You are on page 1of 20

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: J. Chem. Theory Comput. 2019, 15, 1652−1671 pubs.acs.org/JCTC

GFN2-xTBAn Accurate and Broadly Parametrized Self-Consistent


Tight-Binding Quantum Chemical Method with Multipole
Electrostatics and Density-Dependent Dispersion Contributions
Christoph Bannwarth,*,†,‡ Sebastian Ehlert,† and Stefan Grimme*,†

Mulliken Center for Theoretical Chemistry, Universität Bonn, Beringstr. 4, 53115 Bonn, Germany

Department of Chemistry, Stanford University, Stanford, California 94305, United States
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via RWTH AACHEN HOCHSCHUL BIBL on August 17, 2023 at 12:56:25 (UTC).

ABSTRACT: An extended semiempirical tight-binding


model is presented, which is primarily designed for the fast
calculation of structures and noncovalent interaction energies
for molecular systems with roughly 1000 atoms. The essential
novelty in this so-called GFN2-xTB method is the inclusion of
anisotropic second order density fluctuation effects via short-
range damped interactions of cumulative atomic multipole
moments. Without noticeable increase in the computational
demands, this results in a less empirical and overall more
physically sound method, which does not require any classical
halogen or hydrogen bonding corrections and which relies
solely on global and element-specific parameters (available up to radon, Z = 86). Moreover, the atomic partial charge dependent
D4 London dispersion model is incorporated self-consistently, which can be naturally obtained in a tight-binding picture from
second order density fluctuations. Fully analytical and numerically precise gradients (nuclear forces) are implemented. The
accuracy of the method is benchmarked for a wide variety of systems and compared with other semiempirical methods. Along
with excellent performance for the “target” properties, we also find lower errors for “off-target” properties such as barrier heights
and molecular dipole moments. High computational efficiency along with the improved physics compared to its precursor
GFN-xTB makes this method well-suited to explore the conformational space of molecular systems. Significant improvements
are furthermore observed for various benchmark sets, which are prototypical for biomolecular systems in aqueous solution.

1. INTRODUCTION architectures along with accordingly adjusted codes have


The accurate description of large molecular systems partic- become available in recent years,12−25 there is still need for
ularly in the condensed phase still remains one of the main inherently simple, reasonably accurate, and efficient electronic
challenges faced in theoretical chemistry.1,2 While Kohn− structure methods, which are applicable to large systems also
Sham density functional theory (DFT) can routinely provide without specialized hardware.
gas-phase (or continuum embedded) structures and energies Semiempirical quantum mechanical (SQM) methods26−28
for roughly a few hundred atoms, long-time molecular are the physically motivated approach to bridge the gap
dynamics simulations or conformational sampling are still between ab initio quantum mechanical (QM) and FF methods.
not feasible in a reasonably sized atomic orbital (AO) basis set The foundation of SQM methods typically is a valence-only
for systems of this size, e.g., within a day of computation time minimal basis self-consistent field method, derived from either
on a standard desktop computer. Recent efforts of our3−6 and Hartree−Fock theory or Kohn−Sham DFT. Since a QM
other groups7−10 have thus been directed toward the description for the valence electrons is used in SQM models,
development of low-cost DFT and Hartree−Fock methods their scope of applications is much broader than that of FFs. At
to allow such calculations. the same time, they are faster by at least 2 orders of magnitude
However, many interesting problems in biochemistry, compared to ab initio QM methods, due to drastic integral
materials science, and supramolecular and macromolecular approximations.28 These, however, come at the price of a
chemistry would require methods that can handle several significantly reduced accuracy and robustness. Nevertheless, if
thousands of atoms, which is beyond the scope even for the parameters are adjusted carefully, SQM methods can be used
aforementioned low-cost methods. Though classical force- sufficiently well for screening purposes of a desired
fields (FFs) can routinely be used for nanosecond time scale property.29−34 Nowadays, frequently used sophisticated
molecular dynamics simulations, 11 their limitations are
manifold and they are not suited for general use. Though Received: November 20, 2018
faster computers and more importantly parallel computing Published: February 11, 2019

© 2019 American Chemical Society 1652 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

Hartree−Fock-based zero-differential overlap (ZDO) methods or halogen bonds. These are reasonably well described
include PM635−40 and OM2.41−44 While the former mostly within the multipole-extended electrostatics.
finds use in the computation of ground state structures and
energies,45−48 the latter is still extensively used in excited state • Different from other semiempirical methods, GFN2-xTB
dynamics studies.49−53 In the past 20 years, density functional strictly follows a global and element-specific parameter
tight binding (DFTB) methods have found much atten- strategy. No element pair-specific parameters are
tion.54−61 Here, the Kohn−Sham DFT energy is expanded in employed.
terms of density fluctuations δρ relative to a superposition of
atomic reference densities. The highest-level variant (DFTB3) • As for the predecessor, properties around energetic
employs a self-consistent charge treatment including up to minima, such as geometries, vibrational frequencies, and
third order density fluctuation terms55,56 and has been noncovalent interactions, are the target quantities.
parametrized for a number of chemical elements.58−62 Already in our work for the GFN-xTB scheme, we
In particular, the element pair-specific parametrization used have identified that the preference for geometries
in all aforementioned methods has complicated the para- (instead of covalent bond energies) in the fit procedure
metrization procedure and only PM6 covers large parts of the resulted in systematically overestimated covalent bond
periodic table of elements (70 elements). Recently, we have energies. We have not deviated from this strategy, as
presented a DFTB3 variant, termed GFN-xTB, which mostly covalent bond energies are not of primary interest for
follows a global and element-specific parameters only strategy this method, and furthermore, the errors are very
and is parametrized to all elements through radon.34 The systematic as in GFN-xTB. This way, the errors will be
original purpose of the method and main target for the less random and more useful, once the systematic errors
parameter optimization has been the computation of molecular have been removedsee ref 45 for such a correction
scheme which in combination with GFN-xTB outper-
geometries, vibrational frequencies, and non-covalent inter-
forms any other semiempirical method of comparable
action energies. As such, it has successfully been used in
complexity. By keeping the focus on geometries as for
structure optimizations of organometallic complexes63,64 and
GFN-xTB, the GFN2-xTB method would likewise be
structural sampling.65−67 Apart from that, the method
well-suited for applications in high (electronic) temper-
performed well in high temperature molecular dynamics
ature molecular dynamics simulations to compute
simulations of electron impact mass spectra,68 and the
electron impact mass spectra.68 However, due to the
electronic structure information on GFN-xTB can furthermore
more sophisticated physical interaction terms, we find
serve as input to generate intermolecular FFs.69 Of particular that some “off-target” properties related to the electronic
significance is its contribution to a newly developed protocol structure also improve.
for the automated computation of nuclear magnetic resonance
(NMR) spectra. Therein, structural sampling with GFN-xTB is After a detailed description of the GFN2-xTB Hamiltonian,
employed to generate the conformer-rotamer ensemble, which technical details of the calculations are given. Then the method
in turn defines the detailed shape of an NMR spectrum. While is benchmarked on standard sets for molecular structures,
the GFN-xTB could successfully identify the relevant noncovalent energies, and conformational energies. The
structures for less polar molecules, polar and strongly performance for barrier heights and molecular dipole moments
hydrogen-bonded systems like sugars were not described as is also assessed. Other semiempirical methods are used for
well. This motivated the present work. Here, we attempt to comparison, namely, GFN-xTB,34 PM6-D3H4X,35,38,39 and
improve upon the GFN-xTB Hamiltonian in a physically DFTB3-D3(BJ).56,58,71
sound manner. The theory will be described below, but in a
nutshell, the method, which will be termed GFN2-xTB from 2. THEORY
now on, has the following characteristics:
Most previous derivations of density functional tight-binding
• The GFN2-xTB basis set consists of a minimal valence (DFTB) are formally based on a (semi)local density
basis set of atom centered, contracted Gaussian approximation (LDA) of DFT.28,54−56 Though nonlocal
functions, which approximate Slater functions (STO- exchange variants have been presented,72−74 the electron
mG).70 Polarization functions for most main group correlation functional has also been local in those cases. This
elements (typically second row or higher) are employed, however neglects long-range correlation effects, which are
which are particularly important to describe hypervalent responsible for the important dispersion interactions. Typi-
states. Different from GFN-xTB, hydrogen is only cally, corrections for the latter are then added in an a posteriori
assigned a single 1s function. fashion.71,75 Without explicitly referring to a specific functional
approximation, we will assume as starting point a general
• The GFN2-xTB Hamiltonian closely resembles that of density functional approximation, which includes a nonlocal
GFN-xTB or of the well-known DFTB3 method. (NL) VV10-type76 correlation contribution. The total Kohn−
However, GFN2-xTB represents the first broadly Sham energy expression is then given by (atomic units are used
parametrized tight-binding method to include electro-
ÄÅ
throughout)

ÅÅ
static interactions and exchange-correlation effects up to

E[ρ] = ∫ ρ(r)ÅÅÅT[ρ(r)] + Vn(r) + εXC


ÅÅÇ
second order in the multipole expansion. Furthermore,
LDA

ÉÑ
the simultaneously developed density-dependent D4 [ρ(r)]

1 ij 1 yz ÑÑ
+ ∫ jj + ΦC (r, r′)zzρ(r′) dr′ÑÑÑdr + Enn
dispersion model in a self-consistent formulation is an

2 k |r − r′| { ÑÑÖ
inherent part of the method. Different from either GFN- NL
xTB or DFTB3, GFN2-xTB does not employ other
classical FF-type corrections, e.g., to describe hydrogen (1)

1653 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

A,core ≔ 0. EA,valence is the


equivalently they are set to zero, i.e., E(0) (0)
Here, T[ρ(r)] is the kinetic energy per particle and Vn(r) the
(external) potential due to the nuclei. In writing so, we are sum of the valence orbital energies, which represent
using the short-hand notation for eigensolutions of the Hamiltonian for the free atom.

