also on two additional parameters α and β,
(69)
with the former controlling the transition from VPT2 to DCPT2 and the latter its smoothness. The values of the parameters proposed in Ref. 93 were α = 1.0 and β = 5.0 × 105.
Then, it is possible to apply the transformation introduced in Eq. (68) to the elements of the matrix S,
(70)
In analogy to Eq. (67), the fundamental frequencies within the HDCPT2 framework can be evaluated,
(71)
and used in conjunction with Eq. (55) for the calculation of the vibrational partition function deprived of singularities. The latter serves as a starting point to derive the thermochemical properties of interest, such as the vibrational contributions to the internal energy (Uv), entropy (Sv), and specific heat (cv) accounting for anharmonic effects;
(72)

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:

  1. Geometry optimization and calculation of the force constant matrix through a selected QC program.

  2. 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.

  3. Extraction of the output data, construction of the anharmonic force constants and derivatives of the matrix g through the standalone program.

  4. 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.

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 

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.

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).

FIG. 1.

Molecular structures of (a) formaldehyde, (b) oxirane, (c) vinyl radical, (d) fluoroamine, and (e) imidazole.

FIG. 1.

Molecular structures of (a) formaldehyde, (b) oxirane, (c) vinyl radical, (d) fluoroamine, and (e) imidazole.

Close modal

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.

TABLE II.

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-RVPT2-C
Harm.aAnh.bW+CcTotalAnh.dExPot.eTotal
Formaldehyde 5 874 −80 −1 5 793 −77 −4 5 793 
Oxirane 12 637 −172 12 466 −185 14 12 466 
Imidazole 15 620 −189 15 432 −192 15 432 
Fluoroamine 6 058 −101 −1 5 957 −93 −8 5 957 
Vinyl radical 8 041 −123 7 921 −106 −14 7 921 
VPT2-RVPT2-C
Harm.aAnh.bW+CcTotalAnh.dExPot.eTotal
Formaldehyde 5 874 −80 −1 5 793 −77 −4 5 793 
Oxirane 12 637 −172 12 466 −185 14 12 466 
Imidazole 15 620 −189 15 432 −192 15 432 
Fluoroamine 6 058 −101 −1 5 957 −93 −8 5 957 
Vinyl radical 8 041 −123 7 921 −106 −14 7 921 
a

Harmonic ZPVE.

b

Anharmonic correction arising from the expansion of the potential energy.

c

Anharmonic correction arising from both the Coriolis effect and Watson’s terms.

d

Anharmonic correction arising from the expansion of both the potential and kinetic energies.

e

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.

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 stateSymm.VPT2DCPT2HDCPT2VPT2DCPT2HDCPT2CCaExpt.b
ZPVE578557855785578557855785
|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 
cpe  8.45 8.45 8.45 8.45 8.45 8.45 ⋯ 8.46d 
Curvilinear coords.Rectilinear coords.
Rectilinear stateSymm.VPT2DCPT2HDCPT2VPT2DCPT2HDCPT2CCaExpt.b
ZPVE578557855785578557855785
|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 
cpe  8.45 8.45 8.45 8.45 8.45 8.45 ⋯ 8.46d 
a

GVPT2-R with CCSD(T)/aug-cc-pVTZ force field from Ref. 108.

b

Reference 109.

c

Molar entropy at 298.15 K and 1 atm.

d

Reference 110.

e

Specific heat at 298.15 K and 1 atm.

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)].

FIG. 2.

Molecular structures of (a) formic acid, (b) methanol, and (c) acetaldehyde.

FIG. 2.

Molecular structures of (a) formic acid, (b) methanol, and (c) acetaldehyde.

Close modal

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.

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.
AssignmentSymm.Harm.HDCPT2HDCPT2GVPT2aExpt.b
ZPVE7443734373437343
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 ⋯ 
Curvilinear coords.Rectilinear coords.
AssignmentSymm.Harm.HDCPT2HDCPT2GVPT2aExpt.b
ZPVE7443734373437343
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 ⋯ 
a

Anharmonic calculation based on the GVPT2 scheme accounting for both FRs and DDRs effects.

b

References 112 and 113.

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.

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.HDCPT2HDCPT2GVPT2aExpt.b
ZPVE11 32411 15211 15211 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 ⋯ 
Curvilinear coords.Rectilinear coords.
Ass.Symm.Harm.HDCPT2HDCPT2GVPT2aExpt.b
ZPVE11 32411 15211 15211 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 ⋯ 
a

Anharmonic calculation based on the GVPT2 scheme accounting for both FRs and DDRs effects.

b

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).

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.

MoleculeHarm.aVPT2bHDCPT2cExpt.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)  
MoleculeHarm.aVPT2bHDCPT2cExpt.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)  
a

Harmonic value of the thermochemical property, with the vibrational contribution being evaluated within the Wilson GF method.

b

Curvilinear-based formulation of VPT2.

c

Curvilinear-based formulation of HDCPT2.

d

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.

e

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.

FIG. 3.

Molecular structure of formamide.

FIG. 3.

Molecular structure of formamide.

Close modal

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.

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.
AssignmentSymm.Harm.VPT2HDCPT2HDCPT2GVPT2aExpt.b
ZPVE99869936993699369936
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  
MAEe  68 57 34 29 29  
MAEf  (29) (14) (14) (14) (57)  
Curvilinear coords.Rectilinear coords.
AssignmentSymm.Harm.VPT2HDCPT2HDCPT2GVPT2aExpt.b
ZPVE99869936993699369936
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  
MAEe  68 57 34 29 29  
MAEf  (29) (14) (14) (14) (57)  
a

Anharmonic calculation based on the GVPT2 scheme accounting for both FRs and DDRs effects.

b

Reference 115.

c

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.

d

Mean absolute error of the theoretical wave numbers, with the fundamental transition of the out-of-plane motion of the NH2 group being neglected.

e

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.

f

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.

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.

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.

Close modal

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.

FIG. 5.

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.

FIG. 5.

