In this work, we develop a new semiempirical method, dubbed NOTCH (Natural Orbital Tied Constructed Hamiltonian). Compared to existing semiempirical methods, NOTCH is less empirical in its functional form as well as parameterization. Specifically, in NOTCH, (1) the core electrons are treated explicitly; (2) the nuclear–nuclear repulsion term is calculated analytically, without any empirical parameterization; (3) the contraction coefficients of the atomic orbital (AO) basis depend on the coordinates of the neighboring atoms, which allows the size of AOs to depend on the molecular environment, despite the fact that a minimal basis set is used; (4) the one-center integrals of free atoms are derived from scalar relativistic multireference equation-of-motion coupled cluster calculations instead of empirical fitting, drastically reducing the number of necessary empirical parameters; (5) the (AA|AB) and (AB|AB)-type two-center integrals are explicitly included, going beyond the neglect of differential diatomic overlap approximation; and (6) the integrals depend on the atomic charges, effectively mimicking the “breathing” of AOs when the atomic charge varies. For this preliminary report, the model has been parameterized for the elements H–Ne, giving only 8 empirical global parameters. Preliminary results on the ionization potentials, electron affinities, and excitation energies of atoms and diatomic molecules, as well as the equilibrium geometries, vibrational frequencies dipole moments, and bond dissociation energies of diatomic molecules, show that the accuracy of NOTCH rivals or exceeds those of popular semiempirical methods (including PM3, PM7, OM2, OM3, GFN-xTB, and GFN2-xTB) as well as the cost-effective ab initio method Hartree–Fock-3c.
I. INTRODUCTION
Semiempirical methods are quantum chemistry methods with intermediate cost, accuracy, and degree of empiricism between density functional theory (DFT) and molecular mechanics (MM) methods and have found wide use in the study of large molecules (>1000 atoms) and the molecular dynamics/conformational search/pre-optimization of moderately sized molecules, to name a few.1,2 As approximate, parameterized versions of minimal basis Hartree–Fock (HF) or DFT calculations, semiempirical methods are more robust and describe more of the underlying physics than classical MM methods while typically requiring only a few hundredths of the time of a DFT calculation with a double-zeta basis set, even when efficient approximations, such as resolution of the identity (RI)3 and chain-of-spheres exchange (COSX),4–7 are applied to the latter. When properly parameterized, semiempirical methods can often achieve useful accuracy on a handful of properties of sufficiently “well-behaved” molecules, although like any other method with a fair amount of empiricism, it is notoriously difficult to transfer the accuracy to properties and types of molecules that are not used in the training procedure. Intriguingly, few methods are known whose timing and accuracy are between those of most semiempirical methods and RI-DFT. Thus, even if slightly more computational budget is available than what is required for a semiempirical calculation, one usually cannot take advantage of this to improve the accuracy of the computational results.
The most popular semiempirical methods mainly fall into two categories: the zero-differential overlap (ZDO) methods (e.g., the PMx8–10 and OMx11 families of methods) and tight-binding (TB) methods (e.g., the DFTB12,13 and xTB14–16 families of methods), respectively. In the ZDO family of methods,17,18 the following approximations are made (among others): (1) Only the (AA|AA)- and (AA|BB)-type two-electron integrals are retained, i.e., , which is the chief reason for the low cost of the ZDO methods; (2) consistent with the above approximation, the Roothaan–Hall equation FC = SCϵ—where F, C, S, and ϵ are the Fock, molecular orbital (MO) coefficient, overlap, and orbital energy matrices, respectively—is replaced by FC = Cϵ, i.e., the overlap matrix S is set to the unit matrix. On the other hand, the TB methods12–14 usually do not approximate S as the unit matrix, while they describe the electrostatic (and exchange-correlation) interaction as a damped interaction between atomic charges (usually based on a Mulliken population analysis), thereby achieving a similar cost as the ZDO methods. However, most TB methods lack an exact exchange contribution, which introduces considerable self-interaction errors that can be difficult to remedy and makes them unable to describe finer electronic structure effects, such as electronic multiplets or spin state splittings.
The present work aims to develop a new semiempirical method, NOTCH (Natural Orbital Tied Constructed Hamiltonian), which is designed to fill up the timing and accuracy gap between the common semiempirical methods and RI-DFT. It makes fewer approximations and, at the same time, introduces less empiricism than both the ZDO and the TB methods. As its name suggests, the NOTCH atomic orbital (AO) basis functions are designed to mimic atomic natural orbitals (ANOs) as closely as possible. Even more importantly, the NOTCH one-center one-electron and two-electron integrals of free atoms, as well as those of monocationic and monoanionic atomic ions, are required to reproduce the ANO basis integrals of an effective ab initio Hamiltonian that implicitly includes dynamic correlation and scalar relativity, i.e., the second-order Douglas–Kroll–Hess (DKH2)19,20 similarity-transformed multireference equation-of-motion coupled cluster (MR-EOM-CC)21–24 Hamiltonian. In this way, the parameterization of atomic integrals is solved in a convenient, systematic, and completely nonempirical way. The nonempirical atomic integrals serve also as a starting point for building the one-electron and two-electron integrals in molecules.
A few other salient points of our construction include the following:
In NOTCH, the core electrons are treated explicitly, rather than approximated by absorbing their contribution into the nuclear attraction potential, as what is done in most semiempirical methods. This obviates the need for complicated treatments of the core–core repulsion energy, which have often been “abused” to absorb the errors of other parts of the semiempirical method and contributed significantly to the highly empirical nature of semiempirical methods and a partially severe lack of transferability.
In NOTCH, the radial distributions of the AOs depend on the molecular geometry, so that, e.g., an AO becomes more compact in a molecular environment than in the free atom. By this way, we mimic a double-zeta basis calculation within a minimal basis framework. Semiempirical methods employing geometry-dependent basis sets are rare, one prominent example being sTDA-xTB.25
The radial distributions of the AOs also depend on the atomic charges in NOTCH, i.e., the more positive charge an atom carries, the more compact its AOs will be. A simple way to account for this is to add charge-dependent corrections to the one-center one-electron and two-electron integrals. Similar ideas have been explored in ZDO26–28 and TB15,16,29,30 methods.
The NOTCH AO two-electron integrals are first calculated analytically from the AO basis functions, which are subsequently corrected in order to incorporate correlation effects. This stands in contrast to the common practice of using crude approximations like the Dewar–Sabelli–Klopman formula31–35 to approximate the integrals and then hope to absorb dynamic correlation effects and various errors by parameterization. Note that the left–right correlation (LRC) of bonds, which is inherently nonlocal, is also explicitly accounted for.
In NOTCH, we retain not only the AO basis (AA|AA) and (AA|BB)-type two-electron integrals but also the (AA|AB) and (AB|AB)-type ones.
In NOTCH, the dispersion interaction is not described as a post-self-consistent field (SCF) correction to the energy but as a correction to the (AA|BB)-type two-electron integrals. The main rationale for this choice is that dispersion is inherently a correlation effect and thus represents an interaction between electrons, not between atoms. Similar ideas have been pursued in some semiempirical methods36 but have not gained wide attention.
In NOTCH, a classical charge polarization term is added as a correction to the one-electron integrals to account for the underestimation of polarizability by the minimal basis set formalism. The idea of adding a classical charge polarization term has been explored before, but as a post-SCF correction to the total energy instead of a correction to the integrals.37,38 The same correction also mimics the charge polarization due to bonding and therefore improves the description of bonding as well.
It should be noted that the present work is preliminary, as only the elements from H to Ne are parameterized, and only results for diatomic molecules are reported. Nevertheless, our benchmark results on all the possible neutral diatomic molecules that can form between the elements H–Ne show that NOTCH provides equilibrium bond length, vibrational frequency, dipole moment, ionization potential (IP), electron affinity (EA), bond dissociation energy, and excitation energy results that are at least on par with, and frequently outperform, the best existing semiempirical methods. More importantly, NOTCH is characterized by a lack of severe outliers, which can at least partly be traced back to NOTCH’s greatly reduced empiricism compared to other semiempirical methods. The possible ways to extend NOTCH to multiatomic molecules will be briefly discussed at the end of the manuscript.
II. THEORY
A. Notations
Throughout the manuscript, the following notations are used: Atoms are denoted by A, B, C, …, AOs and orthogonalized AOs (OAOs) are denoted by μ, ν, κ, λ, …, occupied and virtual MOs are labeled by i, j, … and a, b, …, respectively; spin indices are denoted by σ, τ, …. The notation μA stands for a primitive AO, a contracted AO, or an OAO centered on atom A.
Integrals without any superscript, e.g., , hμν, Tμν, or , denote analytic, uncorrected AO integrals calculated over ano-pc-0 basis functions (vide infra). Integrals with overbars [e.g., ] stand for corrected integrals that are explained in detail below.
As the present manuscript deals with monoatomic and diatomic systems only, it will sometimes be assumed that the number of atoms in the system is no larger than 2, in order to simplify the expressions. Likewise, for the sake of simplicity, we may sometimes assume that all the elements in the molecule belong to the first two periods, but it is trivial to generalize the conclusions to heavier elements.
B. Overview
Another special feature of NOTCH is that the integrals depend on the atomic charges (and therefore on the density matrix D). In this way we are able to mimic the expansion/shrinking of AOs as a result of the atom acquiring negative/positive charge, respectively (which otherwise requires at least a split valence basis to describe), despite using a minimal basis set.
For the sake of simplicity, in the remaining parts of the manuscript, we use and to denote the “zero-charge” core Hamiltonian and two-electron integrals, respectively, i.e., the integrals one obtains in the case that all atomic charges are exactly zero.
C. Basis set
A scalar relativistic second-order DKH2 state-averaged (SA)-complete active space self-consistent field (CASSCF) calculation with the decontracted pc-0 basis set is carried out for the neutral atoms of the elements Li–Ne, where the CAS active space includes all valence orbitals, and all states are averaged. (The DKH2 scalar relativistic treatment is of course unnecessary for the elements H–Ne, but will be necessary for the heavy elements that will be parameterized in the future; we have thus used DKH2 here as well, for consistency reasons.) For Li and Be, we consider the 2p orbitals as belonging to the valence orbitals, and furthermore we add the diffuse p function from the aug-pc-0 basis set to improve the description of the 2p orbitals. The converged inactive orbital is then taken as the 1s function of the ano-pc-0 basis set.
The valence functions of the ano-pc-0 basis set are also given as contractions of the pc-0 primitive functions (plus the aug-pc-0 diffuse p functions for Li and Be). We first variationally optimized the contraction coefficients of the valence functions to minimize the DKH2 state specific (SS)-CASSCF ground state energies of the homodimers of elements H–Ne at the bond lengths 3.0, 2.9, …, 0.1 Å, under the constraint that the 2s function (if it exists) is orthogonal to the 1s function. We then interpolate the optimized contraction coefficients as a function of the bond length, with the following formula:
The radial parts of some selected ano-pc-0 basis functions of atoms and diatomic molecules are plotted in Fig. 1 as a function of the bond length. As expected, the basis functions are generally more compact in covalently bonded diatomic molecules than in free atoms, and become progressively more compact when the bond shortens or (in the case of H 1s) when the other atom becomes more and more electronegative. Thus, ano-pc-0 mimics a double-zeta basis set despite being a minimal basis set, thanks to the geometry dependence of the contraction coefficients. It is also interesting to note that the radial distribution of the C 2s AO is nearly invariant with respect to the chemical environment, probably because it has to remain orthogonal to the 1s AO, which restricts its ability to shrink in a molecular environment. However, there are two important effects that ano-pc-0 cannot yet capture: (1) the px, py, and pz functions of an atom can have different radial distribution profiles, a frequently cited example being linear molecules where the p orbital parallel to the bond axis (pσ) is typically more compact than those perpendicular to the bond axis (pπ); and (2) the basis functions may deform by mixing with polarization functions. These effects are crucial for an accurate prediction of equilibrium bond lengths and vibrational frequencies, and will be discussed in Sec. II G.