∫ ρ(r)T[ρ(r)] dr = −∑ ni ψi
1 2
∇i ψi
EA(0),valence = ∑ ∑ Pκκ0 HAl
2 l∈A κ∈l (4)
i
0
The zeroth order density matrix P is diagonal, as the atomic
with ψi being the molecular orbital and ni the corresponding orbitals with energy HlA are assumed to be eigenstates of the
occupation number. The last three terms in the integral over dr free atom. Erep and Edisp (0)
will be described by classical
arise from the electronic contributions to the mean-field expressions between “clamped” atoms in GFN2-xTB (see eqs
potential: the (semi)local exchange-correlation (XC), the 9 and 32). It can be seen that the zeroth order energy in the δρ
Coulomb, and the nonlocal correlation energy, respectively. expansion is thus related to the well-known Buckingham- or
The nonlocal correlation kernel ΦNL C (r,r′) captures long-range Lennard-Jones-type potentials, which can be a reasonable
correlation effects. Enn is the classical nuclear repulsion energy. approximation to treat noncovalent interactions of noble gas
In density functional tight-binding (DFTB) theory, the total atoms. It does not contribute to the tight-binding electronic
energy is expanded in terms of density fluctuations around a energy, which depends only on the fluctuations δρ.
superposition of (neutral) atomic reference densities 2.1.2. First order terms. For the choice of spherical atomic
ρ0 = ∑A ρA :28,54 reference densities ρA0, the description of covalent bonds in
0

E[ρ] = E(0)[ρ0 ] + E(1)[ρ0 , δρ] + E(2)[ρ0 , (δρ)2 ] tight-binding theories becomes possible at first order in the δρ
expansion. This is typically achieved by changes in the
+ E(3)[ρ0 , (δρ)3 ] + ··· (2) occupation δP of the atomic energy levels by invoking the
well-known extended Hückel approximation. Since δρ is
A fixed reference density ρ0 is employed, and the calculation of assumed to be nonzero only in the valence electron space
the electronic structure is done in terms of the density (see above), the latter will be restricted to the valence orbitals.
fluctuations δρ. The most widely used variants truncate this Due to the density fluctuations to first order, the atoms can
expansion after the third order term.34,55,56 The same is true obtain a net charge and are no longer neutral. This has no
for the GFN2-xTB approach presented here. The first effect on the interatomic electrostatic (ES) interactions, since
approximation made in tight-binding methods is the the electrostatic potential from the remaining neutral atoms
assumption that the density fluctuations are restricted to the (with ρA0) is still zero. However, since starting from the DFT-
valence orbital space, while the core electron density remains
frozen. The individual approximations for the energy NL expression in eq 1 in GFN2-xTB, the energy to first order
contributions to different orders in δρ have been described in δρ is augmented with a first order dispersion contribution:
before,28,54,56,72,73 but we briefly outline the origin of newly E(1)[ρ0 , δρ] ≈ Edisp
(1)
+ ∑ EA(1),valence
introduced terms in GFN2-xTB. A (5)
2.1. Contributions Relevant to the GFN2-xTB energy.
2.1.1. Zeroth Order Terms. In previous derivations of DFTB, The last term describes first order changes in the occupation
the zeroth order term is reduced to a classical repulsion and energies of the electronic valence levels. If combined with
term.28,54−56 Due to the different starting point (DFT-NL), an eq 4, this is essentially equivalent to extended Hückel theory
additional term will arise at first order: (EHT), i.e., EEHT = ∑A(E(0) (1)
A,valence + EA,valence). The explicit
expression to approximate this term in GFN2-xTB will be
E(0)[ρ0 ] ≈ Erep(0)
+ Edisp (0)
+ ∑ (EA(0),core + EA(0),valence) given in Section 2.2. Due to the nonvanishing zeroth order
´ÖÖÖÖÖÖÖÖÖÖÖÖÖÖ≠ÖÖÖÖÖÖÖÖÖÖÖÖÖÖÆ
Buckingham/Lennard ‐ Jones ‐ type term A (3) London dispersion potential of the other charge-neutral atoms,
Here, the total energy in eq 1 is partitioned into same-center the dispersion energy changes at first order in δρ. The
and off-center energy terms. All one-electron and two-electron expansion/contraction of the atomic density will increase/
terms in eq 1 involving particles at the same atom reduce to a decrease the magnitude of the interaction. In GFN2-xTB these
single number, the atomic energy EA, which can furthermore and second order effects (see below) will be taken into account
be partitioned into a core and a valence contribution (see last within the self-consistent D4 dispersion model (see Section
two terms in eq 3). Since the reference densities refer to 2.2.6).77,78 It should be mentioned that likewise first and
neutral, spherically symmetric atoms, the summed Coulomb higher order effects for the repulsion energy exist, which are
interactions (e−e, n−n, e−n) between distant atoms vanish at generally neglected in DFTB methods. The same is true for the
this level of approximation. However, in regions with GFN2-xTB method presented here, though these neglected
overlapping atomic densities, repulsive forces due to effects can partially be absorbed in the parametrization of the
exchange-correlation and charge penetration effects are other terms.
present. Similar to previous works on DFTB, this term is 2.1.3. Second Order Terms. At second order in δρ, energy
modeled by a classical repulsion term.28,54,56,72,73 Different contributions arise, which require a self-consistent tight-
from DFTB but in line with our previous method GFN-xTB,34 binding calculation.
no element pair-specific parametrization will be employed for E(2)[ρ0 , (δρ)2 ] ≈ E ES
(2) (2)
+ E XC (2)
+ Edisp (6)
this term. Due to the presence of a NL correlation functional in
the energy expression in eq 1, a pairwise London dispersion At second order, interatomic electrostatic and one-center
contribution is also present at zeroth order (see Supporting exchange-correlation terms occur. In DFTB schemes, these are
Information, Section 1.5). The sum of the atomic energies is a generally condensed in a Mataga−Nishimoto−Ohno−Klop-
constant for a given system, and the total tight-binding energy man damped Coulomb interaction between atomic or shell
is given relative to the sum of atomic core energies, or charge monopoles.79−81 The same will be done in GFN2-xTB
1654 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

(see Section 2.2). However, we will go beyond the monopole electronic temperature Tel due to Fermi smearing.83 This term
approximation for both E(2) (2)
ES and EXC (see Section 2.2.5) and is necessary to provide a variational solution for fractional
include anisotropic effects up to second order in the multipole occupations and is given by
expansion. Isotropic second order δρ effects in the dispersion
energy are also included within the self-consistent D4 G Fermi = kBTel ∑ ∑ [niσ ln(niσ ) + (1 − niσ ) ln(1 − niσ )]
dispersion model in GFN2-xTB. σ=α ,β i

2.1.4. Third Order Terms. Basically all energy contributions (10)


of the Kohn−Sham energy expression in eq 1 that require a kB is Boltzmann’s constant and niσ refers to the (fractional)
self-consistent solution at the DFT level of theory will also occupation number of the spin-MO ψiσ. These are given by
contribute to energy corrections of higher order (≥2) in δρ. As
in previous works on DFTB and GFN-xTB,34,55 we will neglect 1
niσ =
all third order terms but an isotropic on-site term, which exp[(ϵi − ϵFσ )/(kBTel)] + 1 (11)
mostly originates from short-ranged Coulomb and XC effects.
ϵi is the orbital energy of the orbital ψi and = 0.5 ϵσF + (ϵσHOMO
(3) 3 (3) (3) 1 3 ϵσLUMO) is the Fermi level for the respective orbital space (α or
E [ρ0 , (δρ) ] ≈ E ES + E XC ≈ ∑ ∑ ΓA ,lqA ,l β). Tel the electronic temperature, which by default is equal to
3 A l∈A (7)
300 K as in GFN-xTB.34
Different from those previous works, however, the third order The occupation ni for the spatial molecular orbital ψi (which
term is expressed in terms of partial shell charges (Mulliken is the same for ψiα and ψiβ) is given as
approximation) in GFN2-xTB.
2.2. GFN2-xTB Method. Having related the individual ni = niα + niβ (12)
energy components to different orders in δρ, we will outline It should be stressed that this finite temperature treatment
the specific contributions to the GFN2-xTB energy in the predominantly has the purpose of enabling fractional orbital
following. The total GFN2-xTB energy expression is given by occupations in static correlation cases. The analytical gradients
EGFN2 − xTB = Erep + Edisp + E EHT + E IES + IXC + EAES of eq 8 obtained with fractionally occupied orbitals are
numerically exact and can be used without any problems for
+ EAXC + G Fermi (8) very tightly converged structure optimizations or in high-
temperature molecular dynamics simulations.
The newly introduced abbreviations in the subscripts indicate As usual, the spatial molecular orbitals (MOs) ψi are
the isotropic electrostatic (IES) and isotropic XC (IXC), and expressed as linear combinations of atom-centered orbitals
likewise the anisotropic electrostatic (AES) and anisotropic XC (LCAO),84,85
(AXC) energies, respectively. Before going into detail, it is
NAO
important to note that no halogen- or hydrogen-bond
corrections are included in GFN2-xTB. The description of ψi = ∑ cκiϕκ (ζκ , STO‐mG)
these interactions is already improved by the AES energy EAES, κ (13)
which is described in Section 2.2.5. Following Stewart’s Gaussian expansions,70 ϕκ refer to
2.2.1. Classical Repulsion Energy. For the repulsion energy contracted Gaussian atomic orbitals, which are used to
in eq 8, we employ an atom pairwise potential similar to the approximate a spherical Slater-type orbital with exponent ζκ.
one proposed in refs 34 and 82. The number of primitives m varies between 3 and 6explicit
numbers of primitives are given in the Supporting Information
Y Aeff Y Beff −(αAαB)1/2 (RAB)k rep
Erep = ∑ e (SI). In GFN2-xTB, a minimal spherical valence basis set is
AB
RAB (9) employed, whereas most heavier main group elements (Z > 9)
are provided with a polarization function. AOs located on the
where Yeff eff
A and YB define the magnitude of the repulsive same atom are always orthogonal to each other in GFN2-xTB,
interaction. Like αA and αB, they are element-specific thus formally forming a basis of eigenstates of the free atom.
parameters. krep is a parameter which is equal to one if both The basis set employed is given in Table 1, whereas the Slater
atoms are either H or He and equal to 3/2 otherwise. The exponents are listed in the SI. As in GFN-xTB, the “f-in-core”
different krep parameters for the very light element pairs turned approximation is employed for lanthanides. Variational
out to be beneficial for torsion barriers in alkanes, without minimization of the energy expression in eq 8 with respect
sacrificing the accuracy for noncovalent complexes. The to the linear coefficients cκi in eq 13 leads to the general
repulsion energy in GFN2-xTB is described classically and is
independent of changes in the electronic structure. It should be
Table 1. Slater-Type AO Basis Sets Employed for the
mentioned that, compared to GFN-xTB,34 the Yeff values
Different Elementsa
correlate less with the atomic number in GFN2-xTB.
2.2.2. Choice of Electronic Wave Function and Finite element basis functions
Temperature Treatment. As in the predecessor GFN-xTB, we H ns
are working with a formally spin-restricted wave function He ns, (n + 1)p
throughout and no spin density dependent terms are present. group 1, Be−F, Zn, Cd, Hg−Po nsp
Hence, α and β orbitals always have identical spatial parts and Ne nsp, (n + 1)d
orbital energies but possibly different occupation. In order to group 2, 13−18 nspd
handle static correlation (nearly degenerate states) via transition metals and lanthanides nd, (n + 1)sp
fractional orbital occupations, a finite temperature treatment
a
is used. The last term GFermi in eq 8 formally refers to the n denotes the principal quantum number of the valence shell of the
entropic contribution of an electronic free energy at finite element.