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.

Close modal

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.

The analysis presented in this section has been limited to relatively small molecular systems. When the focus shifts to large systems, an additional issue comes into play, namely, the problem of determining accurate vibrational properties while retaining an affordable computational cost. An effective solution is offered by the so-called dual-level method1,20 in which the harmonic terms and the corresponding anharmonic corrections are evaluated at different levels of theory, since the accuracy of the former (generally much larger) contributions iscule usually more critical. Furthermore, the standard finite-difference approach employed for the construction of the quartic force field requires the computation of 2N′ + 1 analytical force constant matrices, whose cost increases steeply with molecular size, becoming rapidly prohibitive even for double hybrid functionals. Within the dual-level framework, the value of a generic vibrational property P is
(73)
where PH is the harmonic term, ΔPA is the anharmonic correction, while HL and LL refer to generic high- and low-level methods, respectively. In order to clearly understand the impact of the harmonic contribution, it is convenient to rewrite Eq. (73) in the following way:
(74)
where ΔPH=PHLHPLLH is the corrective factor that leads from the LL value of the harmonic term to its HL counterpart. In the present work, rDSD/3F12 and B3/SVP (or alternatively, B3/6-31G*) have been used systematically as the HL and LL methods, respectively.

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.

FIG. 6.

Molecular structures of (a) Tc conformer of pyruvic acid and (b) norbornane.

FIG. 6.

Molecular structures of (a) Tc conformer of pyruvic acid and (b) norbornane.

Close modal

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 10 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.

TABLE VIII.

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.

StateSymm.Harm.aΔHarm.bΔAnharm.cAnharm.dExpt.e
ZPVE39 092−98−132137 673
|11⟩ A1 3113 −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 −23 998 989 
|111⟩  935 −19 925 916 
|112⟩  890 −19 874 864 
|113⟩  828 −16 814 818 
|114⟩  765 −15 754 758 
|115⟩  406 −3 −9 394 401 
|116⟩ A2 3086 −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 −19 963 956 
|124⟩  967 −4 −27 936 949 
|125⟩  548 −3 −11 534 546 
|126⟩  158 −13 −8 137 152 
|127⟩ B1 3108 −149 2961 ⋯ 
|128⟩  3102 −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 −27 1024 1027 
|135⟩  972 −23 954 957 
|136⟩  905 −21 893 877 
|137⟩  811 −7 −19 785 786 
|138⟩  453 −6 449 435 
|139⟩ B2 3103 −2 −137 2964 ⋯ 
|140⟩  3090 −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 −17 958 956 
|149⟩  829 −18 819 814 
|150⟩  769 −19 755 760 
|151⟩  337 −5 −16 316 336 
MAEf  31 ⋯ ⋯ ⋯ 
StateSymm.Harm.aΔHarm.bΔAnharm.cAnharm.dExpt.e
ZPVE39 092−98−132137 673
|11⟩ A1 3113 −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 −23 998 989 
|111⟩  935 −19 925 916 
|112⟩  890 −19 874 864 
|113⟩  828 −16 814 818 
|114⟩  765 −15 754 758 
|115⟩  406 −3 −9 394 401 
|116⟩ A2 3086 −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 −19 963 956 
|124⟩  967 −4 −27 936 949 
|125⟩  548 −3 −11 534 546 
|126⟩  158 −13 −8 137 152 
|127⟩ B1 3108 −149 2961 ⋯ 
|128⟩  3102 −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 −27 1024 1027 
|135⟩  972 −23 954 957 
|136⟩  905 −21 893 877 
|137⟩  811 −7 −19 785 786 
|138⟩  453 −6 449 435 
|139⟩ B2 3103 −2 −137 2964 ⋯ 
|140⟩  3090 −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 −17 958 956 
|149⟩  829 −18 819 814 
|150⟩  769 −19 755 760 
|151⟩  337 −5 −16 316 336 
MAEf  31 ⋯ ⋯ ⋯ 
a

B3/6-31G* level of theory.

b

Difference between the harmonic contributions at the rDSD/3F12 level and those at the B3/6-31G* level.

c

Anharmonic corrections at the B3/6-31G* level.

d

Fully anharmonic results.

e

Reference 121.

f

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.

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.

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.

Close modal

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).

TABLE IX.

Comparison of the experimental and theoretical fundamental wave numbers (in cm−1) of the Tc conformer of pyruvic acid.

ωaδνbνc
Ass.Symm.B3drDSDeCCfB3dB3drDSD//B3gCC//B3hrDSDeexp.i
ZPVE15 67415 74315 743−19115 48315 55215 55215 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 ⋯ 
ωaδνbνc
Ass.Symm.B3drDSDeCCfB3dB3drDSD//B3gCC//B3hrDSDeexp.i
ZPVE15 67415 74315 743−19115 48315 55215 55215 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 ⋯ 
a

Harmonic result.

b

Anharmonic correction.

c

Full anharmonic result.

d

B3/SVP level of theory.

e

rDSD/3F12 level of theory.

f

CCSD(T)/cc-pVTZ level of theory from Ref. 118.

g

Anharmonic fundamental wave numbers obtained by combining rDSD/3F12 harmonic frequencies with B3/SVP anharmonic corrections.

h

Anharmonic fundamental wave numbers obtained by combining CCSD(T)/cc-pVTZ harmonic frequencies with B3/SVP anharmonic corrections.

i

References 119 and 120.

j

Correction computed through the HR model.

k

Fundamental wave number computed through the HR model.

l

Mean absolute error of the computed anharmonic wave numbers with respect to their experimental counterparts.

TABLE X.

Comparison of harmonic and anharmonic ZPVE (in kcal mol−1) for medium-sized molecules.

MoleculeHarm.ΔAnharm
B3arDSDbB3 scal.cB3a
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 
MoleculeHarm.ΔAnharm
B3arDSDbB3 scal.cB3a
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 
a

B3/SVP level of theory.

b

rDSD/3F12 level of theory.

c

B3/SVP value scaled by a constant factor of 0.996.

d

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.