Radial parts of some ano-pc-0 basis functions in different diatomic molecules with different bond lengths. The C 2p (F-C, R = 1.3 Å) curve nearly coincides with the C 2p (C-C, R = 1.5 Å) one.
Radial parts of some ano-pc-0 basis functions in different diatomic molecules with different bond lengths. The C 2p (F-C, R = 1.3 Å) curve nearly coincides with the C 2p (C-C, R = 1.5 Å) one.
D. NOTCH integrals of free atoms and atomic ions
To keep the empiricism of NOTCH at the lowest possible level while still obtaining decent and uniform accuracy, the core Hamiltonian and two-electron NOTCH integrals of free atoms are determined from DKH2 multireference equation-of-motion coupled cluster singles and doubles (MR-EOM-CCSD)/ANO-RCC-Full45–47 calculations of neutral atoms. We make full use of the fact that the ano-pc-0 AOs resemble the atomic natural orbitals (since they are optimized for atomic CASSCF calculations), so that for atomic calculations, the integrals of the former can be replaced by those of the latter to take advantage of the higher basis set quality and (more importantly) the implicit inclusion of dynamic correlation of the latter. This construction is at the heart of the NOTCH idea and is the reason why we insisted on using a generally contracted basis set. In addition, we require that the 1s and 2s functions of every element be orthogonal: a segmented contracted or non-orthogonal basis set may expand the correct s-space, but its basis functions generally do not resemble the exact atomic natural orbitals as orthogonal generally contracted basis sets can do.
The rationale of this construction is that, a multireference configurational interaction singles (MR-CIS) calculation with NOTCH integrals will then closely reproduce the MR-EOM excitation energies (the reproduction is not exact, because the NOTCH MR-CIS calculation does not include Rydberg orbitals). This is corroborated by our benchmark results of atomic excitation energies, as shown in Table I: the NOTCH MR-CIS excitation energies of atoms reproduce the MR-EOM ones (which in turn agrees with the experimental results48 to about 0.1 eV) to within a few tenth of an eV, with only one exception (fluorine). Thus, the NOTCH two-electron integrals can be said to implicitly include dynamic correlation with close to MR-CCSD quality. While a slightly better accuracy is achieved by the INDO/S intermediate neglect of differential overlap/spectroscopy49,50 method, this is mostly due to the fact that the INDO/S integrals were fitted against this exact set of experimental spectra, unlike the NOTCH integrals that were derived from higher-level calculations without any empiricism. Semiempirical methods that were not fitted to excitation energies at all, such as PM38 and OM2,51,52 perform very poorly and yield mean absolute deviations (MADs) around 2–3 eV. The largest outlier, the 12P → 12S excitation energy of the fluorine atom, is predicted to be 17.615 and 9.474 eV too low by PM3 and OM2, respectively. This is probably due to the fact that ground state, fluorine-containing molecules generally do not have a hole in the fluorine 2s orbital, making typical ground state molecule training sets incapable of properly constraining the integrals over 2s orbitals. While PM3 is well-known to be unsuitable for excitation energies in general,53 the failure of OM2 is surprising since random phase approximation (RPA) and multireference configuration interaction (MRCI)54 calculations with OM2 integrals are routinely used to calculate the excited states of molecules. This again highlights the importance of deriving the one-center integrals from the MR-EOM Hamiltonian, instead of fitting against reference data.
Atomic excitation energies (eV) of elements Li–F computed by MR-CIS using NOTCH, INDO/S, PM3 and OM2 integrals, compared with DKH2 MR-EOM-CCSD/ANO-RCC-Full and experimental results.
Atom (ground state) . | Excited state . | NOTCH . | INDO/S . | PM3 . | OM2 . | MR-EOM . | Expt. . |
---|---|---|---|---|---|---|---|
Li [1s22s1 (2S)] | 1s22p1 (2P) | 1.846 | 1.850 | 2.407 | N.A.a | 1.846 | 1.848 |
Be [1s22s2 (1S)] | 1s22s12p1 (3P) | 2.908 | 3.258 | N.A.a | N.A.a | 2.866 | 2.725 |
1s22s12p1 (1P) | 5.961 | 5.810 | N.A.a | N.A.a | 5.280 | 5.277 | |
B [1s22s22p1 (2P)] | 1s22s12p2 (4P) | 2.623 | 4.154 | 6.998 | N.A.a | 3.703 | 3.551 |
1s22s12p2 (2D) | 5.706 | 6.789 | 9.988 | N.A.a | 5.985 | 5.932 | |
1s22s12p2 (2S) | 7.054 | 8.043 | 11.758 | N.A.a | 8.092 | 7.879 | |
1s22s12p2 (2P) | 9.168 | 9.555 | 12.428 | N.A.a | 8.916 | 8.991 | |
C [1s22s22p2 (3P)] | 1s22s22p2 (1D) | 1.217 | 1.082 | 1.754 | 1.240 | 1.148 | 1.260 |
1s22s22p2 (1S) | 3.042 | 2.706 | 3.583 | 2.265 | 2.646 | 2.680 | |
1s22s12p3 (5S) | 3.114 | 5.168 | 3.847 | 4.516 | 4.310 | 4.179 | |
1s22s12p3 (3D) | 7.737 | 9.090 | 8.768 | 8.806 | 7.949 | 7.942 | |
N [1s22s22p3 (4S)] | 1s22s22p3 (2D) | 2.278 | 2.325 | 1.421 | 2.085 | 2.251 | 2.384 |
1s22s22p3 (2P) | 2.934 | 3.876 | 2.268 | 2.891 | 3.519 | 3.576 | |
1s22s12p4 (4P) | 10.731 | 11.180 | 8.593 | 14.026 | 10.870 | 10.927 | |
1s22s12p4 (2D) | 15.624 | 15.717 | 10.677 | 18.556 | 14.952 | 15.027 | |
O [1s22s22p4 (3P)] | 1s22s22p4 (1D) | 1.808 | 1.657 | 1.248 | 1.540 | 1.867 | 1.958 |
1s22s22p4 (1S) | 4.521 | 4.142 | 3.094 | 2.816 | 4.204 | 4.180 | |
1s22s12p5 (3P) | 16.385 | 15.896 | 18.336 | 20.669 | 15.546 | 15.650 | |
F [1s22s22p5 (2P)] | 1s22s12p6 (2S) | 23.104 | 20.257 | 3.506 | 11.647 | 21.121 | 21.039 |
MDb | 0.036 | 0.289 | −0.485 (0.585c) | 0.056 (0.923c) | |||
MADb | 0.519 | 0.420 | 2.861 (1.938c) | 2.117 (1.448c) |
Atom (ground state) . | Excited state . | NOTCH . | INDO/S . | PM3 . | OM2 . | MR-EOM . | Expt. . |
---|---|---|---|---|---|---|---|
Li [1s22s1 (2S)] | 1s22p1 (2P) | 1.846 | 1.850 | 2.407 | N.A.a | 1.846 | 1.848 |
Be [1s22s2 (1S)] | 1s22s12p1 (3P) | 2.908 | 3.258 | N.A.a | N.A.a | 2.866 | 2.725 |
1s22s12p1 (1P) | 5.961 | 5.810 | N.A.a | N.A.a | 5.280 | 5.277 | |
B [1s22s22p1 (2P)] | 1s22s12p2 (4P) | 2.623 | 4.154 | 6.998 | N.A.a | 3.703 | 3.551 |
1s22s12p2 (2D) | 5.706 | 6.789 | 9.988 | N.A.a | 5.985 | 5.932 | |
1s22s12p2 (2S) | 7.054 | 8.043 | 11.758 | N.A.a | 8.092 | 7.879 | |
1s22s12p2 (2P) | 9.168 | 9.555 | 12.428 | N.A.a | 8.916 | 8.991 | |
C [1s22s22p2 (3P)] | 1s22s22p2 (1D) | 1.217 | 1.082 | 1.754 | 1.240 | 1.148 | 1.260 |
1s22s22p2 (1S) | 3.042 | 2.706 | 3.583 | 2.265 | 2.646 | 2.680 | |
1s22s12p3 (5S) | 3.114 | 5.168 | 3.847 | 4.516 | 4.310 | 4.179 | |
1s22s12p3 (3D) | 7.737 | 9.090 | 8.768 | 8.806 | 7.949 | 7.942 | |
N [1s22s22p3 (4S)] | 1s22s22p3 (2D) | 2.278 | 2.325 | 1.421 | 2.085 | 2.251 | 2.384 |
1s22s22p3 (2P) | 2.934 | 3.876 | 2.268 | 2.891 | 3.519 | 3.576 | |
1s22s12p4 (4P) | 10.731 | 11.180 | 8.593 | 14.026 | 10.870 | 10.927 | |
1s22s12p4 (2D) | 15.624 | 15.717 | 10.677 | 18.556 | 14.952 | 15.027 | |
O [1s22s22p4 (3P)] | 1s22s22p4 (1D) | 1.808 | 1.657 | 1.248 | 1.540 | 1.867 | 1.958 |
1s22s22p4 (1S) | 4.521 | 4.142 | 3.094 | 2.816 | 4.204 | 4.180 | |
1s22s12p5 (3P) | 16.385 | 15.896 | 18.336 | 20.669 | 15.546 | 15.650 | |
F [1s22s22p5 (2P)] | 1s22s12p6 (2S) | 23.104 | 20.257 | 3.506 | 11.647 | 21.121 | 21.039 |
MDb | 0.036 | 0.289 | −0.485 (0.585c) | 0.056 (0.923c) | |||
MADb | 0.519 | 0.420 | 2.861 (1.938c) | 2.117 (1.448c) |
The semiempirical method is not parameterized for this element.
With respect to the MR-EOM results.
Statistics after the worst outlier, the 12P → 12S excitation energy of fluorine, is removed.
E. NOTCH integrals of diatomic molecules
The parameters are pre-computed for every pair of atoms AB, so that when the spherically averaged NOTCH ROHF atomic density matrices of atoms A and B are plugged into Eq. (31), the DFT-D3 and coefficients of the atom pair AB are reproduced. The remaining parameters are determined by fitting, except for s6 which is fixed at the theoretical value of 1.0, as will be detailed in Sec. II I.
F. Charge dependence of the molecular NOTCH integrals
G. Basis set incompleteness correction
Even with the above constructions, the NOTCH method still suffers from a systematic overestimation of equilibrium bond length (by 0.03–0.1 Å), as well as IPs and EAs (by 1.0–4.0 eV). This was found to be due to fundamental restrictions of the minimal basis set that are not remedied by the geometry-dependent re-contraction, i.e., (1) the constraint that the radial distributions of the px, py and pz orbitals of a given atom must be the same, and (2) the lack of polarization functions. These effects prove to be very difficult to capture without adding more basis functions to the minimal basis set.
H. Left–right correlation correction
The only remaining important contribution to the total energy is the left–right correlation of the bonded electrons, also known as the non-dynamic correlation. It is well known that semilocal DFT correlation functionals mainly describe dynamic correlation and severely underestimate (if not completely miss) the left–right correlation contribution,64 especially in elongated bonds. This is one of the major reasons why DFT functionals with 100% HF exchange typically perform poorly. As the NOTCH correlation correction described in Sec. II E uses an LDA-inspired form, it suffers from the same lack of left–right correlation, which is primarily manifested as a systematic overestimation of vibrational frequencies and IPs for certain molecules, like N2, O2, and F2, as well as an erroneous prediction of dipole moments. Therefore, we now add a final correction, , to approximately account for left–right correlation.
We conclude by a few remarks of the functional form Eq. (43):
In the DFT community, the left–right correlation is usually modeled by using a less than 100% amount of HF exchange; the difference of DFT exchange and HF exchange approximates the left–right correlation energy because it makes the exchange–correlation hole more local (as it usually should) than if the functional comprises 100% HF exchange plus a semilocal correlation functional. However, this strategy introduces the notorious self-interaction error problem, since the so-modeled left–right correlation is nonzero even for single-electron densities. The same problem obviously also exists in all self-consistent TB methods. In NOTCH, however, by construction vanishes for single-electron densities, since the 2-RDM matrix elements are zero when there is only one electron. The theory thus remains self-interaction free.