1655 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

eigenvalue problem, which is likewise encountered in Hartree− Hκκ = HAl − HCN


l
A
CN′A , (κ ∈ l ∈ A ) (17)
Fock and Kohn−Sham density functional theory
with
FC = SCϵ (14)
Natoms
The elements of the tight-binding Hamiltonian or “Fock” CN′A = ∑ (1 + e−10(4(RA ,cov + RB,cov)/3RAB− 1))−1
matrix F will be given after presentation of the individual B≠A
energy terms. S is the overlap matrix, and C is the LCAO-MO
coefficient matrix. ϵ is a diagonal matrix containing the orbital × (1 + e−20(4(RA ,cov + RB,cov + 2)/3RAB− 1))−1 (18)
energies. Here, HlA and HlCNA are element-specific parameters. Compared
2.2.3. Extended Hückel-Type Energy. The extended Hückel
to the D3 coordination number,86 which has been used in
contribution is the crucial ingredient to describe covalent
GFN-xTB,34 CNA′ is smoother and slightly more long-ranged.
bonds in tight-binding methods.
The covalent radii of the atoms RA,cov/RB,cov are the rescaled
E EHT = ∑ ni⟨ψi|Ĥ 0|ψi⟩ = ∑ ∑ ∑ nicκicλi⟨ϕλ|Ĥ 0|ϕκ ⟩ radii from ref 87 as used in the D386 and D477 (see below)
i κ λ i
model. Sκλ is the overlap between the two AOs. The last three
terms in the product on the right-hand side of eq 16 modify
≡ ∑ ∑ PκλHλκ the magnitude of the interaction in certain situations. ζκ and ζλ
κ λ (15) are the Slater exponents of the two AOs. This factor is unity for
ζκ = ζλ and otherwise reduces the magnitude of the Hκλ matrix
Here, the density matrix element Pκλ = P0κλ + δPκλ. Due to its
element. Similarly, the second to last term reduces the
origin in the first order δρ fluctuation term, the operator Ĥ 0 is “covalent” interaction for two atoms with different electro-
formally a one-electron operator, which should however negativities. While the latter term exclusively adjusts nonpolar
provide the zeroth order energy (eq 4) for neutral non- vs polar (or ionic) binding, the former term adds more
interacting atoms. The corresponding off-diagonal matrix flexibility for orbitals with different compactness and, thus, is
elements Hκλ are thus approximated in the following way: also relevant in homonuclear situations. The explicit depend-
ij 2 ζκζλ yz
Hκλ = kll′ (Hκκ + Hλλ)Sκλjjjj zz
z
ence on orbital exponents introduces effects present for HF/
j ζκ + ζλ zz
1/2
1
k {
(1 + kEN ΔENAB2)Π(RAB ,ll′), DFT in the kinetic energy one-electron integrals. As in GFN-
2 xTB,34 the distance and shell dependent polynomial Π(RAB,ll′)
(κ ∈ l ∈ A , λ ∈ l′ ∈ B) (16) is based on element-specific parameters kpoly poly
A,l /kB,l′ , which are
fitted, while the RA,cov′/RB,cov′ are taken from ref 88.
jij i R yz zyzjij i R yz zyz
j polyj z j polyj
j j z z j j zz zzzz
z
kll′ is the effective scaling factor common to all EHT methods.
Π(RAB ,ll′) = jj1 + kA , l jjj z zj1 + kB , l′ jj
j Rcov ′ ,AB zzz zzzjjj jj R zz zz
1/2 1/2

jj
Its value depends on the angular momentum quantum number
k { {k k { {
AB AB

k
of the interacting AOs (see Table 2 for explicit values). Similar
cov ′ ,AB
to GFN-xTB,34 the atomic energy levels, i.e., the diagonal
elements of the EHT Hamiltonian matrix, are made flexible by (19)
being proportional to the coordination number, The key purpose of this term is the distance-dependent
adjustment of the EHT-type interaction, which provides a
Table 2. Global Empirical Parameters of the GFN2-xTB better balance between short- (covalent) and long-ranged
Methoda (noncovalent) effects.
parameter value
The EHT-type term in GFN2-xTB is mostly responsible for
covalent binding. Via the coordination number dependence of
kss 1.85
the valence energy levels, these obtain additional flexibility
kpp, kdd 2.23
beyond the δρ-based expansion up to first order. To some
ksp 2.04b
extent, the fitted element- and shell-specific parameters for HlA
ksd, kpd 2.00
and HlCNA can implicitly account for the formally neglected first
krep 1.5c
Ks 1.0 and second order on-site δρ effects. This procedure allows
Kp 0.5 short-ranged many-body effects to be captured, formally up to
Kd 0.25 infinite order, by a linearized empirical model. Noteworthy is
kEN 0.02 the importance of that term for atoms that can become
multipole parameters hypervalent: the energy level for the d-polarization functions
Δval 1.2 are typically high in energy for small CNA′ . Then, neither do
Rmax 5.0 these functions interfere strongly with the valence functions in
a3 3.0 “standard” covalent binding nor do they significantly affect
a5 4.0 noncovalent interactions in a way which is reminiscent to the
dispersion parameters well-known basis set superposition error (BSSE) in AO-based
a1 0.52 ab initio calculations. For large CN′A values, however, this d-
a2 5.0 level is lowered in energy, which then enables a much better
s6 1.0 description of hypervalency.
s8 2.7 2.2.4. Isotropic Electrostatic and Exchange-Correlation
s9 5.0 Energy. For charged and polar systems, the density ρ deviates
from the reference density ρ0. In that case, the net partial
a
The parameters are either dimensionless or in atomic units. charges on the individual atoms are nonzero. In GFN2-xTB,
b
Obtained as ksp = 0.5 (kss + kpp). ckrep = 1.0 for H/He pairs. the isotropic electrostatic and XC terms are treated with shell-
1656 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

Figure 1. Schematic overview of the employed approximations for the second order ES and XC energy in tight-binding theory. While the isotopic
approximation is well-established,34,54,56 the inclusion of anisotropic effects by higher order multipole terms has rarely been described in the
literature.91−93 For simplicity, an atomic charge partitioning is used throughout in this scheme.

wise partitioned Mulliken partial charges (cf. ref 89). The for some transition metals without diminishing the accuracy
following contribution to the energy is then obtained: for main group elements.
The shell-wise treatment requires the definition of reference
1 1
E IES + IXC = ∑ ∑ ∑ qA ,lqB,l′γAB,ll′ + ∑ ∑ ΓA ,lqA ,l 3 valence shell occupations. For the occupation of elements of
2 A , B l ∈ A l ′∈ B
3 A l∈A group 1, 2, 12, 13, 16, 17, and 18, we follow the aufbau
(20) principle, whereas for transition metals a modified aufbau
principle of the form ndx−2(n + 1)s1(n + 1)p1 (x denotes the
Here, the first term on the right-hand side is derived from the group) is used. Lighter elements of groups 14 and 15 are
second order energy, while the last term is due to third order handled slightly different to better reflect the occupations in
density fluctuations. qA,l refers to an isotropic monopole charge bonded atoms. For this purpose, carbon is treated with a
of the l shell on atom A. The distance dependence of the reference valence shell occupation of 2s12p3, whereas fractional
Coulomb interaction within the first term is described by a reference occupations of type ns1.5npx−11.5 (x denotes the
generalized form of the well-established Mataga−Nishimoto− group) are used for N, Si, Ge, P, and As. The remaining
Ohno−Klopman79−81,90 formula:

ij yz
elements of groups 14 and 15 follow the standard aufbau

γAB , ll′ = jjj zz


j R 2 + η−2 zz
1/2 principle. All element-specific parameters are given in the SI.
1
k AB {
2.2.5. Multipole-Extended Electrostatic and Exchange-
(21) Correlation Energy. The approximate expression for the ES
and XC energy commonly employed in tight-binding theory is
Here, RAB is the interatomic distance and η is the average of the derived from the first two terms of the second order energy
effective chemical hardnesses of the two shells l and l′ on the (see eq 6):56
ij yz
atoms A and B:
jj 1 zz
jj + zzδρ(r )δρ(r ) dr dr
jj zzz
jj rij
∂ 2E XC
z
1 1
j ρ = ρ0 z
1
η = (ηA , l + ηB , l′) = ((1 + κAl)ηA + (1 + κBl′)ηB) (2)
E ES (2)
+ E XC = ∬
k {
i j i j
2 2 2 ∂ρ(ri)∂ρ(rj)
(22)
(24)
ηA and ηB are treated as element-specific fit parameters. and κlA
κlB′ are fitted element-specific scaling factors for the individual Anisotropic Electrostatic Interactions. In Figure 1, it is
shells (note that κlA = 0 for l = 0). This way, electrostatic schematically shown how eq 24 is typically approximated by
interactions between distant atoms and on-site changes in the purely isotropic energy terms in DFTB.54,56 γAB is the
isotropic electrostatic and XC energy are treated in a seamless interatomic Coulomb interaction, which is damped to a finite
manner. The third order term is restricted to shell-wise on-site value at short range (see eq 21). This short-range damping
terms. As discussed in the literature on DFTB,56 this term can then also includes effects due to the second order changes in
help to stabilize charged atomic states and partially remedy the semilocal XC energy E(2)
XC. Apart from different partitioning
shortcomings from missing, e.g., diffuse functions in the AO schemes (atomic or shell-wise, see Section 2.2.4), the same
basis. functional form for the second order ES/XC energy is used in
The shell-wise parameter ΓA,l is obtained from the element- DFTB,54,56 GFN-xTB,34 and the GFN2-xTB method pre-
specific parameter ΓA and the global, shell-specific parameters sented here. The basic possibility of including higher multipole
Kl. ES interactions in DFTB has been suggested in ref 92, and an
implementation within the so-called “Gaussian tight binding”
ΓA , l = Kl ΓA (23) method has been described recently93 (for an earlier multipole-
extended tight binding model to describe zirconia crystals, see
ΓA is formally related to ∂ηA/∂qA|qA=0 but is a fitted parameter ref 91). In GFN2-xTB, we pioneer in going beyond this
in GFN2-xTB. At variance with GFN-xTB, we use a shell-wise, monopole approximation in a broadly parametrized tight
though parameter-economic, treatment by treating Kl as global binding method for both ES and XC terms including all terms
parameters. The shell-wise treatment appeared to be beneficial up to second order in the multipole expansion. The newly
1657 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

incorporated terms are schematically shown in Figure 1 and R max − R 0A


outlined below (for a derivation, see SI). The AES energy in R 0A′ = R 0A +
1 + exp[−4(CN′A − Nval − Δ val )] (29)
GFN2-xTB is given by
EAES = Eqμ + EqΘ + Eμμ (25a) Δval = 1.2 and Rmax = 5.0 bohrs. Aside from those light
elements (see SI), RA0 ′ = 5.0 bohrs for all elements. This
1 flexible RA0 ′ value reduces the strength of the AES interactions
= ∑ {f3 (RAB)[qA(μBT RBA) + qB(μAT RAB)] (25b)
for strongly coordinated atoms and was found to increase the
2 A ,B
robustness of the SCF convergence for inorganic clusters. In
+ f5 (RAB)[qA RTABΘBRAB + qBRTABΘA RAB (25c) case RA0 ′ ≠ 5.0 bohrs, an extra parameter Nval in eq 29 is used,
which correlates with the typical valency of the element.
− 3(μAT RAB)(μBT RAB) + (μAT μB )RAB2]} (25d) Primarily, the AES terms are intended to improve the
noncovalent interactions between the outer, i.e., less
Here, μA is the cumulative atomic dipole moment of atom A coordinated atoms. This way, no extra hydrogen or halogen
and ΘA is the corresponding traceless quadrupole moment bond corrections nor any element-specific bond adaptations
are required.
3 αβ δαβ xx
Θαβ
A = θA − (θA + θAyy + θAzz) Anisotropic XC Energy. While the basic idea of including
2 2 (26) AES terms in a TB model has been suggested before,92 no such
94 extension for second order XC effects was mentioned so far. In
The cumulative atomic multipole moments (CAMM) up to
second order are computed from the context of excitation energies, the approach of Dominguez
et al. to include INDO-like terms in DFTB is somewhat
qA = ZA − ∑ ∑ Pκλ ´⟨ÖÖÖÖÖÖ
ϕλ|ϕκ ⟩
Ö≠ÖÖÖÖÖÖÖÆ related, though its basis is rather founded in a semiempirical
κ∈A λ Sλκ (27a) hybrid-like density functional. Here, we take the second term
of eq 24 as the starting point. The second order XC energy
μAα = ∑ ∑ Pκλ(αASλκ − ⟨ϕλ|αi|ϕκ ⟩ ) contribution takes the form of a static XC kernel. In the local
´ÖÖÖÖÖÖÖÖÖ
Ö≠αÖÖÖÖÖÖÖÖÖÖÆ
κ∈A λ Dλκ (27b) density approximation, this term reduces to a pure same-site
energy in a tight-binding scheme. Going up to second order
θAαβ = ∑ ∑ Pκλ(αADλκβ + βADλκα − αAβASλκ − ⟨ϕλ|αiβi |ϕκ ⟩ ) multipolar terms, E(2)XC can then be simplified to (see SI for the
´ÖÖÖÖÖÖÖÖÖÖÖ≠αβÖÖÖÖÖÖÖÖÖÖÖÆ derivation):
κ∈A λ Q λκ
(27c) (2) q
qA 2 μA
|μA |2 + f XC ΘA
||ΘA ||2 )
E XC ≈ ∑ ( ´ÖÖÖÖÖÖÖ
f XC A
Ö≠ÖÖÖÖÖÖÖ
ÖÆ
+ f XC
´ÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖ
Ö≠ÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖÖ
ÖÆ
α and β are Cartesian components. Dαλκ and Qαβλκ are the electric A isotropic XC anisotropic XC (30)
dipole and quadrupole moment integrals between the AOs ϕκ
and ϕλ. The expressions for the CAMMs directly originate The isotropic monopolar term is already included in the shell-
from the multipole expansion (in Cartesian coordinates) and wise isotropic XC energy (see Section 2.2.4). To our
guarantee that the respective overall molecular moments are knowledge, the other terms are proposed here for the first
correctly preserved. These expressions give the atomic time in a tight-binding context. They form the anisotropic XC
contribution (Mulliken approximation) of the particular energy in GFN2-xTB:
atom to the overall multipole moment. The CAMMs are
μ Θ
defined as such that their respective origin is located at the EAXC = ∑ (f XC |μA |2 + f XC
A ||ΘA ||2 ) A