TABLE XI.

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.

PropertyHarm.aΔAnharm.bAnharm.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 
PropertyHarm.aΔAnharm.bAnharm.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 
a

Harmonic value of the thermochemical property, with the vibrational contribution being evaluated within the Wilson GF method.

b

Anharmonic correction to the thermochemical property.

c

Curvilinear-based formulation of HDCPT2.

d

In parenthesis: harmonic values scaled by 0.996.

FIG. 8.

Molecular structure of (a) androsterone, (b) β-estradiol, (c) molecular motor, and (d) uridine.

FIG. 8.

Molecular structure of (a) androsterone, (b) β-estradiol, (c) molecular motor, and (d) uridine.

Close modal

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.

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.

The supplementary material includes the definition of the anharmonic χ matrix within the different frameworks of VPT2.

The authors acknowledge the funding from Gaussian, Inc.

The authors have no conflicts to disclose.

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).

The data that support the findings of this study are available within the article and its supplementary material.

Inspection of Eq. (30) shows that the extra-potential term (Vg) depends on first- and second-order derivatives of the determinant g̃ (or the g̃ij 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 (gij=giiδij). Therefore, it is possible to restrict the analysis to the coefficients g̃i and g̃ii.

One possible approach for calculating the desired terms relies entirely on the numerical differentiation of analytical determinants,
(A1)
(A2)
where δqi is the displacement along the normal mode i. Alternatively, the coefficients g̃ij can be related to the derivatives of the matrix g. In particular,
(A3)
where N′ and Nr are the numbers of active vibrations and of rotational degrees of freedom, respectively. Similarly, the second-order terms can be obtained using the following expression:
(A4)

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.