Equation (43) is a quartic functional of the density matrix, so it cannot be accurately absorbed into any ansatz where the energy only depends on the density matrix up to third order. The energy expressions of all mainstream ZDO methods are quadratic functionals of the density matrix, while some TB methods include terms cubic in the density matrix (i.e., cubic charge dependence terms of the on-site energy).15,16,29,30 This in part explains why heavily parameterized core–core repulsion potentials are ubiquitous in existing semiempirical methods because the employed functional forms are not flexible enough to describe the left–right correlation, so that many more empirical parameters have to be introduced to fit the reference data.
The parameterization procedure of and ζAB implies that NOTCH can dissociate homodiatomic molecules correctly even under a restricted HF formalism, which even many ab initio methods struggle to achieve. The dissociation is in general not exactly correct for heterodiatomic molecules or diatomic molecular ions but in general still better than if was not present. While this may not be directly relevant in real-world calculations (since it is usually acceptable to use broken-symmetry wavefunctions, which naturally lead to the correct dissociation limits), it is responsible for curing the systematic overestimation of energy at long but finite bond lengths, leading to better vibrational frequencies at equilibrium geometries (and potentially also the systematic overestimation of barrier height, which is common for methods with 100% HF exchange).
The use of the uncorrelated 2-RDM [Eq. (42)] is crucial here. Take H2 as an example: The exact correlated 2-RDM element decays to zero in the RAB → +∞ limit, giving zero correlation correction, while the uncorrelated RHF 2-RDM shows that there is still one pair of shared electrons, and thus correctly yields a correlation correction that restores the UHF dissociation limit (see Fig. S1 for the uncorrelated and correlated 2-RDM elements of H2 as a function of the bond length). Thus, Eq. (42) should not be understood as an approximation of the correlated 2-RDM. This is reminiscent of the fact that, the Wiberg bond order of, e.g., H2 calculated from the correlated density matrix goes to zero in the RAB → +∞ limit and therefore does not reflect the number of shared electrons in the RHF wavefunction.
To illustrate the effect of the LRC correction (especially near the dissociation region), we have plotted the potential energy curves of six representative diatomic molecules (H2, N2, F2, CH, HF, LiF) with NOTCH, both with and without the LRC correction (Fig. 2). As LRC strongly stabilizes the RHF wavefunction in the dissociation region but less so near the equilibrium geometry, it corrects the energy overestimation of the other terms of the NOTCH method in the dissociation region, including the highly ionic molecule LiF, although the correction is sometimes too large (especially for F2 and CH). Perhaps more importantly, the preferential stabilization of the dissociation region also increases the equilibrium bond length and reduces the vibrational frequency at the equilibrium geometry, which is especially obvious for F2 and CH. Therefore, despite the fact that the LRC correction does not consistently improve the bond energy and UHF potential energy curves in the dissociation region, we expect that it should in general be beneficial for geometries and vibrational frequencies, as will be confirmed in Secs. III B–III C.
Potential energy curves of selected diatomic molecules, calculated by NOTCH with and without the LRC correction, compared against DKH2 UCCSD(T)/ANO-RCC-TZP results.
Potential energy curves of selected diatomic molecules, calculated by NOTCH with and without the LRC correction, compared against DKH2 UCCSD(T)/ANO-RCC-TZP results.
Another property of the LRC correction is that the RHF instability point of the NOTCH method occurs later than if the LRC correction is not present, especially in F2 and CH; this can be compared to the well-known fact that the ground state wavefunction instabilities of DFT methods generally occur at longer bond lengths than the HF method. For H2, the restricted NOTCH wavefunction even remains stable all the way until dissociation, so that the RHF NOTCH and UHF NOTCH curves coincide. For N2 and F2, the LRC correction also makes the RHF NOTCH dissociation limit coincide with the UHF one (by construction), but an RHF instability is still observed when the molecules are not fully dissociated; thus, while the RHF NOTCH curves exhibit unphysical bumps, the UHF NOTCH ones remain qualitatively correct. In the case of the HF molecule, the RHF NOTCH dissociation limit is wrong, but again the UHF NOTCH curve is at least as well behaved as when the LRC correction is not present.
Overall, our preliminary version of NOTCH is a semiempirical method featuring minimal empiricism and explicitly describes much more physics than existing semiempirical methods: (1) the local dynamic correlation and scalar relativistic effects are included through nonempirical MR-EOM calculations (Sec. II D) and the former is properly scaled down in a molecular environment (Sec. II E); (2) the nonlocal component of dynamic correlation, i.e., dispersion, is accounted for as a correction to the (AA|BB)-type two-electron integrals (Sec. II E); (3) the non-dynamic or left–right correlation is described by a separate term inspired by Wiberg bond orders (Sec. II H); (4) the radial distortions of AOs as a function of geometry (Sec. II C) and atomic charge (Sec. II F) are accounted for by corrections to the AO contraction coefficients and one-center integrals, respectively; and (5) the angular distortions of AOs are approximated by adding a classical charge polarization term (Sec. II G). It can be seen that although our method is built upon a HF/minimal basis framework, all relativity, correlation, and basis set incompleteness effects are included in a physically transparent way, and no attempt is made to absorb any effect into terms whose functional forms are not designed to capture that effect.
I. Parameterization
The present form of NOTCH contains eight global empirical parameters: (1) the parameter αcorr responsible for removing the double-counting of dynamic correlation [Eq. (27)]; (2) the dispersion correction parameters s8, a1, and a2 [Eq. (31)]; (3) the parameter βCDI used for damping the charge dependence correction of integrals [Eq. (37)]; (4) the parameter βpol used for damping the basis set incompleteness correction [Eq. (40)]; and (5) the parameters and involved in the resonance integral in the LRC correction [Eq. (47)]. They were fitted via the following procedure:
The parameters αcorr, s8, a1, and a2 were fitted against the domain-based local pair natural orbital (DLPNO)-coupled-cluster singles and doubles with perturbative triples [CCSD(T)]66,67 local energy decomposition (LED)68,69 dispersion energies of all 55 diatomic molecules constructible from the elements H–Ne, at bond distances that are 0.9, 1.0, 1.2, 1.5, and 2.0 times the sum of the respective van der Waals radii (for details, see Appendix B).
The parameters βpol, , and are obtained by fitting the RHF NOTCH potential energy curve of H2 against DKH2 CCSD(T)/ANO-RCC-TZP.45–47
The parameter βCDI is fitted against the DKH2 HF/ANO-RCC-TZP dissociation energy of H2+.
Since there are some couplings between the parameters, the above procedure is iterated a few times until all the parameters converge. Although the parameters βpol, , , and βCDI are fitted against H2 and H2+ only, the benchmark results (Sec. III) do not improve noticeably when we add more molecules (such as N2 and F2) to the training set, suggesting that the parameters are well transferable among different diatomic molecules.
In addition, NOTCH has many parameters that are not derived from fitting against high-level reference data, including (1) the parameters used for calculating the geometry-dependent contraction coefficients of the ano-pc-0 basis set (Sec. II C), fitted from CASSCF calculations of ground state homodiatomic molecules; (2) the Gaussian exponents and coefficients used for expanding the spherically averaged density of atoms, (Sec. II C), and atomic monocations, (Sec. II G), fitted against spherically averaged ROHF atomic densities; (3) the MR-EOM one-electron and two-electron integrals (Sec. II D), generated from atomic MR-EOM calculations; (4) the location of the radial grid point in the numerical integration of Eq. (29) (Sec. II E), fitted from numerically integrating Eq. (30) for spherically averaged atoms; (5) the atomic polarizability correction (Sec. II G), obtained from polarizability calculations of atoms; and (6) the element-specific parameters and ζAA (Sec. II H) involved in the LRC correction, determined by matching the RHF and UHF NOTCH energies for homodiatomic molecules at the dissociation limit. These parameters, as well as detailed descriptions for the procedures to generate them, are tabulated in the supplementary material for convenience. Note that all the nonempirical parameters are derived from the calculations of either atoms or homodiatomic molecules; thus, it is easy to extend the parameterization to heavier elements. While the global, empirical parameters (Table II) may need to be re-fitted when we parameterize more elements, we expect that the reference data of a very few number of heavy element-containing molecules should suffice for the fitting.
The global NOTCH parameters. All parameters are given in atomic units.
Parameter . | αcorr . | s8 . | a1 . | a2 . | βCDI . | βpol . | . | . |
---|---|---|---|---|---|---|---|---|
Value | 0.080 | 0.711 | 0.752 | 5.773 | 0.793 | 0.240 | 0.130 | 0.094 |
Parameter . | αcorr . | s8 . | a1 . | a2 . | βCDI . | βpol . | . | . |
---|---|---|---|---|---|---|---|---|
Value | 0.080 | 0.711 | 0.752 | 5.773 | 0.793 | 0.240 | 0.130 | 0.094 |
Finally, based on the comments of a referee to this paper, we would like to briefly discuss the meaning of the term “empirical” in semiempirical theories. Empiricism, as a deviation from pure ab initio theories, may enter a semiempirical method in various ways:
by fitting physically more or less well-defined parameters that are atom or even atom pair specific to a set of molecular reference data;
by fitting a set of element-specific parameters to experimental data obtained on free atoms and ions;
by replacing well-defined quantities in ab initio theory, for example, analytical AO integrals, with simplified and possibly parameterized expressions—the latter may or may not respect the known asymptotic behavior of the corresponding ab initio quantities; or
by inventing correction terms that have no counterpart in ab initio theory and are based on plausibility arguments.
Based on these levels of empiricism, we wish to comment on the level of empiricism in NOTCH:
In NOTCH, the one-center integrals are derived from high-level calculations on atoms and ions. There is no fitting involved in this process. The parameters are taken directly from the twice similarity-transformed Hamiltonian of MR-EOM calculations in an ANO basis. This is manifestly less empirical than fitting to experimental data of free atoms and ions, let alone the experimental data of molecules. We therefore regard our current approach as nonempirical.
In NOTCH, the “breathing” of basis functions due to the chemical environment is modeled by fitting them to basis functions optimized in an ab initio fashion for homodiatomic molecules. Importantly, the latter basis functions are not optimized to reproduce high-level data obtained either experimentally or computationally (like the case of most semiempirical methods, whose basis set exponents are treated as free parameters and fitted against experimental data) but are rather optimized to minimize the diatomic CASSCF energy computed using that basis set (like the case of any ab initio basis set whose basis functions are optimized variationally on atoms and possibly molecules). Thus, we regard this form of fitting also as nonempirical, even though molecular calculations are involved.
In NOTCH, there are no element-specific empirical parameters that are derived from fitting to atomic or molecular reference data. There are a handful of semiempirical parameters that are fitted to molecular reference data, which is indeed an empirical element of the method. However, the total number of parameters determined in this way (8) is much smaller as for example, in PM79 (225) or GFN2-xTB16 (194) for the elements H–Ne.
In NOTCH, physically motivated, semiempirical expressions are used to approximately incorporate well-defined physical phenomena that all have a counterpart in ab initio theory. This obviously is a form of empiricism, but here we stress the close ties to ab initio theory. Most of the NOTCH correction terms can be turned on or off with well-defined and physically meaningful changes in the results of the methods (Sec. III); this is in stark contrast with many other semiempirical methods or even some density functionals, where individual energy contributions do not have rigorous ties to any physical quantity, since only the sum of the contributions was constrained by the parameterization process. Our expressions reproduce most of the known asymptotic behaviors of their ab initio counterparts to the extent that semiempirical methods can possibly achieve.
Based on these criteria, we believe that NOTCH is significantly less empirical than practically all semiempirical methods in use, while at least up to this point in the development, NOTCH is characterized by a far more uniform accuracy across a large array of molecular properties, as will be shown in Sec. III. We are hopeful that this behavior will translate to larger molecules.
J. Implementation
The NOTCH method has been implemented in a development version of ORCA.70 The program workflow of a NOTCH single point calculation closely resembles a minimal basis HF calculation and is given as follows:
The ano-pc-0 basis set contraction coefficients of the molecule in question, , are generated on the fly, via the procedure detailed in Sec. II C.