particular atom; i.e., they are origin-independent, as enforced A (31)


by the “shift” contributions from lower order moments (eqs
27b and 27c). In eq 25, we have gone up to second order in Again, μA and ΘA are the cumulative atomic dipole and
the multipole expansion of the Coulomb energy, and thus all traceless quadrupole moments, which have already been
terms that decay with RAB−3 or slower are included (see eq 28). introduced in eqs 27 and 26. f μXCA
and f ΘXCA are fitted element-
The monopole−monopole term, which is the term of lowest specific parameters. Formally, these terms are supposed to
order (see SI), has already been included in the shell-wise capture changes in the atomic XC energy, which result from
isotropic ES energy described in Section 2.2.4. The terms anisotropic density distributions (polarization). To some
containing higher order multipoles should improve the extent, they may also alleviate shortcomings of the small AO
description of the anisotropic electron density around the basis set (e.g., insufficient polarization functions).
atoms. We chose to employ an atomic partitioning in GFN2- 2.2.6. Density-Dependent Dispersion Energy. In GFN2-
xTB for these terms. To avoid divergence for the AES energy xTB, we treat dispersion interactions by means of a self-
(eq 25), we damp the corresponding terms at short distances. consistent variant of the recently published D4 dispersion
The distance dependence including damping is given by model.77,78 In typical semiempirical mean-field methods,
London dispersion interactions are generally treated by
fdamp (an , RAB) 1 1 means of post-SCF corrections (see refs 28 and 75 for
fn (RAB) = n = n · an reviews). Though the widely employed D3 dispersion model
RAB RAB R AB
1 + 6 R0 ( )
AB (28)
takes environmental effects via the geometric coordination
number into account, electronic structure effects are missing.
The damping function is related to the zero damping function In a tight-binding context, the D3 dispersion energy should be
in the original D3 dispersion model.86 an are adjusted global regarded as a zeroth order term; i.e., it corresponds to E(0) disp.
parameters, whereas RAB0 = 0.5 (R0 ′ + R0 ′) determines the
A B
Here, we go beyond this model and include effects up to
damping of the AES interaction. RA0 ′ is made dependent on the second order within the self-consistent formulation of the D4
coordination number (see eq 18) for many lighter elements. dispersion model:78

1658 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

(0) (1) (2)


Edisp = Edisp + Edisp + Edisp

CnAB(qA , CNcov
A B
, qB , CNcov )
≈ −∑ ∑ sn n f ndamp,BJ (RAB)
A > B n = 6,8 RAB
(3cos(θABC)cos(θBCA)cos(θCAB) + 1)C9ABC(CNcov
A B
, CNcov C
, CNcov )
− s9 ∑ × f 9damp,zero (RAB , RAC , RBC)
A>B>C (RABRACRBC)3 (32)

ÄÅ ÄÅ ÉÑ|ÉÑÑ
ÅÅ l ÅÅ i ÑÑoÑÑ
ÅÅ o o ÅÅ jj y
zzÑÑÑo
A ,r z oÑÑÑ
ÅÅ o o Å j zzÑÑo
ξA(qA , qA , r ) = expÅÅ3m 1 − expÅÅ4ηA jjj1 − eff }ÑÑ
The last term is the charge-independent three-body (also

ÅÅ o Å ZA + qA zÑÑÑo z Ñ
eff

o ÅÅ j oÑÑÑÑ
o
Z + q

ÅÅ o
called Axilrod−Teller−Muto or ATM) dispersion term,95,96
o {ÑÖo
A

Å
ÅÇ k
r

ÅÇ n ~ÑÖ
which is added to incorporate the dominant part of the many-
body dispersion energy. Different from the charge-dependent
two-body term, this three-body contribution does not affect (36)
the electronic energy. where ηA is the chemical hardness taken from ref 98. Zeff
A is the
The damping functions f damp,BJ
n and f damp,zero
9 in eq 32 have effective nuclear charge of atom A, which has been determined
been defined in refs 97 and 86, respectively. It is, however, by subtracting the number of core electrons represented by the
important to note that, in the D4 model, the BJ-type cutoff def2-ECPs in the time-dependent DFT reference calculations
radii97 are used in both f damp,BJ
n and f damp,zero
9 (see below). While (see ref 77 for details). Due to the charge dependency, the
the C8 are calculated recursively86 from the lowest order CAB
AB
6 pairwise dispersion energy in GFN2-xTB enters the electronic
coefficients, the latter are computed from a numerical Casimir- energy and is self-consistently optimized. A similar expression
Polder integration. is used in the standard form of the DFT-D4 method,77 but
therein, the partial charges are obtained by purely geometrical
3
C6AB = A
∑ wjαA̅ (iωj , qA , CNcov B
)αB̅ (iωj , qB , CNcov ) means. The rational damping function used in the DFT-D4
π j (33) model is given by
n
wj are the integration weights, which are derived from a RAB
f ndamp,BJ (RAB) =
trapezoidal partitioning between the grid points j (j ∈ [1, 23]). n
RAB crit.
+ (a1·RAB + a 2)6
The isotropically averaged, dynamic dipole−dipole polar-
izabilites α̅ A at the jth imaginary frequency iωj are obtained crit. C8AB
with RAB =
from the self-consistent D4 model; i.e., they are depending on C6AB (37)
the covalent coordination number78 and are also charge
dependent. The method thus relies on precomputed atomic The zero damping function for the ATM dispersion is defined
polarizabilities at a certain molecular geometry, i.e., with the slightly different from the previous implementations of DFT-
atom having a GFN2-xTB computed atomic partial charge of D3, namely, the factor 4/3 is dropped and the cutoff radii are
qA,r and a covalent coordination number CNA,r cov (the index r calculated consistently with the two-body terms.

jij ij Rcrit.R crit.R crit. yz zyz


indicates values for the reference structures). Similar to D3, a
jj jj 3 AB BC CA zz zz
(RAB , RAC , RBC) = jj1 + 6jj z z
−1

jj RABRBCR CA zzz zzz


16
Gaussian-weighting scheme based on the covalent coordina-
jj
j k { z{
damp,zero
tion number CNAcov via the WrA terms (see ref 77) is employed.
k
f9
NA,ref
A
αA̅ (iωj , qA , CNcov ) = ∑ ξAr(qA , qA ,r )αA̅ ,r(iωj , qA ,r , CNcov
A ,r
) (38)
r
2.3. GFN2-xTB Hamiltonian Matrix. As mentioned
W Ar(CNcov
A A ,r
, CNcov ) (34) before, GFN2-xTB includes energy terms of second and
third order in δρ. Therefore, the energy expression in eq 8
The Gaussian weighting for each reference system is given by needs to be solved self-consistently. To compute the density
matrix, the Roothaan-Hall-type eigenvalue problem (eq 14) is
Ngauss
1 then solved.
W Ar(CNcov
A A ,r
, CNcov )= ∑ A
exp[−6j ·(CNcov A ,r 2
− CNcov )] As for the total energy, the matrix elements of the GFN2-
j=1
5
xTB Hamiltonian can be decomposed into individual
NA ,ref contributions
with ∑ W Ar(CNcov
A A ,r
, CNcov )=1 IES + IXC AES AXC D4
r Fκλ = Hκλ + Fκλ + Fκλ + Fκλ + Fκλ (39)
(35)
Due to the analogy to Hartree−Fock, we will denote this
where 5 is a normalization constant. The number of Gaussian matrix simply as TB-Fock matrix in the following. The
functions per reference system Ngauss is mostly one, but is equal extended Hückel matrix elements Hκλ have already been
to three for CNA,rcov = 0 and reference systems with similar described in eq 16. The general derivation for the isotropic ES
coordination number (see ref 77 for details). The charge- and XC contributions to the TB-Fock matrix has been shown
dependency is included via the empirical scaling function ξrA. in ref 56. In GFN2-xTB, these are given by
1659 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

where indices κ and λ denote the AOs with corresponding


IES + IXC 1 angular momenta l and l′ and the second sum runs over the
Fκλ = − Sκλ ∑ ∑ (γAC , ll ″ + γBC , l ′ l ″)qC , l ″
2 C l″
atoms C and their shells l″.
1 The last three terms in eq 39 are new, and their derivation
− Sκλ(qA , l 2 ΓA , l + qB , l′2 ΓB , l′) can be found in the SI. FAES AXC
κλ and Fκλ both involve terms that
2
include electric dipole and quadrupole, as well as overlap,
(κ ∈ l(A), λ ∈ l′(B)) (40) integrals. In a condensed notation, they can be expressed as

AES AXC 1
Fκλ + Fκλ = Sκλ[VS(R B) + VS(R C)] (41a)
2
1
+ DTκλ [VD(R B) + VD(R C)] (41b)
2
1 αβ
+ ∑ Q κλ [VQαβ(R B) + VQαβ(R C)] , ∀ κ ∈ B, λ ∈ C (41c)
2 α ,β

NA ,ref NB,ref
∂ξAr CnAB ,ref damp,BJ
Here, the respective integral (overlap, dipole, and quadrupole) dA = ∑ ∑∑ ∑ W ArW BsξBs· sn n fn (RAB)
∂qA RAB
proportional potential terms are given as r B s n = 6,8

l
o
o T
VS(R C) = ∑ o
(46)
mR C[f5 (RAC)μA RAC
o
o
A o
n
2
− RAC3f5 (RAC)(μAT RAC) Here, the dispersion coefficient for two reference atoms CAB,ref
n
is evaluated at the reference points, i.e., for qA = qr, qB = qs,
− f3 (RAC)qA RAC] − f5 (RAC)RTACΘA RAC − f3 (RAC)μAT RAC CNAcov = CNrcov, and CNBcov = CNscov.
|
o
o
+ qAf5 (RAC) R C 2RAC 2 − qAf5 (RAC) ∑ αABβABαCβC o
}
2.4. Technical Details. The global parameters in Table 2
o
o
o
1 3

~
and the element-specific parameters (see SI) have been
2 2 α ,β determined by minimizing the root-mean-square deviation
μC T
+ 2f XC ΘC T
R CμC − f XC R C[3ΘC − Tr(ΘC )I]R C (42) (RMSD) between reference and GFN2-xTB-computed data.