1.
C.
Puzzarini
,
J.
Bloino
,
N.
Tasinato
, and
V.
Barone
, “
Accuracy and interpretability: The devil and the holy grail. New routes across old boundaries in computational spectroscopy
,”
Chem. Rev.
119
,
8131
8191
(
2019
).
2.
J. M. L.
Martin
and
M. K.
Kesharwani
, “
Assessment of CCSD(T)-F12 approximations and basis sets for harmonic vibrational frequencies
,”
J. Chem. Theory Comput.
10
,
2085
2090
(
2014
).
3.
D.
Agbaglo
and
R. C.
Fortenberry
, “
The performance of explicitly correlated methods for the computation of anharmonic vibrational frequencies
,”
Int. J. Quantum Chem.
119
,
e25899
(
2019
).
4.
Z. N.
Heim
,
B. K.
Amberger
,
B. J.
Esselman
,
J. F.
Stanton
,
R. C.
Woods
, and
R. J.
McMahon
, “
Molecular structure determination: Equilibrium structure of pyrimidine (m-C4H4N2) from rotational spectroscopy ((reSE)) and high-level ab initio calculation (re) agree within the uncertainty of experimental measurement
,”
J. Chem. Phys.
152
,
104303
(
2020
).
5.
K.
Raghavachari
,
G. W.
Trucks
,
J. A.
Pople
, and
M.
Head-Gordon
, “
A fifth-order perturbation comparison of electron correlation theories
,”
Chem. Phys. Lett.
157
,
479
483
(
1989
).
6.
Q.
Ma
and
H. J.
Werner
, “
Explicitly correlated local coupled-cluster methods using pair natural orbitals
,”
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
8
,
e1371
(
2018
).
7.
P. R.
Nagy
and
M.
Kállay
, “
Approaching the basis set limit of CCSD(T) energies for large molecules with local natural orbital coupled-cluster methods
,”
J. Chem. Theory Comput.
15
,
5275
5298
(
2019
).
8.
D. G.
Liakos
,
Y.
Guo
, and
F.
Neese
, “
Comprehensive benchmark results for the domain based local pair natural orbital coupled cluster method (DLPNO-CCSD(T)) for closed- and open-shell systems
,”
J. Phys. Chem. A
124
,
90
100
(
2020
).
9.
M.
Kállay
,
R. A.
Horváth
,
L.
Gyevi-Nagy
, and
P. R.
Nagy
, “
Basis set limit CCSD(T) energies for extended molecules via a reduced-cost explicitly correlated approach
,”
J. Chem. Theory Comput.
19
,
174
189
(
2023
).
10.
V.
Barone
,
G.
Ceselin
,
M.
Fusè
, and
N.
Tasinato
, “
Accuracy meets interpretability for computational spectroscopy by means of hybrid and double-hybrid functionals
,”
Front. Chem.
8
,
584203
(
2020
).
11.
Q.
Yang
,
M.
Mendolicchio
,
V.
Barone
, and
J.
Bloino
, “
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.
8
,
665232
(
2021
).
12.
H. H.
Nielsen
, “
The vibration-rotation energies of molecules
,”
Rev. Mod. Phys.
23
,
90
136
(
1951
).
13.
I. M.
Mills
, “
Vibration-rotation structure in asymmetric- and symmetric-top molecules
,” in
Molecular Spectroscopy: Modern Research
, edited by
K. N.
Rao
and
C. W.
Mathews
(
Academic Press
,
New York
,
1972
), Chap. 3.2, pp.
115
140
.
14.
D. A.
Clabo
, Jr.
,
W. D.
Allen
,
R. B.
Remington
,
Y.
Yamaguchi
, and
H. F.
Schaefer
III
, “
A systematic study of molecular vibrational anharmonicity and vibration-rotation interaction by self-consistent-field higher-derivative methods. Asymmetric top molecules
,”
Chem. Phys.
123
,
187
239
(
1988
).
15.
W. D.
Allen
,
Y.
Yamaguchi
,
A. G.
Császár
,
D. A.
Clabo
, Jr.
,
R. B.
Remington
, and
H. F.
Schaefer
III
, “
A systematic study of molecular vibrational anharmonicity and vibration-rotation interaction by self-consistent-fied higher-derivative methods. Linear polyatomic molecules
,”
Chem. Phys.
145
,
427
466
(
1990
).
16.
V.
Barone
, “
Vibrational zero-point energies and thermodynamic functions beyond the harmonic approximation
,”
J. Chem. Phys.
120
,
3059
3065
(
2004
).
17.
V.
Barone
, “
Anharmonic vibrational properties by a fully automated second-order perturbative approach
,”
J. Chem. Phys.
122
,
014108
(
2005
).
18.
S. V.
Krasnoshchekov
,
E. V.
Isayeva
, and
N. F.
Stepanov
, “
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
116
,
3691
3709
(
2012
).
19.
A. M.
Rosnik
and
W. F.
Polik
, “
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.
112
,
261
300
(
2014
).
20.
P. R.
Franke
,
J. F.
Stanton
, and
G. E.
Douberly
, “
How to VPT2: Accurate and intuitive simulations of CH stretching infrared spectra using VPT2+K with large effective Hamiltonian resonance treatments
,”
J. Phys. Chem. A
125
,
1301
1324
(
2021
).
21.
C.
Peng
,
P. Y.
Ayala
,
H. B.
Schlegel
, and
M. J.
Frisch
, “
Using redundant internal coordinates to optimize equilibrium geometries and transition states
,”
J. Comput. Chem.
17
,
49
56
(
1996
).
22.
V.
Barone
, “
Quantum chemistry meets high-resolution spectroscopy for characterizing the molecular bricks of life in the gas-phase
,”
Phys. Chem. Chem. Phys.
26
,
5802
5821
(
2024
).
23.
F.
Lazzari
,
A.
Salvadori
,
G.
Mancini
, and
V.
Barone
, “
Molecular perception for visualization and computation: The proxima library
,”
J. Chem. Inf. Model.
60
,
2668
2672
(
2020
).
24.
C.
Bannwarth
,
S.
Ehlert
, and
S.
Grimme
, “
GFN2-xTB—An accurate and broadly parametrized self-consistent tight-binding quantum chemical method with multipole electrostatics and density-dependent dispersion contributions
,”
J. Chem. Theory Comput.
15
,
1652
1671
(
2019
).
25.
P.
Pracht
,
F.
Bohle
, and
S.
Grimme
, “