The zero-charge one-electron and two-electron integrals, and , are calculated under the AO basis (Sec. II E), transformed to the OAO basis, and stored on disk. The transformation to the OAO basis is not necessary for diatomic calculations (since the Fock builds can be done under the AO basis instead) but may become essential for multiatomic molecules (see Sec. IV).
- A standard integral conventional HF calculation is then carried out. During each SCF iteration, the usual one-electron and two-electron components of the Fock matrix, FOAO, are assembled under the OAO basis using the pre-calculated, zero-charge integrals and the OAO basis density matrix DOAO. The contributions from the , , and terms in Eq. (5) are then calculated under the OAO basis and added onto FOAO. The derivatives of the integrals with respect to the density matrix elements are given by the chain rule, e.g.,(50)
It is then elementary to derive the working formulas of the derivatives of the integrals, which are given in the supplementary material.
The contribution of the charge dependence of integrals (Sec. II G) and the LRC correction (Sec. II H) to the energy are then calculated and added to the total energy [Eq. (1)].
The final Fock matrix can be transformed back to the AO basis to yield the Roothaan–Hall equation (3), which is then solved in the standard way to give the molecular orbitals of the next SCF iteration. For large systems, however, it may be advantageous to directly diagonalize the Fock matrix under the OAO basis to save a few matrix operations.
III. RESULTS AND DISCUSSIONS
In this section, we calculate a few properties of atoms and diatomic molecules using NOTCH and some other semiempirical methods (PM3,8 PM7,9 OM2,51,52 OM3,71 GFN-xTB,15 GFN2-xTB16), as well as two cost-effective ab initio methods (HF-3c43 and B3LYP-D360,72,73/def2-SVP62). All calculations at the NOTCH, PM3, GFN-xTB, GFN2-xTB, HF-3c, and B3LYP-D3/def2-SVP levels were conducted by a development version of ORCA,70 while the PM7 and OM2/OM3 calculations were performed with the MOPAC74 and MNDO75 programs, respectively. The reference values for all ground state properties were generated at the DKH2 CCSD(T)/ANO-RCC-Full (for atomic calculations) and DKH2 CCSD(T)/ANO-RCC-TZP (for molecular calculations) levels of theories. The reference excitation energies were generated at the DKH2 multireference average coupled-pair functional (MRACPF)76/ANO-RCC-TZP level of theory; herein MRACPF was used instead of the more rigorous MR-EOM-CCSD method due to the better numerical stability of the former.
A. Atomic IPs and EAs
As a first check, we report the atomic IPs and EAs calculated at the NOTCH level and compare them against the reference results. All IPs and EAs are calculated as differences of single-reference unrestricted SCF energies. As can be seen from Tables III and IV, the MADs of NOTCH atomic IPs and EAs are on the order of 0.4 eV and do not exhibit systematic bias. However, if the neutral atom integrals are used in the calculations of the atomic cations and anions, the MADs increase to 1.63 and 2.42 eV for IPs and EAs, respectively, with the IPs being systematically overestimated and the EAs being systematically underestimated; in particular, all the atomic monoanions are predicted to be unbound. This clearly illustrates the importance of introducing the charge dependence to the integrals. It is interesting to note that none of the other semiempirical methods outperform NOTCH for either the IPs or the EAs, and while the B3LYP-D3/def2-SVP IPs are more accurate than the NOTCH ones, the corresponding EAs are inferior to the NOTCH results. The latter shows that the charge dependence of NOTCH integrals introduces a radial flexibility in the basis set that goes beyond that of even def2-SVP. While the GFN-xTB and GFN2-xTB methods do effectively account for the charge dependence of integrals owing to an energy term that depends cubically on the atomic charges, in practice GFN-xTB and GFN2-xTB perform very poorly for the IPs and EAs of atoms. Indeed, it has been recommended to use a specially parameterized version of GFN-xTB to calculate the IPs and EAs of molecules; even then, an empirical shift on the order of 5 eV must be applied to the IPs and EAs.15 Also of note is the presence of several severe outliers in various semiempirical methods [e.g., the IPs of Ne (PM7), He, O (GFN-xTB), and the EAs of Li (PM7) and Be (GFN2-xTB)], which clearly represent corner cases that are not found in typical training sets. The nonempirical atomic integrals of NOTCH, however, ensure that NOTCH remains reliable even in these cases. It must be noted, however, that atoms (and many of the diatomic molecules that will be mentioned in the following sections) are not the target of most semiempirical methods.
Atomic IPs (eV) of elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-Full results.
Element . | NOTCH . | NOTCHa . | PM3 . | PM7 . | OM2 . | OM3 . | GFN-xTB . | GFN2-xTB . | HF-3c . | B3LYP-D3/def2-SVP . | CCSD(T) . |
---|---|---|---|---|---|---|---|---|---|---|---|
H | 13.61 | 13.61 | 13.07 | 11.07 | 12.65 | 12.46 | 17.32 | 17.10 | 12.68 | 13.55 | 13.61 |
He | 24.56 | 26.09 | N.A.b | 22.33 | N.A.b | N.A.b | 43.09 | 34.27 | 24.59 | 24.78 | 24.56 |
Li | 5.42 | 5.40 | 5.13 | 4.80 | N.A.b | N.A.b | 11.00 | 9.42 | 5.26 | 5.46 | 5.39 |
Be | 8.52 | 8.73 | N.A.b | 7.84 | N.A.b | N.A.b | 14.09 | N.A.c | 7.74 | 9.00 | 9.31 |
B | 8.41 | 8.62 | 5.32 | 4.04 | N.A.b | N.A.b | 13.54 | N.A.c | 6.88 | 8.73 | 8.26 |
C | 11.51 | 12.24 | 9.86 | 8.22 | 10.01 | 9.87 | 17.23 | N.A.c | 9.87 | 11.51 | 11.23 |
N | 14.90 | 16.41 | 13.28 | 14.09 | 13.63 | 13.42 | 19.56 | 18.91 | 13.79 | 14.58 | 14.52 |
O | 12.17 | 16.67 | 12.76 | 10.54 | 13.43 | 13.41 | 26.17 | 22.39 | 12.87 | 13.91 | 13.51 |
F | 17.99 | 20.91 | 16.39 | 17.25 | 17.06 | 17.29 | 28.58 | 24.84 | 16.90 | 17.44 | 17.34 |
Ne | 22.35 | 25.70 | N.A.b | 5.30 | N.A.b | N.A.b | 24.76 | 31.86 | 21.47 | 21.35 | 21.53 |
MD | 0.02 | 1.51 | −2.17 | −3.38 | −0.68 | −0.75 | 7.61 | 6.90 | −0.72 | 0.11 | |
MAD | 0.45 | 1.63 | 2.17 | 3.38 | 0.68 | 0.75 | 7.61 | 6.90 | 0.72 | 0.22 |
Element . | NOTCH . | NOTCHa . | PM3 . | PM7 . | OM2 . | OM3 . | GFN-xTB . | GFN2-xTB . | HF-3c . | B3LYP-D3/def2-SVP . | CCSD(T) . |
---|---|---|---|---|---|---|---|---|---|---|---|
H | 13.61 | 13.61 | 13.07 | 11.07 | 12.65 | 12.46 | 17.32 | 17.10 | 12.68 | 13.55 | 13.61 |
He | 24.56 | 26.09 | N.A.b | 22.33 | N.A.b | N.A.b | 43.09 | 34.27 | 24.59 | 24.78 | 24.56 |
Li | 5.42 | 5.40 | 5.13 | 4.80 | N.A.b | N.A.b | 11.00 | 9.42 | 5.26 | 5.46 | 5.39 |
Be | 8.52 | 8.73 | N.A.b | 7.84 | N.A.b | N.A.b | 14.09 | N.A.c | 7.74 | 9.00 | 9.31 |
B | 8.41 | 8.62 | 5.32 | 4.04 | N.A.b | N.A.b | 13.54 | N.A.c | 6.88 | 8.73 | 8.26 |
C | 11.51 | 12.24 | 9.86 | 8.22 | 10.01 | 9.87 | 17.23 | N.A.c | 9.87 | 11.51 | 11.23 |
N | 14.90 | 16.41 | 13.28 | 14.09 | 13.63 | 13.42 | 19.56 | 18.91 | 13.79 | 14.58 | 14.52 |
O | 12.17 | 16.67 | 12.76 | 10.54 | 13.43 | 13.41 | 26.17 | 22.39 | 12.87 | 13.91 | 13.51 |
F | 17.99 | 20.91 | 16.39 | 17.25 | 17.06 | 17.29 | 28.58 | 24.84 | 16.90 | 17.44 | 17.34 |
Ne | 22.35 | 25.70 | N.A.b | 5.30 | N.A.b | N.A.b | 24.76 | 31.86 | 21.47 | 21.35 | 21.53 |
MD | 0.02 | 1.51 | −2.17 | −3.38 | −0.68 | −0.75 | 7.61 | 6.90 | −0.72 | 0.11 | |
MAD | 0.45 | 1.63 | 2.17 | 3.38 | 0.68 | 0.75 | 7.61 | 6.90 | 0.72 | 0.22 |
The charge dependence corrections , , , and are set to zero.
The element is not parameterized.
The SCF calculation fails to converge.
Atomic EAs (eV) of elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-Full results. The EAs of He and Ne are omitted since their anions are either not calculable (due to the use of a minimal basis set) or unbound at all levels of theories.
Element . | NOTCH . | NOTCHa . | PM3 . | PM7 . | OM2 . | OM3 . | GFN-xTB . | GFN2-xTB . | HF-3c . | B3LYP-D3/def2-SVP . | CCSD(T) . |
---|---|---|---|---|---|---|---|---|---|---|---|
H | 0.72 | −2.32 | −1.72 | −3.08 | −0.20 | −0.39 | 4.53 | 5.91 | −8.60 | −0.42 | 0.72 |
Li | 0.27 | −0.72 | −2.17 | −14.10 | N.A.b | N.A.b | 5.41 | 2.75 | −1.64 | 0.41 | 0.60 |
Be | −0.61 | −1.79 | N.A.b | −3.95 | N.A.b | N.A.b | 2.54 | −14.87 | −7.05 | −0.82 | −0.73 |
B | 0.32 | −1.54 | −1.89 | 2.39 | N.A.b | N.A.b | 4.07 | −1.93 | −4.70 | −0.30 | 0.14 |
C | 1.43 | −0.74 | 1.70 | −0.63 | 0.79 | 0.65 | 4.74 | 3.37 | −4.53 | 0.38 | 1.24 |
N | −1.27 | −3.07 | 0.58 | −0.03 | −0.74 | −0.95 | 6.30 | 5.86 | −8.21 | −1.71 | −0.24 |
O | 1.87 | −1.64 | 0.98 | 0.99 | 1.22 | 1.20 | 9.65 | 8.20 | −6.23 | −0.39 | 1.40 |
F | 4.03 | −0.14 | 2.17 | 3.36 | 3.05 | 3.28 | 8.85 | 7.90 | −4.78 | 1.21 | 3.37 |
MD | −0.06 | −2.42 | −0.86 | −2.69 | −0.47 | −0.54 | 4.95 | 1.33 | −6.53 | −1.02 | |
MAD | 0.38 | 2.42 | 1.36 | 3.31 | 0.47 | 0.54 | 4.95 | 6.31 | 6.53 | 1.02 |
Element . | NOTCH . | NOTCHa . | PM3 . | PM7 . | OM2 . | OM3 . | GFN-xTB . | GFN2-xTB . | HF-3c . | B3LYP-D3/def2-SVP . | CCSD(T) . |
---|---|---|---|---|---|---|---|---|---|---|---|
H | 0.72 | −2.32 | −1.72 | −3.08 | −0.20 | −0.39 | 4.53 | 5.91 | −8.60 | −0.42 | 0.72 |
Li | 0.27 | −0.72 | −2.17 | −14.10 | N.A.b | N.A.b | 5.41 | 2.75 | −1.64 | 0.41 | 0.60 |
Be | −0.61 | −1.79 | N.A.b | −3.95 | N.A.b | N.A.b | 2.54 | −14.87 | −7.05 | −0.82 | −0.73 |
B | 0.32 | −1.54 | −1.89 | 2.39 | N.A.b | N.A.b | 4.07 | −1.93 | −4.70 | −0.30 | 0.14 |
C | 1.43 | −0.74 | 1.70 | −0.63 | 0.79 | 0.65 | 4.74 | 3.37 | −4.53 | 0.38 | 1.24 |
N | −1.27 | −3.07 | 0.58 | −0.03 | −0.74 | −0.95 | 6.30 | 5.86 | −8.21 | −1.71 | −0.24 |
O | 1.87 | −1.64 | 0.98 | 0.99 | 1.22 | 1.20 | 9.65 | 8.20 | −6.23 | −0.39 | 1.40 |
F | 4.03 | −0.14 | 2.17 | 3.36 | 3.05 | 3.28 | 8.85 | 7.90 | −4.78 | 1.21 | 3.37 |
MD | −0.06 | −2.42 | −0.86 | −2.69 | −0.47 | −0.54 | 4.95 | 1.33 | −6.53 | −1.02 | |
MAD | 0.38 | 2.42 | 1.36 | 3.31 | 0.47 | 0.54 | 4.95 | 6.31 | 6.53 | 1.02 |
The charge dependence corrections , , , and are set to zero.