ÅÄÅ
The procedure is the same as in GFN-xTB and relies on the
Å
∑ ÅÅÅÅÅRAC3f5 (RAC)(μAT RAC) − f5 (RAC)μA RAC 2 + f3 (RAC)qA RAC
Levenberg−Marquardt algorithm.99,100
ÅÅ
ÅÇ
VD(R C) = The global parameters have been determined along with the
ÉÑ
ÑÑ
element-specific parameters for the elements H, C, N, and O.
Ñ
A

+ 3qAf5 (RAC)RAC ∑ αCαAC ÑÑÑÑ


ÑÑ
Then the element-specific parameters for the other elements
ÑÖ
2
− qAf5 (RAC)R CRAC
α
were subsequently determined keeping all existing parameters
μC ΘC
fixed. For the lanthanides, only the parameters for Ce and Lu
− 2f XC μC + 2f XC [3ΘC − Tr(ΘC )I]R C (43) were freely fitted, while a linear interpolation with the nuclear
ÅÄÅ 3 ÑÉÑ
VQαβ(R C) = −∑ qAf5 (RAC)ÅÅÅÅ αACβAC − RAB2 ÑÑÑÑ
charge Z has been used for the other elements.

ÅÇ 2 ÑÖ
1 In general, the reference data, which was employed in the

ÄÅ ÉÑ
2 parametrization of GFN-xTB, has been extended and used for

ÅÅ ÑÑ
A
fitting. The data consisted of molecular equilibrium and
ΘC Å αα Ñ
Å
− f XC ÅÅ3ΘC − δαβ ∑ ΘC ÑÑÑ
ÅÅ ÑÑ
nonequilibrium structures (energies and forces), harmonic

ÅÇ ÑÖ
αβ
force constants for up to medium sized systems (<30 atoms),
α (44) and noncovalent interaction energies and structures (mainly
potential energy curves and a few full optimizations). No
The last line in each of the previous equations describes the atomic partial charges or molecular dipole moments have been
AXC potential, whereas the remaining terms define the AES used in the fit.
potential experienced at position RC. It is stressed that the A mixed level of theory for the reference data is used. If
electric multipole moment integrals are given with origin at O systems from standard benchmark sets have been used, basis
= (0 0 0)T. Hence, VS(RC) and VD(RC) also include higher set extrapolated CCSD(T) energies are typically used, while
order potential terms due to the “shift” terms in the CAMM forces and structures were mostly computed by the more
definition (see eq 27). Analytical first nuclear derivatives are efficient, though sufficiently accurate, composite methods
not shown here but are derived and given in the SI. However, PBEh-3c4 or B97-3c.6 We used the TURBOMOLE suite of
it is noted here that the expressions for the analytical gradients programs101−103 (version 7.0) to conduct most of the ground
become much simpler if the multipole integrals are given with state DFT reference calculations and geometry optimizations.
origin at the respective atomic position. The exchange-correlation functional integration grid m4 and
The TB-Fock matrix contribution from self-consistent D4 is the SCF convergence criterion (10−7 Eh) along with the
given by resolution of the identity (RI) integral approximation104−106
D4 1 has been used.
Fκλ = − Sκλ(dA + dB), ∀ κ ∈ A, λ ∈ B The GFN2-xTB method parameters have first been
2 (45)
determined with a D3-variant for the dispersion energy.
where dA is given by (we are dropping the dependency on q Then the D4 parameters and reference values (i.e., qA,r) were
and CNcov for brevity) determined while keeping all other parameters fixed.
1660 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

Calculations for comparison with other semiempirical methods OM2-D3(BJ) and PM6-D3H4X.34 GFN2-xTB ranks
methods were conducted with the DFTB+107 (DFTB356 second from the considered methods and is outperformed only
with the 3OB parametrization58−60) and MOPAC16108 (PM6- slightly by its predecessor GFN-xTB. This likely results from
D3H4X35,38,39) codes. The DFTB3 method was used in the fact that the relative weight of geometries in the fitting
combination with the 3OB parametrization58−60 and D3- procedure has been larger for the GFN-xTB than for the
(BJ)71,86,97 dispersion correction. PM6-D3H4X, i.e., with zero- GFN2-xTB method. Furthermore, it should be noted that
damped D3 dispersion, as implemented in MOPAC16, is although GFN-xTB is mostly constructed from global and
used.38,86 Our standalone dftd3 code109 was used for the element-specific parameters, there still are a few element pair-
calculations of the D3(BJ) corrections. The Fermi smearing specific “fine-tuning” parameters (e.g., between N−H or H−H
technique (Tel = 300 K) has also been employed in the DFTB3 pairs). GFN2-xTB relies solely on element-specific and global
calculations. parameters; thus, its performance for these organic structures
In the following, we will use the abbreviations mean can be regarded as excellent. In Figure 3, the MADs for three
deviation (MD), mean absolute deviation (MAD), standard
deviation (SD), mean relative deviation (MRD), mean
unsigned relative deviation (MURD), standard relative
deviation (SRD), maximum unsigned deviation (MAX),
maximum unsigned relative deviation (MAXR), and regular-
ized relative root-mean-square error (RMSE).

3. RESULTS AND DISCUSSION


3.1. Molecular Structures. The ROT34 benchmark
set110,111 has become an established set to assess the
performance of quantum chemical methods to compute gas
phase structures of organic molecules. The quantity to be
compared is the spectroscopically accessible rotational
constant B0, which can quantum chemically be corrected for
nuclear vibrational effects to Be. This “clamped” nuclei
rotational constant can then be compared directly to the
local Born−Oppenheimer minimum energy structure of the
investigated method. As long as conformational changes can be
excluded, smaller rotational constants then typically indicate
elongated covalent bonds relative to the reference. Since the Figure 3. MADs in pm for bond lengths computed with GFN2-xTB,
isoamyl−acetate molecule in the ROT34 set was found to be GFN-xTB, and PM6-D3H4X (see SI for detailed results). The LB12
problematic w.r.t. conformational changes for many semi- and HMGB11 sets are taken from ref 4. For LB12, the MADs are
empirical methods, it is excluded in this work. The results for scaled by factor of 0.5. The transition metal containing systems
HAPPOD and KAMDOR in LB12 have been discarded for PM6-
this set are plotted in Figure 2. The data for the other SQM D3H4X. For GFN2-xTB, the S2+ 8 system in LB12 is neglected.
methods is taken from ref 34, whereas the detailed results for
GFN2-xTB are given in the SI. It has been discussed before
structure benchmark sets are shown, which are more difficult
that the tight-binding methods show significantly smaller
for quantum chemical methods. The LB124 consists of 12
standard relative deviations (SRDs) compared to the ZDO
molecules with an “unusually” long bond between two atoms.
HMGB114 contains heavy main group, and TMC32112
contains 50 bond distances in a total of 32 transition metal
complexes. GFN-xTB and GFN2-xTB clearly outperform
PM6-D3H4X for this purpose. In particular, it becomes
apparent that GFN2-xTB reproduces the LB12 and HMGB11
bond lengths particularly well. This reflects the consistent
element-specific parametrization in GFN2-xTB. It should,
however, be noted that there exists one outlier in the LB12 set
for GFN2-xTB (385 pm instead of 286 pm for the S82+
system), which is excluded in the statistical analysis.
Presumably, this overestimated bond length is caused by the
strong net charge of the system in combination with the higher
order but truncated multipole expansion (cf. GFN-xTB, which
shows a bond distance underestimated by about 55 pm). This
system is also difficult for many density functional approx-
imations showing similar large deviations from the reference.4
Figure 2. Normal distribution plots for the relative errors in the
computed equilibrium rotational constants Be for the ROT34
For the transition metal complexes (right-hand side of Figure
benchmark, with system 2 (isoamyl-acetate) excluded. The GFN- 3), the three semiempirical methods perform more similarly
xTB (MRD = 0.52%, SRD = 1.10%), DFTB-D3(BJ) (MRD = with GFN-xTB performing best and GFN2-xTB ranking
−1.26%, SRD = 1.28%), and PM6-D3H4X (MRD = −1.60%, SRD = second. For all sets considered here, it is observed that
2.50%) results are taken from ref 34. The values for GFN2-xTB are compared to GFN-xTBthe magnitude of the MD is reduced
computed in this work (MRD = 0.78%, SRD = 1.24%). for GFN2-xTB (see SI and ref 34).
1661 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

In Table 3, the statistical data for structures with focus on xTB clearly outperforms the other methods and marks a
noncovalent interactions are given. Here, the center-of-mass significant improvement over GFN-xTB. Hence, on average,
GFN2-xTB is the best of all SQM methods considered here,
Table 3. Comparison of Structures of Noncovalently Bound which indicates that the Hamiltonian terms responsible for
Systemsa noncovalent interactions are well-behaved.
3.2. Noncovalent Interaction Energies. In order to
GFN2-xTB GFN-xTB PM6-D3H4X DFTB3-D3(BJ)
investigate the performance for noncovalent interaction
S22113 (CMA distance in pm) energies, we consider benchmark sets from the large
MD −5 −5b 3b −11b GMTKN55 benchmark data set.117 The MADs for the
b
MAD 14 15 14b 14b noncovalent interaction energies are collected in Figures 4
SD 16 18b 21b 16b
b
MAX 32 50 59b 45b
114
S66x8 (CMA distance in %)
MRD −0.82 −0.63b 0.58b −2.87b
MURD 1.99 1.96b 2.01b 3.15b
b
SRD 2.58 2.68 2.57b 2.84b
b
MAXR 6.91 6.95 6.74b 10.93b
S22x5115 (CMA distance in %)
MRD −1.40 −0.76 0.03 −3.09
MURD 2.76 2.49 1.97 3.51
SRD 3.11 3.32 2.40 3.59
MAXR 6.85 6.78 4.44 10.70
X40x10116 (CMA distance in %)
MRD −0.17 −1.72 −1.39 −2.16
MURD 2.59 3.47 5.25 3.67
SRD 3.22 3.98 7.40 4.04
MAXR 7.82 8.85 20.09 11.53
a
The CMA distances of the S22113 complexes are obtained from a
Figure 4. Mean absolute deviations (MADs) in kcal mol−1 for the
free optimization of the complex geometries. The deviations are given
noncovalent association energies of different benchmark sets from the
in pm. The CMA distances of the S66,114 S22x5,115 and X40x10116
GMTKN55 database.117
complexes are derived from cubic spline interpolations of energies
computed on the differently separated structures. Here, the relative
errors in % are given. MRD = mean relative deviation, MURD = mean and 6. Except for the weak interactions in the alkane dimer set
unsigned relative deviation, SRD = standard relative deviation, MAXR ADIM6,117 GFN2-xTB always ranks first or second among the
= maximum unsigned relative deviation, MD = mean deviation, MAD considered methods. The benefit of including higher multipole
= mean absolute deviation, SD = standard deviation, MAX = moments is already observable for the often considered (also
maximum absolute deviation. bData taken from ref 34. for fitting) S22113,118 and S66114 sets. Here, the GFN2-xTB
method outperforms GFN-xTB and DFTB-D3(BJ) without
(CMA) distance deviation for fully optimized complexes from specific Hamiltonian terms to treat hydrogen bonds. Only the
the S22113 set are given. Furthermore, the relative deviations PM6-D3H4X method shows a vanishingly lower MAD in both
for the extrapolated CMA distances for the S66x8,114 S22x5,115 sets, which is expected as these sets had been used to adjust the
and X40x10116 sets are given. These are determined by cubic D3 and H4 parameters.38
spline interpolations of the respective interactions at the However, the advantages of including higher multipole
different distances, which has previously been done already for electrostatic terms become particularly obvious when looking
the S66x8 set.4,34 Regarding the first two sets (fully optimized at the HAL59116,117,119 and PNICO23117,120 benchmarks. In
S22 and S66x8), GFN2-xTB performs much like GFN-xTB but these benchmarks, the anisotropic electron density of the
shows slightly stronger underestimation for the CMA distances bonded halogen and pnicogen atoms is very important.
of S66x8. Furthermore, the deviations appear to be a bit more Though the monopole-based tight-binding methods can
systematic for both sets as indicated by the smaller standard perform well for either of the sets due their parametrization
relative deviation (SRD) for S66x8 and standard deviation or specific halogen bond corrections (in GFN-xTB), a
(SD) for S22. All of the SQM methods considered perform consistently good description is only observed for GFN2-
comparably well with DFTB3-D3(BJ), showing the largest xTB. This demonstrates the more physical behavior of GFN2-
deviations for the S66x8 set. xTB, which shows a better performance without special
The results obtained here are quite important also as an correction terms. PM6-D3H4X performs badly for both
assessment for the balance of the noncovalent interaction benchmark sets, particularly for HAL59 in which seven
terms. If repulsion, but in particular dispersion, and the systems yield an error >10 kcal mol−1.
(anisotropic) ES terms are not described in a balanced way, in The last set considered in Figure 4 is the WATER27 set,
particular the free optimization of the S22 structures should which consists of association energies of differently sized
yield larger deviations for the CMA distances. PM6-D3H4X neutral and charged water clusters. Here, the energies as well as
has been parametrized to the S66x8 energies and, as expected, the errors of the different methods are on average larger than
shows good performance for this set, though not better than for the other sets, which contain mostly smaller and less
GFN2-xTB. Its performance for S22x5 is remarkably good, for strongly interacting systems. The low MAD of GFN2-xTB for
which it shows the smallest deviations of the tested methods, the WATER27 set is noteworthy, in particular, since it is
followed by GFN-xTB and GFN2-xTB. For X40x10, GFN2- obtained without any additional H-bond specific correction
1662 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