Automated exploration of the low-energy chemical space with fast quantum chemical methods
,”
Phys. Chem. Chem. Phys.
22
,
7169
7192
(
2020
).
26.
G.
Mancini
,
M.
Fusè
,
F.
Lazzari
,
B.
Chandramouli
, and
V.
Barone
, “
Unsupervised search of low-lying conformers with spectroscopic accuracy: A two-step algorithm rooted into the island model evolutionary algorithm
,”
J. Chem. Phys.
153
,
124110
(
2020
).
27.
D.
Ferro-Costas
,
I.
Mosquera-Lois
, and
A.
Fernandez-Ramos
, “
TorsiFlex: An automatic generator of torsional conformers. Application to the twenty proteinogenic amino acids
,”
J. Cheminf.
13
,
100
(
2021
).
28.
G.
Mancini
,
M.
Fusè
,
F.
Lazzari
, and
V.
Barone
, “
Fast exploration of potential energy surfaces with a joint venture of quantum chemistry, evolutionary algorithms and unsupervised learning
,”
Digital Discovery
1
,
790
805
(
2022
).
29.
V.
Barone
, “
Accuracy meets feasibility for the structures and rotational constants of the molecular bricks of life: A joint venture of DFT and wave-function methods
,”
J. Phys. Chem. Lett.
14
,
5883
5890
(
2023
).
30.
V.
Barone
, “
PCS/Bonds and PCS0: Pick your molecule and get its accurate structure and ground state rotational constants at DFT cost
,”
J. Chem. Phys.
159
,
081102
(
2023
).
31.
S.
Di Grande
and
V.
Barone
, “
Toward accurate quantum chemical methods for molecules of increasing dimension: The new family of Pisa composite schemes
,”
J. Phys. Chem. A
128
,
4886
4900
(
2024
).
32.
D. G.
Truhlar
and
A. D.
Isaacson
, “
Simple perturbation theory estimates of equilibrium constants from force fields
,”
J. Chem. Phys.
94
,
357
359
(
1991
).
33.
E.
Dzib
and
G.
Merino
, “
The hindered rotor theory: A review
,”
Wiley Interdiscip. Rev.:Comput. Mol. Sci.
12
,
e1583
(
2022
).
34.
V.
Barone
, “
From perception to prediction and interpretation: Enlightening the gray zone of molecular bricks of life with the help of machine learning and quantum chemistry
,”
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
15
,
e70000
(
2025
).
35.
M.
Mendolicchio
,
J.
Bloino
, and
V.
Barone
, “
Perturb-then-diagonalize vibrational engine exploiting curvilinear internal coordinates
,”
J. Chem. Theory Comput.
18
,
7603
7619
(
2022
).
36.
M.
Mendolicchio
, “
Harnessing the power of curvilinear internal coordinates: From molecular structure prediction to vibrational spectroscopy
,”
Theor. Chem. Acc.
142
,
133
(
2023
).
37.
M.
Mendolicchio
and
V.
Barone
, “
Accurate vibrational and ro-vibrational contributions to the properties of large molecules by a new engine employing curvilinear internal coordinates and vibrational perturbation theory to second order
,”
J. Chem. Theory Comput.
20
,
8378
8395
(
2024
).
38.
P. Y.
Ayala
and
H. B.
Schlegel
, “
Identification and treatment of internal rotation in normal mode vibrational analysis
,”
J. Chem. Phys.
108
,
2314
2325
(
1998
).
39.
V.
Barone
,
J.
Lupi
,
Z.
Salta
, and
N.
Tasinato
, “
Development and validation of a parameter-free model chemistry for the computation of reliable reaction rates
,”
J. Chem. Theory Comput.
17
,
4913
4928
(
2021
).
40.
M. A.
Harthcock
and
J.
Laane
, “
Calculation of two-dimensional vibrational potential energy surfaces utilizing prediagonalized basis sets and Van Vleck perturbation methods
,”
J. Phys. Chem.
89
,
4231
4240
(
1985
).
41.
V.
Barone
,
M.
Biczysko
,
J.
Bloino
,
M.
Borkowska-Panek
,
I.
Carnimeo
, and
P.
Panek
, “
Toward anharmonic computations of vibrational spectra for large molecular systems
,”
Int. J. Quantum Chem.
112
,
2185
2200
(
2012
).
42.
M.
Fusè
,
G.
Mazzeo
,
G.
Longhi
,
S.
Abbate
,
Q.
Yang
, and
J.
Bloino
, “
Scaling-up VPT2: A feasible route to include anharmonic correction on large molecules
,”
Spectrochim. Acta, Part A
311
,
123969
(
2024
).
43.
T.
Carrington
, Jr.
, “
Using collocation to study the vibrational dynamics of molecules
,”
Spectrochim. Acta, Part A
248
,
119158
(
2021
).
44.
S.
Manzhos
,
X.
Wang
, and
T.
Carrington
, Jr.
, “
A multimode-like scheme for selecting the centers of Gaussian basis functions when computing vibrational spectra
,”
Chem. Phys.
509
,
139
144
(
2018
).
45.
A. G.
Császár
,
C.
Fábri
,
T.
Szidarovszky
,
E.
Mátyus
,
T.
Furtenbacher
, and
G.
Czakó
, “
The fourth age of quantum chemistry: Molecules in motion
,”
Phys. Chem. Chem. Phys.
14
,
1085
1106
(
2012
).
46.
E.
Mátyus
,
G.
Czakó
, and
A. G.
Császár
, “
Toward black-box-type full- and reduced-dimensional variational (ro)vibrational computations
,”
J. Chem. Phys.
130
,
134112
(
2009
).
47.
D.
Lauvergnat
and
A.
Nauts
, “
Exact numerical computation of a kinetic energy operator in curvilinear coordinates
,”
J. Chem. Phys.
116
,
8560
8570
(
2002
).
48.
D.
Lauvergnat
,
E.
Baloïtcha
,
G.
Dive
, and
M.
Desouter-Lecomte
, “
Dynamics of complex molecular systems with numerical kinetic energy operators in generalized coordinates
,”
Chem. Phys.
326
,
500
508
(
2006
).
49.
Y.
Scribano
,
D. M.
Lauvergnat
, and
D. M.
Benoit
, “
Fast vibrational configuration interaction using generalized curvilinear coordinates and self-consistent basis
,”
J. Chem. Phys.
133
,
094103
(
2010
).
50.
J.
Tennyson
, “