The element is not parameterized.
B. Diatomic equilibrium bond lengths
Encouraged by the success with atoms, we then move on to diatomic molecules. The first thing to look at is the ground state equilibrium bond lengths and vibrational frequencies of the molecules, which are of paramount importance since geometry optimization and molecular dynamics are perhaps the most common applications of semiempirical methods. As only 55 different diatomic molecules can possibly form between the elements H–Ne, we can do an exhaustive analysis of all these molecules to rule out most qualitative failures of NOTCH on the diatomic level. To the best of our knowledge, such a systematic analysis of diatomic molecules has never been pursued for semiempirical methods.
As can be seen from Fig. 3, the NOTCH equilibrium bond lengths agree very well with the reference values. The largest absolute deviations are on the order of 1 Å and occur in the molecules HeLi and LiNe, but since they are van der Waals complexes with very long interatomic distances [6.043 and 5.112 Å at the CCSD(T) level, respectively; see Table S2 for the complete set of equilibrium geometry benchmark results], such deviations should be tolerable for most purposes. The largest negative deviations are similarly found in van der Waals complexes (BeNe, BNe, CNe), whose equilibrium bond lengths are very long. The largest deviations (>0.05 Å) seen in covalent and ionic bonds are the multiply bonded molecules BeN, NO, and O2 (which have appreciable multireference character), and the lithium compounds LiH, LiBe, LiB, LiN, LiO, and LiF (which are highly ionic).
Errors of equilibrium bond lengths (Å) of all neutral, ground state diatomic molecules formed between the elements H–Ne (left), as well as excluding Be2 and all molecules containing He and/or Ne (right), computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. Important outliers are labeled. Method abbreviations: NOTCH a—NOTCH where the geometry dependence of the basis contraction coefficients are neglected, i.e., ; NOTCH b—NOTCH where αcorr is set to 0; NOTCH c—NOTCH where the dispersion correction is set to 0; NOTCH d—NOTCH where the charge dependence of integrals is neglected, i.e., is set to 0; NOTCH e—NOTCH where the basis set incompleteness correction is neglected, i.e., is set to 0; NOTCH f—NOTCH where the LRC correction is set to 0; B3LYP—B3LYP-D3/def2-SVP.
Errors of equilibrium bond lengths (Å) of all neutral, ground state diatomic molecules formed between the elements H–Ne (left), as well as excluding Be2 and all molecules containing He and/or Ne (right), computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. Important outliers are labeled. Method abbreviations: NOTCH a—NOTCH where the geometry dependence of the basis contraction coefficients are neglected, i.e., ; NOTCH b—NOTCH where αcorr is set to 0; NOTCH c—NOTCH where the dispersion correction is set to 0; NOTCH d—NOTCH where the charge dependence of integrals is neglected, i.e., is set to 0; NOTCH e—NOTCH where the basis set incompleteness correction is neglected, i.e., is set to 0; NOTCH f—NOTCH where the LRC correction is set to 0; B3LYP—B3LYP-D3/def2-SVP.
To illustrate the effects of some key ingredients of NOTCH, we next perform benchmark studies where certain terms in NOTCH are turned off or certain parameters are set to zero. When the geometry dependence of the ano-pc-0 contraction coefficients is neglected (“NOTCH a”), we observe a systematic overestimation of bond lengths, especially for hydrogen-containing molecules; this is consistent with the well-known fact that the optimal hydrogen 1s exponents of minimal basis sets for molecular calculations are significantly larger than the free atom value 1.0 (typically around 1.242), while the optimal exponents of other elements differ significantly less from the corresponding free atom values. Meanwhile, the use of a scaled-down version of the overlap matrix elements in Eq. (27) appears essential for the correct description of a few non-covalent complexes containing He and Ne, as otherwise their bond lengths are severely underestimated and sometimes even end up at covalent bond-like values (e.g., the Li–Ne equilibrium bond length is 1.826 Å when αcorr is set to zero, as in “NOTCH b” in Fig. 3, compared to the reference value 5.112 Å). The dispersion correction Eq. (31) is of course essential for the qualitative description of most He- and Ne-containing molecules as well as Be2, which are under-bound or even unbound when the dispersion correction is turned off (“NOTCH c”); the effect is notable even for some weak covalent bonds, such as Li2 (Fig. 3). The charge dependence of integrals does not in general affect the bond lengths appreciably but is responsible for preventing the overbinding of some He- and Ne-containing molecules when the dependence is neglected (“NOTCH d”), which can undergo charge transfers on the 0.1e magnitude at compressed geometries and are thus susceptible to the treatment of the charge dependence of integrals. Neglecting the basis set incompleteness correction (“NOTCH e”) results in a systematic overestimation of bond lengths, especially for hydrogen-containing molecules and molecules containing the elements N, O and F. Finally, neglecting the LRC correction (“NOTCH f”) gives systematically underestimated bond lengths for hydrogen-containing molecules and F2, consistent with the well-known fact that HF tends to underestimate bond lengths due to the lack of left–right correlation. In general, neglecting the various terms in NOTCH results in either worse outliers, wider error spread or increased systematic error, which illustrates the importance of the terms (Fig. 3).
Subsequently, we test the accuracy of some other representative semiempirical methods, as well as HF-3c and B3LYP-D3/def2-SVP (Table S3). All methods, except for the OM2 and OM3 methods parameterized for the elements H, C, N, O, and F only, have severe outliers with qualitatively wrong bond lengths, as shown in Fig. 3 and Table S2. For example, the PM3 method systematically overestimates the bond lengths of Li- and B-containing molecules by as much as 0.452 and 0.210 Å, respectively. The PM7, GFN-xTB, and GFN2-xTB methods, as well as HF-3c and B3LYP-D3/def2-SVP, severely underestimate the bond lengths of several He- and Ne-containing molecules. In particular, the PM7 and GFN2-xTB methods predict that the bond lengths of some He- and Ne-containing molecules are even shorter than typical covalent bonds, for example, HeO [R(GFN2-xTB) = 0.808 Å], BNe [R(GFN2-xTB) = 1.213 Å], HNe [R(PM7) = 0.970 Å], and BeNe [R(PM7) = 0.743 Å]. Even the B3LYP-D3/def2-SVP method, which is widely used in production level geometry optimizations, fails dramatically for LiNe, predicting a bond length of 2.677 Å, which is even shorter than the Li–Li bond length obtained at the same level, 2.766 Å. The widespread failure of semiempirical and cost-effective ab initio methods (especially B3LYP-D3/def2-SVP) for these molecules can be partly attributed to basis set superposition error (BSSE), although the aforementioned failures of GFN2-xTB and PM7 for HeO, BNe, HNe, and BeNe must be primarily due to other deficiencies of the methods, given the magnitudes of the deviations. While the latter deficiencies presumably do not show up in non-covalent complexes in the training sets of the methods, they lead to spurious underestimations of energy near the typical interatomic distance of covalent bonds, a region that is scarcely probed by most He- and/or Ne-containing systems in typical training sets. By contrast, the use of less empirical and more physically grounded constructions in the NOTCH method helped us to avoid unprecedented failures of the method at extremely short bond distances, and the explicit consideration of the charge dependence of integrals (which reflects the radial flexibility of the basis set) makes the NOTCH AOs behave like a much more complete basis set than it would otherwise be (as evidenced by the failure of NOTCH for He- and Ne-containing molecules when is set to zero), therefore reducing the BSSE. Finally, even when non-covalent molecules are excluded from the statistics, NOTCH continues to have a narrower error spread than PM3, PM7, GFN-xTB, GFN2-xTB, and HF-3c (Fig. 3, right), which shows that NOTCH is systematically superior for covalently bonded molecules as well. The excellent performance of OM2 and OM3 can be explained by the fact that they are not parameterized for the elements He, Li, Be, B, and Ne, which are generally “difficult” elements for bond lengths, as evidenced by the outliers shown in Fig. 3.
C. Diatomic vibrational frequencies
The errors of harmonic vibrational frequencies (Fig. 4; for the complete set of data, see Table S4) generally mirror the trend of the equilibrium bond lengths, with the frequencies tending to be overestimated whenever the equilibrium bond length is underestimated, and vice versa. However, as NOTCH has a systematic tendency to overestimate vibrational frequencies, it may appear that some of the terms in NOTCH, such as the basis set incompleteness correction , actually worsen the agreement of the vibrational frequencies with the reference values (cf. the “NOTCH e” result in Fig. 4, where is set to 0); nevertheless, the effect is not very large (on the order of 100 cm−1) and may thus be viewed as a necessary compromise for obtaining satisfactory equilibrium bond lengths. Other terms generally improve the accuracy of the vibrational frequencies in most cases. Neglecting the geometry dependence of the AO contraction coefficients (“NOTCH a”) leads to too small vibrational frequencies for H2 and increases the overestimation of vibrational frequencies for period 2 elements. Setting αcorr to zero (“NOTCH b”) worsens the results of Li- and Be-containing molecules, again reflecting the role of Eq. (27) in preventing the overbinding of weakly bound molecules. Neglecting the charge dependence of integrals (“NOTCH d”) results in too large vibrational frequencies for highly polar molecules like HF and LiF. Finally, the LRC correction (“NOTCH f”) is found to play a major role in the correct prediction of vibrational frequencies of hydrogen-containing molecules as well as O2 and F2.
Errors of harmonic vibrational frequencies at equilibrium geometries (cm−1) of all neutral, ground state diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. Important outliers are labeled. For the meanings of the method abbreviations, please refer to Fig. 3.
Errors of harmonic vibrational frequencies at equilibrium geometries (cm−1) of all neutral, ground state diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. Important outliers are labeled. For the meanings of the method abbreviations, please refer to Fig. 3.
Similar to the case of equilibrium geometries, the vibrational frequencies computed by other semiempirical methods and HF-3c (Table S5) invariably show severe outliers. In addition to the outliers that are already expectable from the bad equilibrium geometries (such as the HeO, BNe, HNe, and BeNe molecules mentioned above), there are a few more outliers, like the BN and O2 molecules at the PM3 level and the H2, HO, and F2 molecules at the OM2 level. One may notice that many of these molecules have significant multireference character, such that single reference methods are prone to yield erroneous results. However, the much smaller errors of NOTCH compared to other semiempirical methods as well as the fact that the B3LYP-D3/def2-SVP frequencies of these multireference molecules are not much worse than the single-reference ones suggest that the failure of the existing semiempirical methods is not an intrinsic problem of the single-reference theoretical framework but rather a problem of the functional forms and parameterization procedures of the methods.