Figure 5. Normal distribution plots for the errors in the computed interaction energies of the S66 benchmark.114 Plot (a) refers to the hydrogen
bonded (first 23) systems, whereas plot (b) refers to van der Waals-type bonded systems.

term. In fact, the MAD is even lower than for well-performing


generalized gradient density functional approximations
(GGAs) like BLYP-D3(BJ) (MAD of 4.1 kcal mol−1)117 or
some hybrid functionals (e.g., PBE0-D3(BJ) with an MAD of
5.9 kcal mol−1).117 Given that GFN2-xTB employs a minimal
basis set for hydrogen and oxygen, the excellent performance
for WATER27 suggests a well-working interplay of the tight-
binding Hamiltonian terms, particularly, of the multipole-
extended electrostatics and on-site third-order terms.
In Figure 5, the results for the S66 benchmark set are
processed in more detail. This figure shows the error
distribution plots subdivided for the hydrogen bonds (left-
hand side) and predominantly van der Waals-type interacting
systems (right-hand side). The MD for GFN2-xTB is
significantly improved compared to the monopole-based
GFN-xTB for the latter, which indicates the importance of
AES for such systems. DFTB3-D3(BJ) also works well there Figure 6. Mean absolute deviations (MADs) in kcal mol−1 for the
but shows larger deviations for the hydrogen bonded systems. noncovalent association energies of different benchmark sets from the
GMTKN55 database.117 These sets contain charged systems or heavy
Different from that, GFN2-xTB produces consistently small main group elements for which DFTB3 does not have parameters in
errors (|MD| < 1 kcal mol−1 for both subsets). PM6-D3H4X the 3OB parametrization58−60 (indicated by an asterisk).
also performs extremely well for both subsets.
In the following, other benchmark sets for noncovalent different elements and interaction types, we next turn our
interactions from the GMTKN55 database are considered (see attention to larger systems. For this purpose, the S30L set122 is
Figure 6). These sets consist of more difficult systems considered, which consists of 30 large noncovalently bound
(charged) or elements other than from the first and second neutral and charged complexes. The results for GFN2-xTB are
row. Though GFN-xTB performs slightly better for the compared with those of PM6-D3H4X and DFTB3-D3(BJ) in
cationic hydrogen bonded systems in CHB6 and the Figure 7. GFN-xTB performs similar to DFTB3-D3(BJ) and is
carbene-hydrogen bonded systems in CARBHB12, GFN2- excluded from this figure for clarity (see SI for the GFN-xTB
xTB overall performs best, and no outlier is observed, also for results). It is seen that GFN2-xTB closely resembles the
the heavy main group and noble gas elements. While DFTB3- reference values, i.e., domain-based pair natural orbital coupled
D3(BJ) shows low MADs for the ionic liquids (IL16) and cluster with singles, doubles, and perturbative triplesshort
anionic hydrogen bonds (AHB21), the lack of parameters DLPNO-CCSD(T)123extrapolated to the complete basis set
severely limits the applicability of the method. PM6-D3H4X (CBS) limit. Overall, it shows the smallest (in magnitude)
on the other hand consistently shows bigger MADs than GFN- MD, MAD, and SD of all methods considered and is roughly
xTB, as well as GFN2-xTB. A cross-check for the C1538 set on a par with some dispersion-corrected density functionals in
confirms the good performance of GFN2-xTB for charged a large quadruple-ζ basis set (e.g., B3LYP-D3(BJ)/def2-
complexes. For the I9 set,121 we find an underbinding QZVP).122 The nearly nondetectable deviation for the charged
tendency, likely due to missing polarization or the minimal systems is striking, and even the largest association energy of
basis set GFN2-xTB, cf. the performance of GFN-xTB about −135.5 kcal mol−1 (system 24) is reproduced to within
(double-ζ on hydrogen) or the chemical potential equalization <1% deviation, while the other semiempirical methods show
DFTB approach.28,121 Nevertheless, the errors are clearly more errors of about 20%. This is a reassuring result and reflects the
systematic than in GFN-xTB. The data for these sets can be well-described electrostatic and polarization interactions in
found in the Supporting Information. GFN2-xTB. Actually, the largest deviations are observed for
Having demonstrated the good performance of GFN2-xTB the van der Waals-dominated complexes of conjugated π-
for noncovalent interactions of small systems including systems (systems 7−12). GFN2-xTB overestimates the
1663 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

GMTKN55117 are given (see SI for detailed values). It can be


seen that the GFN2-xTB method ranks best in five out of eight
sets, namely, for the ACONF,124 Amino20x4,125 ICONF,117
PCONF21,117,126,127 and SCONF.117,128 Particularly positive is
the considerable improvement over GFN-xTB for more polar
and hydrogen-bonded systems in Amino20x4, PCONF21, and
SCONF. This suggests a generally better performance for
sugars and polypeptide systems. Furthermore, the GFN2-xTB
MAD of 1.6 kcal mol−1 for the ICONF stands out, as DFTB3-
D3(BJ), which ranks second, has an MAD that is larger by
almost 1 kcal mol−1. Also for the BUT14DIOL117,129 and
MCONF117,130 sets, the GFN2-xTB method performs quite
similar to the other SQM methods. The only outlier with an
MAD of 2.9 kcal mol−1 is the UPU23 set,117,131 which is about
twice as large as the MADs of either GFN-xTB or DFTB3-
D3(BJ). The latter both perform particularly well for this set,
Figure 7. Association energies of the S30L122 benchmark set whereas PM6-D3H4X is only slightly (<0.5 kcal mol−1) better
computed with GFN2-xTB, PM6-D3H4X, and DFTB3-D3(BJ). than GFN2-xTB. Since GFN2-xTB does not in general show
The values and statistical data are given in kcal mol−1. problems for phosphorus (see, e.g., PNICO23 results in Figure
4), the cause for the larger MAD for the UPU23 set is difficult
magnitude for the association energy of these complexes. This to nail down. Nevertheless, GFN2-xTB shows the best overall
behavior partially results from nonadditivity dispersion effects, performance for the conformer sets considered here.
and the ATM term, which is also included in GFN2-xTB (see Recently, a set of glucose and maltose conformers has been
Section 2.2.6), only partially compensates for this. This has compiled.132 These sugar conformers are particularly challeng-
already been observed and discussed for DFT-D3 methods ing due to the many different hydrogen bonded conformers. In
(see the comment in reference 93 of ref 75) and, hence, does Figure 9, conformational energies of both sets are plotted
not represent a weakness specific to the GFN2-xTB against the high level reference data (CBS extrapolated
Hamiltonian. DLPNO-CCSD(T) values).132 First of all, it is observed
3.3. Conformational Energies. For the description of that, on average, all SQM methods underestimate the
conformational energies, the balance between covalent and conformational energies, in particular of high-lying conformers.
(intramolecular) noncovalent forces is very important. The In agreement with the aforementioned results for the SCONF
proper energy ranking of conformers is essential for SQM set, GFN2-xTB outperforms the other methods in all statistical
methods, which can efficiently be used to sample the measures considered. Remarkable are the small SD values,
conformational space of chemically important, moderately which only DFTB3-D3(BJ) is nearly on a par with. The results
sized systems (<100−200 atoms). As mentioned before, for these sugar conformers are encouraging in that GFN2-xTB
determining the thermostatistically populated conformer- will be able to provide more reliable conformer-rotamer
rotamer ensemble for the calculation of spin-coupled ensembles than GFN-xTB even for polar and hydrogen
nuclear-magnetic resonance spectra has been a driving force bonded systems.
for the development of the GFN2-xTB method, and the AES 3.4. Rotational and Vibrational Free Energy Compu-
term in particular. In Figure 8, MAD bar plots for the tations. Due to computational efficiency in the calculation of
conformational energies of different benchmark sets from the gradients and numerical Hessians, a major application of
GFN2-xTB will likely be the computation of free energy
corrections in thermochemical studies. This has already been
pointed out for GFN-xTB,34 and we will crosscheck the
performance of GFN2-xTB against GFN-xTB and PBEh-3c
(frequencies scaled by 0.95).4 The latter is a a hybrid density
functional composite method, which is computationally
considerably more demanding than the semiempirical tight
binding methods (roughly two orders of magnitude). As a
reference method, another composite method called B97-3c,6
is used, which employs a GGA functional in a triple-ζ basis set.
As a “real-life” measure, differences for reactions in the
summed rotational and vibrational free energies termed
ΔGRRHO at T = 298.15 K are considered. While the rotational
part is a measure for the differences in optimized structures,
the vibrational contribution contains information about the
PES curvature around the minima. We use the systems from
the following sets, which are part of the GMTKN55
benchmark set database:117 AL2x6, DARC, HEAVYSB11,
Figure 8. Mean absolute deviations (MADs) in kcal mol−1 for ISOL24, TAUT15, ALK8, G2RC, and ISO34. Furthermore,
conformational energy benchmark sets. The structures and reference the recently proposed MOR41 set for closed-shell metal
data are taken from the GMTKN55 database.117 Detailed values are organic reactions is included.133 The detailed results are listed
listed in the SI. in the SI. It should be noted that translational free energy
1664 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