Perspective: Accurate ro-vibrational calculations on small molecules
,”
J. Chem. Phys.
145
,
124112
(
2016
).
51.
B. T.
Sutcliffe
and
J.
Tennyson
, “
A general treatment of vibration-rotation coordinates for triatomic molecules
,”
Int. J. Quantum Chem.
39
,
183
196
(
1991
).
52.
S. N.
Yurchenko
,
W.
Thiel
, and
P.
Jensen
, “
Theoretical ROVibrational Energies (TROVE): A robust numerical approach to the calculation of rovibrational energies for polyatomic molecules
,”
J. Mol. Spectrosc.
245
,
126
140
(
2007
).
53.
A. S.
Petit
and
A. B.
McCoy
, “
Diffusion Monte Carlo in internal coordinates
,”
J. Phys. Chem. A
117
,
7009
7018
(
2013
).
54.
I. W.
Bulik
,
M. J.
Frisch
, and
P. H.
Vaccaro
, “
Fixed-node, importance-sampling diffusion Monte Carlo for vibrational structure with accurate and compact trial states
,”
J. Chem. Theory Comput.
14
,
1554
1563
(
2018
).
55.
K.
Yagi
,
M.
Keçeli
, and
S.
Hirata
, “
Optimized coordinates for anharmonic vibrational structure theories
,”
J. Chem. Phys.
137
,
204118
(
2012
).
56.
P. M.
Zimmerman
and
P.
Smereka
, “
Optimizing vibrational coordinates to modulate intermode coupling
,”
J. Chem. Theory Comput.
12
,
1883
1891
(
2016
).
57.
B.
Thomsen
,
K.
Yagi
, and
O.
Christiansen
, “
Optimized coordinates in vibrational coupled cluster calculations
,”
J. Chem. Phys.
140
,
154102
(
2014
).
58.
P. S. T.
Arnaud Leclerc
and
T.
Carrington
, “
Comparison of different eigensolvers for calculating vibrational spectra using low-rank, sum-of-product basis functions
,”
Mol. Phys.
115
,
1740
1749
(
2017
).
59.
S. R.
White
, “
Density matrix formulation for quantum renormalization groups
,”
Phys. Rev. Lett.
69
,
2863
2866
(
1992
).
60.
U.
Schollwöck
, “
The density-matrix renormalization group in the age of matrix product states
,”
Ann. Phys.
326
,
96
192
(
2011
).
61.
A.
Baiardi
and
M.
Reiher
, “
The density matrix renormalization group in chemistry and molecular physics: Recent developments and new challenges
,”
J. Chem. Phys.
152
,
040903
(
2020
).
62.
A.
Baiardi
,
C. J.
Stein
,
V.
Barone
, and
M.
Reiher
, “
Vibrational density matrix renormalization group
,”
J. Chem. Theory Comput.
13
,
3764
3777
(
2017
).
63.
N.
Glaser
,
A.
Baiardi
, and
M.
Reiher
, “
Flexible DMRG-based framework for anharmonic vibrational calculations
,”
J. Chem. Theory Comput.
19
,
9329
9343
(
2023
).
64.
M.
Zou
,
Y.
Hassan
,
T. K.
Roy
,
A. B.
McCoy
, and
M. I.
Lester
, “
Infrared spectroscopy of the syn-methyl-substituted criegee intermediate: A combined experimental and theoretical study
,”
J. Chem. Phys.
160
,
204309
(
2024
).
65.
A. B.
McCoy
and
M. A.
Boyer
, “
Exploring expansions of the potential and dipole surfaces used for vibrational perturbation theory
,”
J. Phys. Chem. A
126
,
7242
7249
(
2022
).
66.
S.
Carter
,
A. R.
Sharma
,
J. M.
Bowman
,
P.
Rosmus
, and
R.
Tarroni
, “
Calculations of rovibrational energies and dipole transition intensities for polyatomic molecules using multimode
,”
J. Chem. Phys.
131
,
224106
(
2009
).
67.
G.
Rauhut
and
T.
Hrenar
, “
A combined variational and perturbational study on the vibrational spectrum of P2F4
,”
Chem. Phys.
346
,
160
166
(
2008
).
68.
O.
Christiansen
, “
Selected new developments in vibrational structure theory: Potential construction and vibrational wave function calculations
,”
Phys. Chem. Chem. Phys.
14
,
6672
6687
(
2012
).
69.
D.
Papp
,
T.
Szidarovszky
, and
A. G.
Császár
, “
A general variational approach for computing rovibrational resonances of polyatomic molecules. application to the weakly bound H2He+ and H2·CO systems
,”
J. Chem. Phys.
147
,
094106
(
2017
).
70.
A.
Erba
,
J.
Maul
,
M.
Ferrabone
,
P.
Carbonnière
,
M.
Rérat
, and
R.
Dovesi
, “
Anharmonic vibrational states of solids from DFT calculations. Part I: Description of the potential energy surface
,”
J. Chem. Theory Comput.
15
,
3755
3765
(
2019
).
71.
A.
Erba
,
J.
Maul
,
M.
Ferrabone
,
R.
Dovesi
,
M.
Rérat
, and
P.
Carbonnière
, “
Anharmonic vibrational states of solids from DFT calculations. Part II: Implementation of the VSCF and VCI methods
,”
J. Chem. Theory Comput.
15
,
3766
3777
(
2019
).
72.
D.
Mitoli
,
J.
Maul
, and
A.
Erba
, “
Anharmonic terms of the potential energy surface: A group theoretical approach
,”
Cryst. Growth Des.
23
,
3671
3680
(
2023
).
73.
R. G.
Schireman
,
J.
Maul
,
A.
Erba
, and
M. T.
Ruggiero
, “
Anharmonic coupling of stretching vibrations in ice: A periodic VSCF and VCI description
,”
J. Chem. Theory Comput.
18
,
4428
4437
(
2022
).
74.
M.
Born
and
R.
Oppenheimer
, “
Zur quantentheorie der molekeln
,”
Ann. Phys.
389
,
457
484
(
1927
).
75.
E.
Mátyus
,
A. M.
Santa Daría
, and
G.
Avila
, “
Exact quantum dynamics developments for floppy molecular systems and complexes
,”
Chem. Commun.
59
,
366
381
(
2023
).
76.
S.
Yurchenko
,
Computational Spectroscopy of Polyatomic Molecules
(
CRC Press
,
2023
).
77.
R.
Meyer
and
H. H.
Gunthard
, “
General internal motion of molecules, classical and quantum-mechanical Hamiltonian
,”
J. Chem. Phys.
49
,
1510
1520
(
1968
).
78.
H. M.
Pickett
, “
Vibration—rotation interactions and the choice of rotating axes for polyatomic molecules
,”