D. Diatomic dipole moments
Errors of signed dipole moments (Debye) of all neutral, ground state heterodiatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results (evaluated on the linearized coupled cluster densities). The sign convention is chosen such that the signs of all DKH2 CCSD(T)/ANO-RCC-TZP dipole moments are positive. Important outliers are labeled. “NOTCH + p” stands for NOTCH dipole moments that include the contribution from Eq. (51); other results do not include the contribution from Eq. (51). For the meanings of the other method abbreviations, please refer to Fig. 3.
Errors of signed dipole moments (Debye) of all neutral, ground state heterodiatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results (evaluated on the linearized coupled cluster densities). The sign convention is chosen such that the signs of all DKH2 CCSD(T)/ANO-RCC-TZP dipole moments are positive. Important outliers are labeled. “NOTCH + p” stands for NOTCH dipole moments that include the contribution from Eq. (51); other results do not include the contribution from Eq. (51). For the meanings of the other method abbreviations, please refer to Fig. 3.
Despite this, many things can still be said about the influence of the various terms on the dipole moment. The error spread of the dipole moments is worsened when the geometry dependence of the ano-pc-0 contraction coefficients is neglected (“NOTCH a”), suggesting that a correct description of the radial extent of the AOs is crucial for the accuracy of dipole moments (Fig. 5). Neglecting the charge dependence of integrals (“NOTCH d”) improves the dipole moments of some boron compounds but worsens the dipole moments of lithium compounds; as the boron compounds generally have strong multireference character, we believe that the former improvement is fortuitous. The basis set incompleteness correction is likewise found to slightly deteriorate the dipole moment results for the molecules formed between the elements B, C, and N but improve the results of the hydrides of B, C, N, O, and F (Table S6). Finally, removing the LRC correction (“NOTCH f”) deteriorates the results of molecules like NO, NF, and OF but also improves the dipole moments of some other molecules, notably LiF (Table S6). Overall, many of the terms in NOTCH are not generally beneficial for the accuracy of dipole moments, unlike the case of geometries and vibrational frequencies.
Nevertheless, the NOTCH dipole moments still prove much more robust than the dipole moments calculated by many semiempirical methods and HF-3c (Table S7). The dipole moments of GFN-xTB, GFN2-xTB, and HF-3c have rather large outliers for uncommon bonds, especially Li- and Be-containing molecules; PM7 additionally performs exceptionally poorly for Ne-containing molecules, which can be traced to its strong underestimation of the IP of Ne (Table III). The PM3, OM2, and OM3 methods give accuracies that are similar to or better than NOTCH, but this may partly be attributed to the fact that beryllium is not parameterized in these methods, which seems to be a difficult case for methods that do parameterize it. Finally, B3LYP-D3/def2-SVP gives excellent predictions for most molecules, with the exception of LiBe, LiC, LiN, LiO, and LiF whose dipole moments are significantly underestimated as well as LiB and BeC whose dipole moments are overestimated (obviously owing to their multireference character). The former suggests that the errors of the NOTCH dipole moments of lithium-containing molecules (Table S6), while large in magnitude, are already remarkably small when compared with other methods. Overall, NOTCH has a smaller error range than all the studied semiempirical methods and HF-3c, except for OM2 and OM3, which are only parameterized for H, C, N, O, and F (Fig. 5).
E. Diatomic IPs
Our next target is the IPs of diatomic molecules. As can be seen from Fig. 6 and Table S8, the errors of the NOTCH IPs generally lie within 1.5 eV, but there are a few positive outliers, namely a few molecules formed between B, C, N, and O, as well as a single negative outlier, F2. All these molecules have significant multireference character, and while one may expect that the LRC correction may play a significant role here, it is evident from Fig. 6 that the LRC correction overcorrects F2 and under-corrects most of the other multireference molecules. In general, however, the LRC correction proves to be beneficial for reducing the largest IP errors. The geometry dependence of the AO contraction coefficient also helps to reduce the systematic overestimation of the IPs, especially for molecules containing carbon and/or nitrogen. The parameter αcorr and the charge dependence of integrals do not contribute as much to the IPs although the former reduces the IPs of lithium-containing molecules by about 0.1–0.5 eV and thus helps to reduce the error.
Errors of vertical IPs (eV) of all neutral, ground state diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. The DKH2 CCSD(T)/ANO-RCC-TZP equilibrium geometries are used. Important outliers are labeled. For the meanings of the method abbreviations, please refer to Fig. 3.
Errors of vertical IPs (eV) of all neutral, ground state diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. The DKH2 CCSD(T)/ANO-RCC-TZP equilibrium geometries are used. Important outliers are labeled. For the meanings of the method abbreviations, please refer to Fig. 3.
The performance of other methods is then briefly discussed (Table S9). The GFN-xTB and GFN2-xTB methods drastically overestimate all the IPs by about 2–10 eV with a large scatter, consistent with what is observed in atoms (Table III). The PM7 and HF-3c methods, on the other hand, systematically underestimate the IPs; while the former is well expected from the large negative mean deviation (MD) of the atomic IPs of PM7 (Table III), the latter is better explained by the tendency of HF-3c to significantly underestimate atomic EAs (Table IV), which make the molecular IPs too low when one of the atoms in the molecules carries significant negative charge (e.g., the molecules LiN, LiO and LiF in Table S9). The PM3, OM2, and OM3 methods perform reasonably well and, in particular, are free from the systematic overestimation of the IPs in B-, C-, N-, and O-containing molecules as found in NOTCH. Whether the latter is physically meaningful or due to error cancellation is unclear at the moment. Finally, we note that the B3LYP-D3/def2-SVP method, while performing excellently for all covalently bonded molecules, underestimates the IPs of He2 and Ne2 by large margins, despite NOTCH performing very well for these systems; this is clearly due to the self-interaction error of B3LYP, which spuriously stabilizes stretched radical ions like He2+ and Ne2+ and illustrates the importance of using 100% HF exchange in the NOTCH method.
F. Diatomic EAs
The EAs of diatomic molecules computed under different levels of theories are summarized in Fig. 7 and Table S10; note that molecules whose anions are unbound at the reference level of theory are excluded from the analysis. The accuracy of the NOTCH EAs is similar to that of the IPs; there are a few outliers (Li2, F2, and boron-containing molecules) that can probably be attributed to some problems of the LRC correction. A few other contributions, such as the geometry dependence of the AO contraction coefficients, the parameter αcorr, and the charge dependence of integrals, prove notably beneficial for the accuracy of the EAs, indicating the crucial role of the radial flexibility of the AOs (and to a lesser extent, the correct treatment of correlation) in the correct prediction of the EAs.
Errors of vertical EAs (eV) of all neutral, ground state diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. The DKH2 CCSD(T)/ANO-RCC-TZP equilibrium geometries are used. Molecules whose DKH2 CCSD(T)/ANO-RCC-TZP EAs are negative are omitted. Important outliers are labeled. For the meanings of the method abbreviations, please refer to Fig. 3.
Errors of vertical EAs (eV) of all neutral, ground state diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. The DKH2 CCSD(T)/ANO-RCC-TZP equilibrium geometries are used. Molecules whose DKH2 CCSD(T)/ANO-RCC-TZP EAs are negative are omitted. Important outliers are labeled. For the meanings of the method abbreviations, please refer to Fig. 3.
Most of the other methods studied herein (Table S11) have difficulties predicting the EAs of some of the diatomic molecules. For example, PM3 predicts much too low EAs for LiB, B2, C2, and F2, while PM7 gives systematically too low EAs for most hydrogen- and lithium-containing molecules, except for Li2 where the EA is significantly overestimated. The OM2 and OM3 methods perform well, with just a single outlier, F2; however this may be partially due to the fact that the more difficult elements Li, Be, and B are not parameterized in these methods. The EAs are systematically overestimated at the GFN-xTB and GFN2-xTB levels, except for some beryllium-containing molecules, whose EAs are predicted to be too negative at the GFN2-xTB level, in accord with the drastically underestimated EA of the beryllium atom (Table IV). The HF-3c and, to a much less extent, the B3LYP-D3/def2-SVP methods predict systematically too low EAs, which can be obviously attributed to the use of a too small basis set and the lack of corrections for the lack of diffuse functions.
G. Diatomic bond dissociation energies
The final ground state property to look at is the bond dissociation energy. As shown in Fig. 8 and Table S12, NOTCH gives bond dissociation energies that are usually accurate to about 1 eV; the worse outliers (BeF, F2, BF, N2) generally possess some multireference character. The terms that affect the NOTCH bond dissociation energies the most are the scaled overlap matrix elements [Eq. (27)] and the LRC correction [Eq. (43)]. When αcorr is set to zero (“NOTCH b”), the bond dissociation energies are overestimated by about 1 eV on average and feature a broader error distribution due to overestimation of dynamic correlation. Interestingly, removing the LRC correction (“NOTCH f”) actually improves the bond dissociation energies (the MAD is reduced from 0.813 to 0.485 eV; see Table S12). The latter suggests that the LRC correction may still have room for improvement.
Errors of bond dissociation energies (eV) of all neutral, ground state diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. The equilibrium geometries optimized at the respective levels of theory are used. Important outliers are labeled. For the meanings of the method abbreviations, please refer to Fig. 3.
Errors of bond dissociation energies (eV) of all neutral, ground state diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical and cost-effective ab initio methods, compared with DKH2 CCSD(T)/ANO-RCC-TZP results. The equilibrium geometries optimized at the respective levels of theory are used. Important outliers are labeled. For the meanings of the method abbreviations, please refer to Fig. 3.
The bond dissociation energies of other semiempirical methods and HF-3c (Table S13) have broader error distributions than NOTCH does, again with the exception of OM2 and OM3, which are parameterized for H, C, N, O, and F only. Naturally, the biggest outliers are typically molecules possessing multireference character, like BO, C2, and N2. The PM3 and HF-3c methods, bearing 100% HF exchange, tend to underbind these multiply bonded, multireferential molecules, while the GFN-xTB and GFN2-xTB methods, which do not have HF exchange, typically overbind them. The PM7 method manages to achieve less systematic bias throughout the set of molecules studied, probably due to it being more extensively parameterized; however, like what we have seen in Sec. III B, this comes at a price that some non-covalent complexes (HeB, BeNe) are predicted to be extremely stable, covalently bonded molecules, with dissociation energies (9.246 and 8.594 eV, respectively) even exceeding that of N2 at the same level of theory (8.369 eV). B3LYP-D3/def2-SVP performs excellently for the bond dissociation energies, as expected.
H. Diatomic excitation energies
The excitation energy benchmark results of NOTCH, INDO/S, PM3, and OM2 are listed in Table V. While NOTCH performs very well for π → π* excitations (e.g., N2 and CO) and some σ → π* excitations [e.g., Li2 (3Πu), HN (3Π), BC (4Π), BF, LiC (4Π), and BeF (2Π)], giving an error of within 1 eV in most cases, it systematically overestimates the energies of excitations into σ* orbitals [e.g., H2 (σ → σ*), F2 (π* → σ*), HN (5Σ−, σ → σ*), HF (π → σ*), LiF (π* → σ*), BeF (2Σ+, σ → σ*), BO (2Π, π → σ*, and 2Σ+, σ → σ*), and CN (2Π, π → σ*, and 2Σ+, σ → σ*)]. Neglecting the geometry dependence of the ano-pc-0 basis set improves the excitation energies of hydrogen-containing molecules by as much as 5 eV, suggesting that optimizing the AO basis set for the ground state leads to H 1s orbitals that are too compact for expanding the σ* orbital. However, the fact that a sizable error still remains even if the (very diffuse) free atom AOs are used and that the excitation energies of F2, LiF, BeF, and BO are not much affected by this change suggests that there is probably no minimal basis set that can provide a reasonable compromise between describing the ground state properties and the σ* orbital. Turning off the charge dependence of the integrals does not in general lead to appreciable changes of the excitation energies but changes the excitation energies of highly polar molecules (e.g., HF, LiF) by up to 1.6 eV; while the changes do not always improve the results, it can be said that the effect of the charge dependence of integrals on the excitation energies is large. The LRC correction is occasionally useful in curing the underestimation of triplet excitation energies (N2), which in extreme cases can lead to triplet instabilities (O2, F2), as already discussed in Sec. II H.