Figure 9. Correlation plots for (a) the conformational energies of 80 α-glucose and 76 β-glucose conformers and (b) the conformational energies
of 205 α-maltose conformers. The energies are given in kcal mol−1 and are computed with different SQM methods. The structures and reference
conformational energies are taken from ref 132.

contribution, which is independent of the electronic structure cluding transition metals complexes). This is particularly
method used, is purposely excluded, since the number of important, since such calculations with ab initio and DFT
reactants differs from the number of products for many of the methods quickly become the computational bottleneck in
reactions considered. This way, the magnitudes for ΔGRRHO typical workflows.
are more similar across the different sets. Furthermore, for all 3.5. Other Properties. In the previous section, we have
methods, the harmonic oscillator/free rotor interpolation from assessed GFN2-xTB for well-established benchmark sets that
ref 134 is applied for harmonic frequencies with magnitudes have also been used to study DFT approximations. Though
smaller than 50 cm−1. The results are plotted in Figure 10. The not training sets, these benchmark sets predominantly
coincided with the target properties of the GFN2-xTB method.
In this section, GFN2-xTB is tested for some off-target
properties. As in ref 34, we investigate the performance of
GFN2-xTB for covalently bonded diatomics in Figure 11.
Similar to GFN-xTB, we observe a systematic overestimation
of typical bond dissociation energies, while the minimum
positions are reproduced rather well (high level reference data

Figure 10. Deviation plot for the rotational and vibrational reaction
free energy computed with the tested methods compared to B97-3c.6
The MD as well as maximum and minimum deviation is shown for
each method. The boxes show the range in which the smallest half of
the errors is found. For PBEh-3c, the frequencies have been scaled by
0.95.

idea about this comparison is that both composite “3c”


methods represent comparably accurate methods, and which
one performs better will depend on the system. As expected,
GFN-xTB shows slighty larger errors and only has a somewhat
larger spread of errors. Similar is true also for the GFN2-xTB
method. In fact the 50% cases with smallest deviations for both
tight binding methods are found within a range that is about
twice as large compared to PBEh-3c. However, in absolute
Figure 11. Potential energy curves computed with GFN2-xTB for the
numbers, this error is on a 0.1 kcal mol−1 scale and hence dissociation of H2, F2, N2, and LiH. The electronic temperature
practically irrelevant, given that some deviations to B97-3c are treatment (Tel = 300 K) allows the homolytic dissociation without a
much larger for the tight binding methods as well as for PBEh- multireference treatment. The points mark the minimum energy
3c. Therefore, GFN2-xTB, just like GFN-xTB, should be a positions from high-level calculations.34,135,136 The energies are given
reasonable method of choice to routinely compute harmonic relative to the free atoms (S = 3/2 for nitrogen, S = 1/2 for the
frequencies for subsequent thermostatistical treatments (in- others).

1665 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

are marked with crosses). This agrees well with the may be well-suited to study biomolecular systems in aqueous
observations made in Section 3.1, i.e., that GFN2-xTB on solution.
average is on a par with GFN-xTB in the description of As a last test, we investigate the ability to reproduce the
covalently bonded molecular structures. At the same time, permanent electric dipole moments of small molecules which is
GFN2-xTB shares the property of overestimating covalent relevant for obtaining good long-range noncovalent inter-
bond energies. However, a slight difference compared to GFN- actions. Such a set with purely theoretical reference values has
xTB (cf. ref 34) is observable for the nonpolar, single-bonded recently been compiled by Hait and Head-Gordon.142
diatomics (H2 and F2). Here the overestimation is slightly less In Figure 13, the correlation plot for the computed
pronounced than with GFN-xTB, while the triple bond energy molecular dipole moments is shown. Due to missing
of N2 and the polar LiH bond energy show about the same
magnitude as GFN-xTB. It is to be determined in the future
how this will affect, e.g., the simulation of mass spectra or
reaction enthalpies (within the correction scheme applied in
ref 45).
Next, we turn our attention to some of the kinetics-oriented
benchmark sets of the GMTKN55117 database. In Figure 12,

Figure 12. Mean absolute deviations (MADs) for reaction barriers (in Figure 13. Permanent molecular dipole moments computed for open
kcal mol−1) computed with different semiempirical methods. Due to as well as closed shell systems with GFN2-xTB, GFN-xTB, and PM6-
missing parameters for silicon, DFTB3 calculations are not possible D3H4X. The benchmark set (structures and reference values) are
for two systems in BHDIV10117 and one system in the BHPERI137 taken from from ref 142. Detailed values are listed in the SI. The root-
set. Furthermore, one extreme outlier (PCl3) in the INV24138 was mean-square error (in %) is regularized with a value of 1 D.
found. In these cases, the respective systems are removed from the
statistical analysis for DFTB3-D3(BJ).
parameters for many elements, DFTB3 is excluded here.
While dipole moments have been used in the parametrization
the MADs for different barrier heights are presented. For three procedure of PM6,35 none were used in the fit of GFN-xTB
sets, systems are neglected in the statistical analysis for and GFN2-xTB. Nevertheless, PM6 does not show signifi-
DFTB3-D3(BJ), i.e., the Si-containing systems in BHDIV10 cantly better agreement with the high level reference. In fact,
and BHPERI137 (missing parameters) and also a severe outlier GFN2-xTB shows a lower MAD, as well as a lower regularized
in the INV24138 set. Here, the planar PCl3 transition state root-mean-square error (RMSE). As in ref 142, a value of 1 D
structure yielded a preposterously large repulsion energy (>104 is used for regularization. The RMSE is close to the one for
Hartree). MP2 (37.5%) as presented in the original work.142 This is an
Along all sets considered, the GFN2-xTB performs best. It encouraging finding, given that the set covers 14 different
shows the lowest MAD for the BHDIV10,117 BHROT27,117 elements and almost half of the systems have an open-shell
INV24,138 and PX13117,139 sets and the second lowest for the ground state. Our findings provide further support for the
WCPT18.117,140 Only for the barrier heights of pericyclic improved physics in the GFN2-xTB Hamiltonian, as well as for
reactions (BHPERI)137 is GFN2-xTB outperformed by the our parametrization strategy.
other SQM methods, though GFN-xTB and PM6-D3H4X are Finally, it should be mentioned that, in comparison to GFN-
only better by about 1−1.5 kcal mol−1. The performance of xTB, no increase in the computational time is expected for
GFN2-xTB for all sets in Figure 12 is remarkable, given that typical organic and biomolecular systems. As such, a single-
none of these (or similar systems) were used for fitting. In point energy calculation for crambin (641 atoms) takes 30 s
particular for the proton transfer sets PX13 and WCPT18, as with GFN2-xTB, but instead the same calculation requires 50 s
well as the single bond rotation set BHROT27, GFN2-xTB with GFN-xTB (single-core run on a laptop with a 1.6 GHz
shows MADs, which are comparable to or better than those of Intel Core i5 CPU and 8 GB RAM). Though the additional
the hybrid functional PBE0.117 Thisalong with the low integral evaluation (dipole and quadrupole one-electron
MAD for the Amino20x4,125 PCONF21,117,126,127 and integrals) in GFN2-xTB is more elaborate, the rate-
WATER27141 sets (see above)indicates that GFN2-xTB determining step in both methods is the diagonalization of
1666 DOI: 10.1021/acs.jctc.8b01176
J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

the tight-binding Hamiltonian matrix. Due to the extra s- Reoptimized geometries of the structure benchmark sets
function for hydrogen,34 the matrix dimension in GFN-xTB is (ZIP)


significantly larger for typical organic and biomolecular systems
(by a factor of about 1.5), and hence the computation time is AUTHOR INFORMATION
even reduced for GFN2-xTB compared to GFN-xTB.
Corresponding Authors
However, for some inorganic clusters (e.g., alkaline earth
oxides with >30 atoms), GFN2-xTB occasionally shows SCF *(C.B.) E-mail: christoph.bannwarth@stanford.edu.
convergence problems, which are not observed for GFN-xTB. *(S.G.) E-mail: grimme@thch.uni-bonn.de Phone: +49-228/
This presumably results from the multipole expansion and 73-2351.
partially polarized basis sets. This problem could not entirely ORCID
be removed without sacrificing the overall higher accuracy of Christoph Bannwarth: 0000-0003-3242-496X
GFN2-xTB. Though such problematic cases are rare, they Stefan Grimme: 0000-0002-5844-4371
should be kept in mind when dealing with inorganic clusters. Funding
The GFN2-xTB method is implemented in the standalone This work was supported by the DFG in the framework of the
xtb code, which can be requested from the authors.143 “Gottfried Wilhelm Leibniz Prize” awarded to S.G.
Notes
4. CONCLUSIONS The authors declare no competing financial interest.
We developed a generally applicable semiempirical quantum
mechanical method, termed GFN2-xTB, which represents the
first broadly parametrized tight-binding method to include
■ ACKNOWLEDGMENTS
The authors would like to thank Prof. Dr. Thomas Bredow, Dr.
electrostatic and exchange-correlation Hamiltonian terms Jan Gerit Brandenburg, and Dr. Andreas Hansen for helpful
beyond the monopole approximation. The method is free discussions during the development of the GFN2-xTB method.
from any hydrogen or halogen bond specific corrections, which Furthermore, the authors are grateful to Eike Caldeweyher for
are a standard add-on in other contemporary semiempirical providing routines, which are used in the self-consistent D4
schemes. Furthermore, the self-consistent D4 dispersion model treatment.
is an inherent part of the GFN2-xTB method and allows
efficiently incorporating electronic structure effects on the two-
body dispersion energy. The GFN2-xTB method relies strictly
■ DEDICATION
This work is dedicated to Professor Walter Thiel on the
on element-specific and global parameters and is parametrized occasion of his forthcoming 70th birthday and in appreciation
for all elements up to radon (Z = 86). Like for its predecessor, of his seminal work on approximate molecular orbital and
GFN-xTB, the parameters were fitted to yield reasonable QM/MM methods.


structures, vibrational frequencies, and noncovalent interac-
tions for molecules across the periodic table. The main focus of REFERENCES
this method are organic, organometallic, and biochemical
(1) Houk, K. N.; Liu, F. Holy Grails for Computational Organic
systems on the order of a few thousand atoms. In particular, Chemistry and Biochemistry. Acc. Chem. Res. 2017, 50, 539−543.
the greatly improved noncovalent interactions will likely trigger (2) Grimme, S.; Schreiner, P. R. Computational Chemistry: The
structural searches and studies of conformational and protein− Fate of Current Methods and Future Challenges. Angew. Chem., Int.
ligand studies in the near future. Ed. 2018, 57, 4170−4176.
Apart from these, the improved electrostatics and more (3) Sure, R.; Grimme, S. Corrected small basis set Hartree-Fock
consistent parametrization procedure has provided a method method for large systems. J. Comput. Chem. 2013, 34, 1672−1685.
that better reproduces the electronic density compared to (4) Grimme, S.; Brandenburg, J. G.; Bannwarth, C.; Hansen, A.
other semiempirical methods. The method may thus qualify to Consistent structures and interactions by density functional theory
be used in docking procedures by providing reasonable with small atomic orbital basis sets. J. Chem. Phys. 2015, 143, 054107.
(5) Brandenburg, J. G.; Caldeweyher, E.; Grimme, S. Screened
electrostatic potentials.
exchange hybrid density functional for accurate and efficient
However, as a difficult-to-quantify drawback compared to structures and interaction energies. Phys. Chem. Chem. Phys. 2016,
GFN-xTB we finally note sometimes less robust SCF 18, 15519−15523.
convergence in particular for metallic systems or polar (6) Brandenburg, J. G.; Bannwarth, C.; Hansen, A.; Grimme, S.
inorganic clusters probably caused by the short-range part of B97−3c: A revised low-cost variant of the B97-D density functional
the AES potential. Further work to improve the GFN family of method. J. Chem. Phys. 2018, 148, 064104.
methods for this field of application is in progress. (7) Otero-de-la Roza, A.; DiLabio, G. A. Transferable Atom-