J. Chem. Phys.
56
,
1715
1723
(
1972
).
79.
B.
Podolsky
, “
Quantum-mechanically correct form of Hamiltonian function for conservative systems
,”
Phys. Rev.
32
,
812
(
1928
).
80.
E. B.
Wilson
, Jr.
, “
Some mathematical methods for the study of molecular vibrations
,”
J. Chem. Phys.
9
,
76
84
(
1941
).
81.
S.
Califano
,
Vibrational States
(
John Wiley and Sons
,
1976
).
82.
J. H.
Van Vleck
, “
On σ-type doubling and electron spin in the spectra of diatomic molecules
,”
Phys. Rev.
33
,
467
(
1929
).
83.
S. V.
Krasnoshchekov
,
E. V.
Isayeva
, and
N. F.
Stepanov
, “
Criteria for first- and second-order vibrational resonances and correct evaluation of the Darling-Dennison resonance coefficients using the canonical Van Vleck perturbation theory
,”
J. Chem. Phys.
141
,
234114
(
2014
).
84.
C. R.
Quade
, “
Internal coordinate formulation for the vibration–rotation energies of polyatomic molecules
,”
J. Chem. Phys.
64
,
2783
2795
(
1976
).
85.
A. D.
Isaacson
, “
Including anharmonicity in the calculation of rate constants. 1. The HCN/HNC isomerization reaction
,”
J. Phys. Chem. A
110
,
379
388
(
2006
).
86.
A. D.
Isaacson
, “
Including anharmonicity in the calculation of rate constants. II. The OH + H2 → H2O + H reaction
,”
J. Chem. Phys.
128
,
134304
(
2008
).
87.
E. L.
Sibert
, “
Theoretical studies of vibrationally excited polyatomic molecules using canonical Van Vleck perturbation theory
,”
J. Chem. Phys.
88
,
4378
4390
(
1988
).
88.
E. L.
Sibert
, “
Modeling vibrational anharmonicity in infrared spectra of high frequency vibrations of polyatomic molecules
,”
J. Chem. Phys.
150
,
054107
(
2019
).
89.
M. A.
Boyer
and
A. B.
McCoy
, “
A flexible approach to vibrational perturbation theory using sparse matrix methods
,”
J. Chem. Phys.
156
,
054107
(
2022
).
90.
J. M. L.
Martin
,
T. J.
Lee
,
P. R.
Taylor
, and
J.-P.
François
, “
The anharmonic force field of ethylene, C2H4, by means of accurate ab initio calculations
,”
J. Chem. Phys.
103
,
2589
2602
(
1995
).
91.
S. V.
Krasnoshchekov
,
E. O.
Dobrolyubov
,
M. A.
Syzgantseva
, and
R. V.
Palvelev
, “
Rigorous vibrational fermi resonance criterion revealed: Two different approaches yield the same result
,”
Mol. Phys.
118
,
e1743887
(
2020
).
92.
K. M.
Kuhler
,
D. G.
Truhlar
, and
A. D.
Isaacson
, “
General method for removing resonance singularities in quantum mechanical perturbation theory
,”
J. Chem. Phys.
104
,
4664
4671
(
1996
).
93.
J.
Bloino
,
M.
Biczysko
, and
V.
Barone
, “
General perturbative approach for spectroscopy, thermodynamics, and kinetics: Methodological background and benchmark studies
,”
J. Chem. Theory Comput.
8
,
1015
1036
(
2012
).
94.
M.
Mendolicchio
,
J.
Bloino
, and
V.
Barone
, “
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.
17
,
4332
4358
(
2021
).
95.
M. S.
Schuurman
,
W. D.
Allen
,
P.
von Ragué Schleyer
, and
H. F.
Schaefer
III
, “
The highly anharmonic BH5 potential energy surface characterized in the ab initio limit
,”
J. Chem. Phys.
122
,
104302
(
2005
).
96.
M. J.
Frisch
,
G. W.
Trucks
,
H. B.
Schlegel
,
G. E.
Scuseria
,
M. A.
Robb
,
J. R.
Cheeseman
,
G.
Scalmani
,
V.
Barone
,
G. A.
Petersson
,
H.
Nakatsuji
,
X.
Li
,
M.
Caricato
,
A. V.
Marenich
,
J.
Bloino
,
B. G.
Janesko
,
R.
Gomperts
,
B.
Mennucci
,
H. P.
Hratchian
,
J. V.
Ortiz
,
A. F.
Izmaylov
,
J. L.
Sonnenberg
,
D.
Williams-Young
,
F.
Ding
,
F.
Lipparini
,
F.
Egidi
,
J.
Goings
,
B.
Peng
,
A.
Petrone
,
T.
Henderson
,
D.
Ranasinghe
,
V. G.
Zakrzewski
,
J.
Gao
,
N.
Rega
,
G.
Zheng
,
W.
Liang
,
M.
Hada
,
M.
Ehara
,
K.
Toyota
,
R.
Fukuda
,
J.
Hasegawa
,
M.
Ishida
,
T.
Nakajima
,
Y.
Honda
,
O.
Kitao
,
H.
Nakai
,
T.
Vreven
,
K.
Throssell
,
J. A.
Montgomery
, Jr.
,
J. E.
Peralta
,
F.
Ogliaro
,
M. J.
Bearpark
,
J. J.
Heyd
,
E. N.
Brothers
,
K. N.
Kudin
,
V. N.
Staroverov
,
T. A.
Keith
,
R.
Kobayashi
,
J.
Normand
,
K.
Raghavachari
,
A. P.
Rendell
,
J. C.
Burant
,
S. S.
Iyengar
,
J.
Tomasi
,
M.
Cossi
,
J. M.
Millam
,
M.
Klene
,
C.
Adamo
,
R.
Cammi
,
J. W.
Ochterski
,
R. L.
Martin
,
K.
Morokuma
,
O.
Farkas
,
J. B.
Foresman
, and
D. J.
Fox
, Gaussian 16, Revision C.01,
Gaussian, Inc.
,
Wallingford, CT
,
2016
.
97.
A. D.
Becke
, “
Density-functional exchange-energy approximation with correct asymptotic behavior
,”
Phys. Rev. A
38
,
3098
3100
(
1988
).
98.
C.
Lee
,
W.
Yang
, and
R. G.
Parr
, “
Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density
,”
Phys. Rev. B
37
,
785
789
(
1988
).
99.
A. D.
Becke
, “
Density-functional thermochemistry. III. The role of exact exchange
,”
J. Chem. Phys.
98
,
5648
5652
(
1993
).
100.
G.
Santra
,
N.
Sylvetsky
, and
J. M. L.
Martin
, “
Minimally empirical double-hybrid functionals trained against the GMTKN55 database: revDSD-PBEP86-D4, revDOD-PBE-D4, and DOD-SCAN-D4
,”
J. Phys. Chem. A
123
,
5129
5143
(
2019
).
101.
S.
Grimme
, “
Density functional theory with London dispersion corrections
,”
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
1
,
211
228
(
2011
).
102.
E.
Caldeweyher
,
C.
Bannwarth
, and
S.
Grimme
, “
Extension of the D3 dispersion coefficient model
,”
J. Chem. Phys.
147
,
034112
(
2017
).
103.
W. J.
Hehre
,
R.