The lowest two vertical excitation energies (eV) of neutral, diatomic molecules formed between the elements H–Ne, computed by NOTCH as well as other semiempirical methods via RPA, compared with DKH2 MRACPF/ANO-RCC-TZP results. Only those molecules that do not contain rare gas atoms, whose ground states are spatially nondegenerate, and whose ground state NOTCH wavefunctions do not have instabilities are listed. The DKH2 CCSD(T)/ANO-RCC-TZP equilibrium geometries are used.
Molecule . | State . | NOTCH . | NOTCH a . | NOTCH b . | NOTCH c . | INDO/S . | PM3 . | OM2 . | MRACPF . |
---|---|---|---|---|---|---|---|---|---|
H2 | 3Σu+ | 18.305 | 15.458 | 18.305 | 15.782 | 13.700 | 6.926 | 5.066 | 10.585 |
1Σu+ | 22.371 | 17.463 | 24.670 | 21.518 | 18.317 | 10.006 | 7.236 | 12.732 | |
Li2 | 3Σu+ | 1.432 | 1.600 | 1.432 | 1.212 | d | d | e | 1.304 |
3Πu | 1.833 | 1.153 | 1.833 | 1.041 | d | d | 1.384 | ||
N2 | 3Σu+ | 7.613 | 7.820 | 7.613 | 6.219 | 3.266 | 5.261 | 5.951 | 7.569 |
3Πg | 7.940 | 7.220 | 7.940 | 7.425 | 4.979 | 5.615 | 6.500 | 7.935 | |
O2 | 3Δu | 6.503 | 6.673 | 6.503 | d | 2.311 | 4.641 | 6.009f | 6.145 |
3Σu+ | 4.045 | 3.171 | 5.757 | d | 2.311 | 4.694 | 6.053f | 6.292 | |
F2 | 3Πu | 6.178 | 6.134 | 6.178 | d | 7.978 | 3.916 | 2.298 | 3.152 |
1Πu | 8.405 | 8.353 | 8.405 | 5.641 | 9.171 | 4.206 | 3.300 | 4.347 | |
HLi | 3Σ+ | 5.394 | 5.501 | 5.403 | 3.936 | 4.937 | 3.297 | e | 3.263 |
1Σ+ | 5.667 | 5.585 | 5.740 | 4.814 | 5.426 | 4.863 | 3.629 | ||
HN | 3Π | 4.207 | 3.632 | 4.208 | 3.968 | 3.976 | 2.876 | 3.959f | 3.701 |
5Σ- | 17.950 | 13.783 | 17.912 | 16.182 | 12.266 | 6.207 | 7.137f | 8.416 | |
HF | 3Π | 16.357 | 15.197 | 17.324 | 19.024 | 14.519 | 8.266 | 9.667 | 10.200 |
1Π | 17.133 | 16.320 | 18.790 | 20.258 | 15.737 | 8.632 | 10.355 | 10.569 | |
LiC | 4Π | 4.445 | 4.763 | 4.311 | 4.462 | 3.442 | 4.102 | e | 3.254 |
4Σ- | 5.156 | 5.490 | 5.041 | 5.172 | 3.718 | 4.293 | 3.778 | ||
LiF | 3Π | 8.570 | 8.956 | 7.557 | 9.381 | 9.502 | 7.854 | e | 6.734 |
1Π | 8.790 | 9.100 | 7.809 | 9.542 | 9.568 | 7.987 | 6.778 | ||
BeF | 2Π | 5.508 | 5.335 | 5.540 | 5.179 | 3.888 | e | e | 4.119 |
2Σ+ | 11.929 | 12.632 | 12.761 | 12.878 | 10.416 | 8.218 | |||
BC | 4Π | 0.360 | 0.316 | d | 0.577 | 2.437 | 3.997 | e | 1.509 |
4Σ- | 4.430 | 4.541 | d | 4.110 | 2.566 | 4.419 | 2.260 | ||
BO | 2Π | 8.172 | 8.003 | 7.684 | 8.131 | 3.339 | 2.981 | e | 3.496 |
2Σ+ | 11.040 | 9.395 | 8.181 | 8.699 | 4.777 | 6.366 | 5.561 | ||
BF | 3Π | 2.638 | 2.521 | 2.464 | 2.134 | 3.216 | 5.073 | e | 3.652 |
1Π | 6.892 | 6.663 | 6.825 | 6.312 | 6.102 | 7.383 | 6.521 | ||
CN | 2Π | 4.596 | 4.649 | 4.295 | 4.696 | 0.886 | 1.392 | 2.209f | 1.329 |
2Σ+ | 5.800 | 5.665 | 5.454 | 5.597 | 1.962 | 2.170 | 4.541f | 3.195 | |
CO | 3Π | 5.789 | 5.269 | 5.965 | 5.389 | 4.975 | 4.771 | 5.040 | 6.191 |
3Σ+ | 8.808 | 8.388 | 8.874 | 8.400 | 6.053 | 6.626 | 7.165 | 8.326 | |
NF | 3Π | 8.664 | 8.631 | 8.668 | 6.996 | 8.247 | 5.004 | 3.803f | 7.107 |
3Π | 10.564 | 10.396 | 10.586 | 9.358 | 10.793 | 5.085 | 6.031f | 7.264 | |
MDg | 2.440 | 1.919 | 2.603 | 2.229 | 0.842 | −0.800 | −1.388 | ||
MADg | 2.723 | 2.354 | 2.724 | 2.608 | 2.238 | 1.672 | 1.649 | ||
MDg,h | 1.450 | 1.299 | 1.348 | 1.005 | 0.200 | −0.101 | −0.711 | ||
MADg,h | 1.807 | 1.845 | 1.504 | 1.494 | 1.804 | 1.228 | 1.125 |
Molecule . | State . | NOTCH . | NOTCH a . | NOTCH b . | NOTCH c . | INDO/S . | PM3 . | OM2 . | MRACPF . |
---|---|---|---|---|---|---|---|---|---|
H2 | 3Σu+ | 18.305 | 15.458 | 18.305 | 15.782 | 13.700 | 6.926 | 5.066 | 10.585 |
1Σu+ | 22.371 | 17.463 | 24.670 | 21.518 | 18.317 | 10.006 | 7.236 | 12.732 | |
Li2 | 3Σu+ | 1.432 | 1.600 | 1.432 | 1.212 | d | d | e | 1.304 |
3Πu | 1.833 | 1.153 | 1.833 | 1.041 | d | d | 1.384 | ||
N2 | 3Σu+ | 7.613 | 7.820 | 7.613 | 6.219 | 3.266 | 5.261 | 5.951 | 7.569 |
3Πg | 7.940 | 7.220 | 7.940 | 7.425 | 4.979 | 5.615 | 6.500 | 7.935 | |
O2 | 3Δu | 6.503 | 6.673 | 6.503 | d | 2.311 | 4.641 | 6.009f | 6.145 |
3Σu+ | 4.045 | 3.171 | 5.757 | d | 2.311 | 4.694 | 6.053f | 6.292 | |
F2 | 3Πu | 6.178 | 6.134 | 6.178 | d | 7.978 | 3.916 | 2.298 | 3.152 |
1Πu | 8.405 | 8.353 | 8.405 | 5.641 | 9.171 | 4.206 | 3.300 | 4.347 | |
HLi | 3Σ+ | 5.394 | 5.501 | 5.403 | 3.936 | 4.937 | 3.297 | e | 3.263 |
1Σ+ | 5.667 | 5.585 | 5.740 | 4.814 | 5.426 | 4.863 | 3.629 | ||
HN | 3Π | 4.207 | 3.632 | 4.208 | 3.968 | 3.976 | 2.876 | 3.959f | 3.701 |
5Σ- | 17.950 | 13.783 | 17.912 | 16.182 | 12.266 | 6.207 | 7.137f | 8.416 | |
HF | 3Π | 16.357 | 15.197 | 17.324 | 19.024 | 14.519 | 8.266 | 9.667 | 10.200 |
1Π | 17.133 | 16.320 | 18.790 | 20.258 | 15.737 | 8.632 | 10.355 | 10.569 | |
LiC | 4Π | 4.445 | 4.763 | 4.311 | 4.462 | 3.442 | 4.102 | e | 3.254 |
4Σ- | 5.156 | 5.490 | 5.041 | 5.172 | 3.718 | 4.293 | 3.778 | ||
LiF | 3Π | 8.570 | 8.956 | 7.557 | 9.381 | 9.502 | 7.854 | e | 6.734 |
1Π | 8.790 | 9.100 | 7.809 | 9.542 | 9.568 | 7.987 | 6.778 | ||
BeF | 2Π | 5.508 | 5.335 | 5.540 | 5.179 | 3.888 | e | e | 4.119 |
2Σ+ | 11.929 | 12.632 | 12.761 | 12.878 | 10.416 | 8.218 | |||
BC | 4Π | 0.360 | 0.316 | d | 0.577 | 2.437 | 3.997 | e | 1.509 |
4Σ- | 4.430 | 4.541 | d | 4.110 | 2.566 | 4.419 | 2.260 | ||
BO | 2Π | 8.172 | 8.003 | 7.684 | 8.131 | 3.339 | 2.981 | e | 3.496 |
2Σ+ | 11.040 | 9.395 | 8.181 | 8.699 | 4.777 | 6.366 | 5.561 | ||
BF | 3Π | 2.638 | 2.521 | 2.464 | 2.134 | 3.216 | 5.073 | e | 3.652 |
1Π | 6.892 | 6.663 | 6.825 | 6.312 | 6.102 | 7.383 | 6.521 | ||
CN | 2Π | 4.596 | 4.649 | 4.295 | 4.696 | 0.886 | 1.392 | 2.209f | 1.329 |
2Σ+ | 5.800 | 5.665 | 5.454 | 5.597 | 1.962 | 2.170 | 4.541f | 3.195 | |
CO | 3Π | 5.789 | 5.269 | 5.965 | 5.389 | 4.975 | 4.771 | 5.040 | 6.191 |
3Σ+ | 8.808 | 8.388 | 8.874 | 8.400 | 6.053 | 6.626 | 7.165 | 8.326 | |
NF | 3Π | 8.664 | 8.631 | 8.668 | 6.996 | 8.247 | 5.004 | 3.803f | 7.107 |
3Π | 10.564 | 10.396 | 10.586 | 9.358 | 10.793 | 5.085 | 6.031f | 7.264 | |
MDg | 2.440 | 1.919 | 2.603 | 2.229 | 0.842 | −0.800 | −1.388 | ||
MADg | 2.723 | 2.354 | 2.724 | 2.608 | 2.238 | 1.672 | 1.649 | ||
MDg,h | 1.450 | 1.299 | 1.348 | 1.005 | 0.200 | −0.101 | −0.711 | ||
MADg,h | 1.807 | 1.845 | 1.504 | 1.494 | 1.804 | 1.228 | 1.125 |
The geometry dependence of the AO contraction coefficients is neglected, i.e., .
The charge dependence of integrals is neglected, i.e., .
The LRC correction is neglected.
The excitation energy cannot be calculated because the molecule has a triplet instability.
The molecule contains an unparameterized element.
A valence space CASSCF calculation is performed instead of RPA since the MNDO program does not support RPA calculations from an unrestricted reference.
Molecules that possess triplet instabilities or unparameterized elements are not included in the statistics.
States whose excitation energies are above 8 eV are excluded in the statistics.