Ditchfield
, and
J. A.
Pople
, “
Self—consistent molecular orbital methods. XII. Further extensions of Gaussian—type basis sets for use in molecular orbital studies of organic molecules
,”
J. Chem. Phys.
56
,
2257
2261
(
1972
).
104.
P. C.
Hariharan
and
J. A.
Pople
, “
The influence of polarization functions on molecular orbital hydrogenation energies
,”
Theor. Chim. Acta
28
,
213
222
(
1973
).
105.
T.
Clark
,
J.
Chandrasekhar
,
G. W.
Spitznagel
, and
P. V. R.
Schleyer
, “
Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21+G basis set for first-row elements, Li–F
,”
J. Comput. Chem.
4
,
294
301
(
1983
).
106.
K. A.
Peterson
,
T. B.
Adler
, and
H.-J.
Werner
, “
Systematically convergent basis sets for explicitly correlated wavefunctions: The atoms H, He, B–Ne, and Al–Ar
,”
J. Chem. Phys.
128
,
084102
(
2008
).
107.
T. H.
Dunning
, “
Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen
,”
J. Chem. Phys.
90
,
1007
1023
(
1989
).
108.
G.
Rauhut
,
G.
Knizia
, and
H. J.
Werner
, “
Accurate calculation of vibrational frequencies using explicitly correlated coupled-cluster theory
,”
J. Chem. Phys.
130
,
054105
(
2009
).
109.
T.
Nakanaga
,
S.
Kondo
, and
S.
Saëki
, “
Infrared band intensities of formaldehyde and formaldehyde-d2
,”
J. Chem. Phys.
76
,
3860
3865
(
1982
).
110.
L. V.
Gurvich
,
I. V.
Veyts
, and
C. B.
Alcock
,
Thermodynamic Properties of Individual Substances
(
Hemisphere Publishing Co.
,
New York
,
1989
).
111.
B. T.
Darling
and
D. M.
Dennison
, “
The water vapor molecule
,”
Phys. Rev.
57
,
128
139
(
1940
).
112.
R. L.
Redington
, “
Vibrational spectra and normal coordinate analysis of isotopically labeled formic acid monomers
,”
J. Mol. Spectrosc.
65
,
171
189
(
1977
).
113.
R. C.
Millikan
and
K. S.
Pitzer
, “
Infrared spectra and vibrational assignment of monomeric formic acid
,”
J. Chem. Phys.
27
,
1305
1308
(
1957
).
114.
T.
Shimanouchi
,
Tables of Molecular Vibrational Frequencies, Consolidated Volume I
(
National Institute of Standards and Technology
,
1972
).
115.
D.
McNaughton
,
C. J.
Evans
,
S.
Lane
, and
C. J.
Nielsen
, “
The high-resolution FTIR far-infrared spectrum of formamide
,”
J. Mol. Spectrosc.
193
,
104
117
(
1999
).
116.
V.
Barone
,
M.
Fusè
,
S. M. V.
Pinto
, and
N.
Tasinato
, “
A computational journey across nitroxide radicals: From structure to spectroscopic properties and beyond
,”
Molecules
26
,
7404
(
2021
).
117.
G.
Ceselin
,
Z.
Salta
,
J.
Bloino
,
N.
Tasinato
, and
V.
Barone
, “
Accurate quantum chemical spectroscopic characterization of glycolic acid: A route toward its astrophysical detection
,”
J. Phys. Chem. A
126
,
2373
2387
(
2022
).
118.
V.
Barone
,
M.
Biczysko
,
J.
Bloino
,
P.
Cimino
,
E.
Penocchio
, and
C.
Puzzarini
, “
CC/DFT route toward accurate structures and spectroscopic features for observed and elusive conformers of flexible molecules: Pyruvic acid as a case study
,”
J. Chem. Theory Comput.
11
,
4342
4363
(
2015
).
119.
H.
Hollenstein
,
F.
Akermann
, and
H. H.
Günthard
, “
Vibrational analysis of pyruvic acid and D−, 13C− and 18O-labelled species: Matrix spectra, assignments, valence force field and normal coordinate analysis
,”
Spectrochim. Acta, Part A
34
,
1041
1063
(
1978
).
120.
K. L.
Plath
,
K.
Takahashi
,
R. T.
Skodje
, and
V.
Vaida
, “
Fundamental and overtone vibrational spectra of gas-phase pyruvic acid
,”
J. Phys. Chem. A
113
,
7294
7303
(
2009
).
121.
R. A.
Shaw
,
C.
Castro
,
R.
Dutler
,
A.
Rauk
, and
H.
Wieser
, “
The vibrational spectra (100–1600 cm−1) and scaled ab initio STO‐3G and 3‐21G harmonic force fields for norbornane, norbornene, and norbornadiene
,”
J. Chem. Phys.
89
,
716
731
(
1988
).
122.
F.
Vazart
,
C.
Latouche
,
P.
Cimino
, and
V.
Barone
, “
Accurate infrared (IR) spectra for molecules containing the C≡N moiety by anharmonic computations with the double hybrid B2PLYP density functional
,”
J. Chem. Theory Comput.
11
,
4364
4369
(
2015
).
123.
T.
Fornaro
,
D.
Burini
,
M.
Biczysko
, and
V.
Barone
, “
Hydrogen-bonding effects on infrared spectra from anharmonic computations: Uracil–water complexes and uracil dimers
,”
J. Phys. Chem. A
119
,
4224
4236
(
2015
).
124.
G.
Wenzel
,
I. R.
Cooke
,
P. B.
Changala
,
E. A.
Bergin
,
S.
Zhang
,
A. M.
Burkhardt
,
A. N.
Byrne
,
S. B.
Charnley
,
M. A.
Cordiner
,
M.
Duffy
,
Z. T. P.
Fried
,
H.
Gupta
,
M. S.
Holdren
,
A.
Lipnicky
,
R. A.
Loomis
,
H. T.
Shay
,
C. N.
Shingledecker
,
M. A.
Siebert
,
D. A.
Stewart
,
R. H. J.
Willis
,
C.
Xue
,
A. J.
Remijan
,
A. E.
Wendlandt
,
M. C.
McCarthy
, and
B. A.
McGuire
, “
Detection of interstellar 1-cyanopyrene: A four-ring polycyclic aromatic hydrocarbon
,”
Science
386
,
810
813
(
2024
).
125.
G.
Wenzel
,
T. H.
Speak
,
P. B.
Changala
,
R. H. J.
Willis
,
A. M.
Burkhardt
,
S.
Zhang
,
E. A.
Bergin
,
A. N.
Byrne
,
S. B.
Charnley
,
Z. T. P.
Fried
,
H.
Gupta
,
E.
Herbst
,
M. S.
Holdren
,
A.
Lipnicky
,
R. A.
Loomis
,
C. N.
Shingledecker
,
C.
Xue
,
A. J.
Remijan
,
A. E.
Wendlandt
,
M. C.
McCarthy
,
I. R.
Cooke
, and
B. A.
McGuire
, “
Detections of interstellar aromatic nitriles 2-cyanopyrene and 4-cyanopyrene in TMC-1
,”
Nat. Astron.
9
,
262
(
2024
).