By comparison, the other semiempirical methods INDO/S, PM3 and (to a lesser extent) OM2 generally underestimate the π → π* excitation energies of N2, O2, and CO, which is a well-known problem of the ZDO formalism itself. In ZDO methods, the bonding and antibonding MOs formed between two AOs are split symmetrically, while in reality the virtual orbitals are higher than what would be expected from a symmetric splitting;51 while this is partially remedied by the orthogonalization corrections in the OMx methods,54 the corrections are only carried out to low order. Thus, while these methods generally perform better than NOTCH for excitations involving σ* orbitals, it is reasonable to believe that the better performance is due to a fortunate cancellation of the negative errors due to the ZDO formalism and the positive errors due to using a minimal basis set. In some cases, the INDO/S, PM3, and OM2 methods significantly underestimate the excitation energies (e.g., H2 at the OM2 level, and NF at the PM3 and OM2 levels), sometimes even resulting in triplet instabilities (e.g., Li2 at the INDO/S and PM3 levels), when one would predict an overestimation of excitation energies due to the use of a minimal basis set. Consequently, despite that the largest outliers occur in NOTCH instead of the other semiempirical methods, we believe that NOTCH should probably be more robust than the other methods for realistically sized organic and organometallic molecules, where the virtual orbitals are less diffuse and much easier to describe by a minimal basis than in diatomic molecules. Moreover, if one desires, it should be relatively easy to improve NOTCH for excited states involving σ* orbitals, since the cause of the overestimation of the excitation energies is well understood.
IV. SUMMARY AND OUTLOOK
The present work describes the preliminary formulation and benchmark results of a new semiempirical method, NOTCH. The method can be considered as a parameterized all-electron HF/minimal basis set calculation, where (contrary to most ZDO methods) all two-center two-electron integrals are retained and all scalar relativity, correlation (local dynamic correlation, dispersion, left–right correlation) and basis set incompleteness (the lack of radial and angular flexibility of the basis set) effects are accounted for by individual, physically motivated corrections.
In our opinion, the theory is based on well-designed and physically appealing approximations and only introduce eight semiempirical parameters which are all global rather than element or even element pair specific. This level of empiricism is even lower than that of many density functionals. Preliminary equilibrium geometry, vibrational frequency, dipole moment, IP, EA, bond dissociation energy, and excitation energy calculations on the atoms and diatomic molecules of the elements H–Ne show that NOTCH is exceptionally robust, with a lack of serious and practically relevant outliers that plague almost all the other semiempirical methods as well as a few cost-effective ab initio methods (e.g., HF-3c), such as the formation of unreasonably short covalent bonds in non-covalent complexes or IPs and EAs that are in error in excess of 5 eV. Meanwhile, the accuracy of NOTCH for the more “conventional” systems, where all methods perform reasonably well, is at least comparable with (and sometimes better than) other semiempirical methods. Moreover, in contrast to other semiempirical methods, NOTCH is balanced in the sense that the quality of the results is consistent across many different properties including the ground state as well as excited states. This feature makes us optimistic that in NOTCH one obtains the right answer for the right reasons. Taken together, the numerical results provide support for the hope that, once fully developed for polyatomic molecules, NOTCH will be useful as a robust, reliable, general-purpose semiempirical method.
While multiple ideas have contributed to the development of NOTCH, we believe that one of the most important ingredients is the parameterization of the one-center one-electron and two-electron integrals on the basis of MR-EOM reference calculations on free atoms and ions, without involving any fitting. The leading idea to approach the design of a semiempirical method from this angle is rooted in the hypothesis that, starting from a good description of the atoms and their excitation and ionization events would be a solid basis for embedding these atoms in a reasonable molecular potential, thus leading to systematically correct electronic structure descriptions. The MR-EOM construction appeared to be particularly promising in this respect given its excellent accuracy relative to experimental data and the fact that the sequence of similarity transformations leads to a form of the theory that can be expressed in terms of “dressed” integrals that can be directly taken as a basis for the parameterization. Thus, MR-EOM does not only provide reliable integrals for atomic calculations, but it also helped us tremendously to avoid large-scale fitting procedures that might have led to overfitting and fortuitously good-looking molecular results.
Insisting on not further manipulating the atomic integrals, we were then forced to identify the real physical reasons for the discrepancies of the calculated results with the reference data, and design physically motivated corrections whose functional forms are genuinely original and have never been proposed or used in existing semiempirical methods. In particular, the geometry dependence of the basis set (Sec. II C), the basis set incompleteness correction (Sec. II G), and the left–right correlation correction (Sec. II H) describe effects that most semiempirical methods do not explicitly consider. In our opinion, this is far more physically satisfying than trying to absorb all of these effects into a large number of parameters that were originally designed to capture other effects. For example, modifying the parameters that define the core–core repulsion potential to absorb, e.g., correlation errors and basis set limitations is particularly dissatisfying to us. While many semiempirical methods incorporate more and more flexible forms of the core–core repulsion potential as they are developed, we have taken the opposite route, i.e., to describe the core–core repulsion exactly using an all-electron approach. Consequently, there is no excuse for adding corrections to the core–core repulsion energy (because it is already exact), and we were forced to develop our other approximations in a way that retains the balance with this “electrostatic cornerstone.” This is clearly a more difficult task, but to us it appears to be more physically appealing than the traditional alternatives.
The obvious next step in the development is the extension of NOTCH to multiatomic molecules. In multiatomic molecules, there are a few issues that are not present in diatomic molecules and that must be properly taken care of:
We would like to omit, to the largest extent possible, three- and four-center two-electron integrals in order to keep the computational performance of NOTCH close to existing semiempirical methods. While simply neglecting these integrals is an unacceptably crude approximation under the AO basis, it is far better justified in the OAO basis.78 If all surviving two-electron integrals can be conveniently derived from one-center and two-center two-electron integrals, then for sufficiently large systems, NOTCH will be at most, say, 2–5 times more expensive than existing semiempirical methods. The additional cost stems partly from the inclusion of the (AA|AB)- and (AB|AB)-type integrals and partly because of the explicit treatment of core electrons.
The and coefficients of a pair of atoms depend on the chemical environments of the atoms. It remains to be seen if ideas from established dispersion correction methods, such as expressing and as functions of the atomic coordination numbers (as used in the DFT-D360 and DFT-D461 methods), can give useful accuracy in conjunction with our newly proposed dispersion correction formalism, Eq. (31).
Of course, analytic gradients and perhaps also analytic Hessians must be implemented such that geometry optimization, frequency analysis, and molecular dynamics simulations of realistically sized molecules become feasible.
Our hope is to be able to extend NOTCH to heavier elements as well. In particular, open shell transition metals appear to be a major challenge for most existing semiempirical methods. Since NOTCH is based on a formulation in which all terms are in balance by construction, and the exchange is treated in a detailed and physically correct fashion without introducing any self-interaction error, we believe that it holds significant promise in the area. A particularly exciting possibility is to combine NOTCH with a minimal active space CASSCF or MR-CIS approach, which is essential for correctly describing the spin multiplet structure of transition metal complexes. When active spaces spanning more than one atom have to be used, one can turn off the contribution of these atoms to the LRC correction in the multireference NOTCH calculation, to avoid double-counting the non-dynamic correlation between these atoms while still retaining the dynamic correlation contribution. This is a key advantage compared to most existing semiempirical methods, which do not separate the non-dynamic and dynamic contributions to correlation.
Eventually, we believe that NOTCH will become a robust, accurate, general-purpose semiempirical method, which offers intermediate cost and accuracy between existing semiempirical methods and DFT methods that allows for a very wide range of applications in (bio)chemistry, material science, and spectroscopy/photochemistry. Our efforts along this way will be reported in due course.
DEDICATION
This paper is dedicated to the memory of Professor Walter Thiel. Walter was a wonderful and kind friend and mentor. One of us (FN) had the privilege to discuss the very early stages of this project with him shortly before his untimely passing. We deeply regret that he could not see this work to fruition, but we hope that this work may help to reinvigorate interest in semiempirical quantum chemistry, a subject that always remained close to Walter’s heart.
SUPPLEMENTARY MATERIAL
See the supplementary material for further computational details of the benchmark studies, derivations of the NOTCH Fock matrix elements and orbital Hessian matrix elements, raw benchmark data, and lists of the nonempirical NOTCH parameters.
ACKNOWLEDGMENTS
The authors gratefully acknowledge generous financial support of this work by the Max Planck society.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Zikuan Wang: Formal analysis (lead); Investigation (lead); Methodology (equal); Software (lead); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (equal). Frank Neese: Conceptualization (lead); Funding acquisition (lead); Methodology (equal); Software (supporting); Supervision (lead); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available within the article and its supplementary material.
APPENDIX A: DERIVATION OF THE ano-pc-0 CONTRACTION COEFFICIENTS AS A FUNCTION OF THE MOLECULAR GEOMETRY
As mentioned in the Introduction, the AOs in molecules tend to be more compact than those in the corresponding free atoms. Two possibly predominant reasons are (a) when the length of a bond A–B is sufficiently short, the AOs of atom A can enjoy more nuclear attraction from atom B by shrinking and (b) the AOs of atom A may overlap more favorably with the AOs of atom B after shrinking, resulting in stronger bonding. While (a) does not necessarily dominate over (b), it is much easier to derive a formula for the contribution (a) than for (b). We have therefore decided to use a functional form that is consistent with (a) and contains a few adjustable parameters and fit the parameters against optimized molecular basis set contraction coefficients so that the effects of contribution (b) are effectively accounted for.
APPENDIX B: FITTING THE NOTCH PARAMETERS αcorr, s8, a1, AND a2 AGAINST DLPNO-CCSD(T) LED DISPERSION ENERGIES
The dispersion correction parameters of existing semiempirical methods are usually fitted against non-covalent complexation energies or other properties related to the total energy, such as the equilibrium geometries. In these approaches, the dispersion correction parameters will absorb some of the errors of the intermolecular repulsion term and other inaccuracies of the non-dispersion contribution to the interaction energy. While this obviously helps to improve the accuracy of the method on the training set, it has a few drawbacks: (1) It will hide the fundamental deficiencies of the non-dispersion terms, which may be masked by error compensation in the training set but show up in systems not seen in the training set; and (2) one cannot easily extract the dispersion contribution from the total energy, since only the total energy but not its components are meaningful. Therefore, in NOTCH we would like to fit the dispersion correction parameters to reproduce the dispersion energy itself rather than the total energy.
The DLPNO-CCSD(T) LED dispersion energy represents an accurate metric of the dispersion contribution to the total interaction energy that is not contaminated by non-dispersive interactions. However, its physical meaning is different from the contribution of the NOTCH dispersion correction to the total energy, Edispcorr [Eq. (32)], because the local dynamic correlation correction (Sec. II E) also describes some of the short-range dispersion. Thus, a naïve approach of fitting Edispcorr against the reference data would result in a drastic overestimate of the magnitude of dispersion. Instead, we have to define a NOTCH dispersion energy that is consistent with the physical picture of Eq. (B3). Herein, we have chosen the following approach:
Perform a usual NOTCH calculation of the dimer XY.
Perform another NOTCH calculation of XY, but all OAO-basis NOTCH two-electron integrals where μ, ν ∈ X and κ, λ ∈ Y are replaced by their HF values . This roughly amounts to selectively turning off the correlation corrections that are responsible for dispersion.
The NOTCH dispersion energy Edisp between the fragments X and Y is defined as the difference between the aforementioned two energies.
There are a few possible variations of this approach, for example, (a) instead of doing a full SCF calculation in step (2), one may also do a single SCF energy evaluation on the converged wavefunction of step (1); and/or (b) one can turn off the correlation under the LMO basis, instead of the OAO basis. The differences between these variations are expected to be small when the interfragment distance is not much shorter than the van der Waals distance.
Thus, we fitted the NOTCH dispersion parameters s8, a1, and a2, as well as the parameter αcorr [which controls the long-range behavior of the local dynamic correlation correction, Eq. (27)], against the diatomic LED dispersion energies at the DKH2 DLPNO-CCSD(T)/ANO-RCC-Full level of theory. When both atoms are open shell, we align their spins in parallel to each other, so as to minimize bonding interactions between the atoms. The resolution of the identity (RI) auxiliary basis sets were generated by the AutoAux method.79 All the core electrons were included in the correlation treatment and the TightPNO setting was enforced to obtain more tightly converged dispersion energies. The sum of strong pair and weak pair contributions to the dispersion energy was used as the reference data in the fitting.