Centered Potentials for the Correction of Basis Set Incompleteness
Errors in Density-Functional Theory. J. Chem. Theory Comput. 2017,
ASSOCIATED CONTENT 13, 3505−3524.
*
S Supporting Information (8) Witte, J.; Neaton, J. B.; Head-Gordon, M. Effective empirical
The Supporting Information is available free of charge on the corrections for basis set superposition error in the def2-SVPD basis:
ACS Publications website at DOI: 10.1021/acs.jctc.8b01176. gCP and DFT-C. J. Chem. Phys. 2017, 146, 234105.
(9) Hostaš, J.; Ř ezác,̌ J. Accurate DFT-D3 Calculations in a Small
The derivations for the newly developed AES and AXC, Basis Set. J. Chem. Theory Comput. 2017, 13, 3575−3585.
as well as the self-consistent D4 dispersion energy (10) Kulik, H. J.; Seelam, N.; Mar, B. D.; Martinez, T. J. Adapting
expressions, along with the respective TB-Fock matrix DFT+U for the Chemically Motivated Correction of Minimal Basis
Set Incompleteness. J. Phys. Chem. A 2016, 120, 5939−5949.
contributions and nuclear gradients, element-specific
(11) Salsbury, F. R. Molecular dynamics simulations of protein
parameters and the detailed results obtained on the dynamics and their relevance to drug discovery. Curr. Opin.
considered test sets (PDF) Pharmacol. 2010, 10, 738−744.

1667 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

(12) Kussmann, J.; Ochsenfeld, C. Pre-selective screening for matrix (33) Grimme, S.; Bannwarth, C. Ultra-fast computation of electronic
elements in linear-scaling exact exchange calculations. J. Chem. Phys. spectra for large systems by tight-binding based simplified Tamm-
2013, 138, 134114. Dancoff approximation (sTDA-xTB). J. Chem. Phys. 2016, 145,
(13) Ufimtsev, I. S.; Martinez, T. J. Quantum Chemistry on 054103.
Graphical Processing Units. 3. Analytical Energy Gradients, Geometry (34) Grimme, S.; Bannwarth, C.; Shushkov, P. A Robust and
Optimization, and First Principles Molecular Dynamics. J. Chem. Accurate Tight-Binding Quantum Chemical Method for Structures,
Theory Comput. 2009, 5, 2619−2628. Vibrational Frequencies, and Noncovalent Interactions of Large
(14) Luehr, N.; Ufimtsev, I. S.; Martinez, T. J. Dynamic Precision for Molecular Systems Parametrized for All spd-Block Elements (Z = 1 −
Electron Repulsion Integral Evaluation on Graphical Processing Units 86). J. Chem. Theory Comput. 2017, 13, 1989−2009.
(GPUs). J. Chem. Theory Comput. 2011, 7, 949−954. (35) Stewart, J. J. P. Optimization of parameters for semiempirical
(15) Schütt, O.; Messmer, P.; Hutter, J.; VandeVondele, J. Electronic methods V: Modification of NDDO approximations and application
Structure Calculations on Graphics Processing Units; John Wiley & Sons, to 70 elements. J. Mol. Model. 2007, 13, 1173.
Ltd.: 2016; pp 173−190. (36) Korth, M.; Pitoňaḱ , M.; Ř ezác,̌ J.; Hobza, P. A Transferable H-
(16) Yasuda, K.; Maruoka, H. Efficient calculation of two electron Bonding Correction for Semiempirical Quantum-Chemical Methods.
integrals for high angular basis functions. Int. J. Quantum Chem. 2014, J. Chem. Theory Comput. 2010, 6, 344−352.
114, 543−552. (37) Kromann, J. C.; Christensen, A. S.; Steinmann, C.; Korth, M.;
(17) Asadchev, A.; Gordon, M. S. New Multithreaded Hybrid CPU/ Jensen, J. H. A third-generation dispersion and third-generation
GPU Approach to Hartree-Fock. J. Chem. Theory Comput. 2012, 8, hydrogen bonding corrected PM6 method: PM6-D3H+. PeerJ 2014,
4166−4176. 2, e449.
(18) Yasuda, K. Accelerating Density Functional Calculations with (38) Ř ezác,̌ J.; Hobza, P. Advanced Corrections of Hydrogen
Graphics Processing Unit. J. Chem. Theory Comput. 2008, 4, 1230− Bonding and Dispersion for Semiempirical Quantum Mechanical
1236. Methods. J. Chem. Theory Comput. 2012, 8, 141−151.
(19) Maia, J. D. C.; Urquiza Carvalho, G. A.; Mangueira, C. P.; (39) Brahmkshatriya, P. S.; Dobeš, P.; Fanfrlík, J.; Ř ezác,̌ J.; Paruch,
Santana, S. R.; Cabral, L. A. F.; Rocha, G. B. GPU Linear Algebra K.; Bronowska, A.; Lepšík, M.; Hobza, P. Quantum Mechanical
Libraries and GPGPU Programming for Accelerating MOPAC Scoring: Structural and Energetic Insights into Cyclin-Dependent
Semiempirical Quantum Chemistry Calculations. J. Chem. Theory Kinase 2 Inhibition by Pyrazolo[1,5-a]pyrimidines. Curr. Comput.-
Comput. 2012, 8, 3072−3081. Aided Drug Des. 2013, 9, 118−129.
(20) Wu, X.; Koslowski, A.; Thiel, W. Semiempirical Quantum (40) Saito, T.; Kitagawa, Y.; Takano, Y. Reparameterization of PM6
Chemical Calculations Accelerated on a Hybrid Multicore CPU− Applied to Organic Diradical Molecules. J. Phys. Chem. A 2016, 120,
GPU Computing Platform. J. Chem. Theory Comput. 2012, 8, 2272− 8750−8760.
2281. (41) Weber, W.; Thiel, W. Orthogonalization corrections for
(21) Kussmann, J.; Ochsenfeld, C. Hybrid CPU/GPU Integral semiempirical methods. Theor. Chem. Acc. 2000, 103, 495−506.
Engine for Strong-Scaling Ab Initio Methods. J. Chem. Theory Comput. (42) Tuttle, T.; Thiel, W. OMx-D: semiempirical methods with
2017, 13, 3153−3159. orthogonalization and dispersion corrections. Implementation and
(22) Kalinowski, J.; Wennmohs, F.; Neese, F. Arbitrary Angular biochemical application. Phys. Chem. Chem. Phys. 2008, 10, 2159−
Momentum Electron Repulsion Integrals with Graphical Processing 2166.
Units: Application to the Resolution of Identity Hartree-Fock (43) Dral, P. O.; Wu, X.; Spörkel, L.; Koslowski, A.; Weber, W.;
Method. J. Chem. Theory Comput. 2017, 13, 3160−3170. Steiger, R.; Scholten, M.; Thiel, W. Semiempirical Quantum-
(23) van Schoot, H.; Visscher, L. Electronic Structure Calculations on Chemical Orthogonalization-Corrected Methods: Theory, Implemen-
Graphics Processing Units; John Wiley & Sons, Ltd.: 2016; pp 101− tation, and Parameters. J. Chem. Theory Comput. 2016, 12, 1082−
114. 1096.
(24) Ufimtsev, I. S.; Martínez, T. J. Graphical Processing Units for (44) Koslowski, A.; Beck, M. E.; Thiel, W. Implementation of a
Quantum Chemistry. Comput. Sci. Eng. 2008, 10, 26−34. general multireference configuration interaction procedure with
(25) Jakowski, J.; Irle, S.; Morokuma, K. In GPU Computing Gems analytic gradients in a semiempirical context using the graphical
Emerald Edition; Hwu, W.-m. W., Ed.; Applications of GPU unitary group approach. J. Comput. Chem. 2003, 24, 714−726.
Computing Series; Morgan Kaufmann: Boston, 2011; pp 59−73. (45) Kromann, J.; Welford, A.; Christensen, A. S.; Jensen, J. Random
(26) Thiel, W. Semiempirical quantum−chemical methods. WIREs Versus Systematic Errors in Reaction Enthalpies Computed Using
Comput. Mol. Sci. 2014, 4, 145−157. Semi-empirical and Minimal Basis Set Methods. ACS Omega 2018, 3,
(27) Yilmazer, N. D.; Korth, M. Enhanced semiempirical QM 4372−4377.
methods for biomolecular interactions. Comput. Struct. Biotechnol. J. (46) Bikadi, Z.; Hazai, E. Application of the PM6 semi-empirical
2015, 13, 169−175. method to modeling proteins enhances docking accuracy of
(28) Christensen, A. S.; Kubař, T.; Cui, Q.; Elstner, M. AutoDock. J. Cheminf. 2009, 1, 15.
Semiempirical Quantum Mechanical Methods for Noncovalent (47) Stewart, J. J. P. Application of the PM6 method to modeling
Interactions for Chemical and Biochemical Applications. Chem. Rev. proteins. J. Mol. Model. 2009, 15, 765−805.
2016, 116, 5301−5337. (48) Miriyala, V. M.; Ř ezác,̌ J. Testing Semiempirical Quantum
(29) Gonzalez-Lafont, A.; Truong, T. N.; Truhlar, D. G. Direct Mechanical Methods on a Data Set of Interaction Energies Mapping
dynamics calculations with NDDO (neglect of diatomic differential Repulsive Contacts in Organic Molecules. J. Phys. Chem. A 2018, 122,
overlap) molecular orbital theory with specific reaction parameters. J. 2801−2808.
Phys. Chem. 1991, 95, 4618−4627. (49) Kazaryan, A.; Lan, Z.; Schäfer, L. V.; Thiel, W.; Filatov, M.
(30) Storer, J. W.; Giesen, D. J.; Cramer, C. J.; Truhlar, D. G. Class Surface Hopping Excited-State Dynamics Study of the Photo-
IV charge models: A new semiempirical approach in quantum isomerization of a Light-Driven Fluorene Molecular Rotary Motor.
chemistry. J. Comput.-Aided Mol. Des. 1995, 9, 87−110. J. Chem. Theory Comput. 2011, 7, 2189−2199.
(31) Jakalian, A.; Jack, D. B.; Bayly, C. I. Fast, efficient generation of (50) Spörkel, L.; Cui, G.; Thiel, W. Photodynamics of Schiff Base
high-quality atomic charges. AM1-BCC model: II. Parameterization Salicylideneaniline: Trajectory Surface-Hopping Simulations. J. Phys.
and validation. J. Comput. Chem. 2002, 23, 1623−1641. Chem. A 2013, 117, 4574−4583.
(32) Cramer, C. J.; Truhlar, D. G. AM1-SM2 and PM3-SM3 (51) Spörkel, L.; Cui, G.; Koslowski, A.; Thiel, W. Nonequilibrium
parameterized SCF solvation models for free energies in aqueous H/D Isotope Effects from Trajectory-Based Nonadiabatic Dynamics.
solution. J. Comput.-Aided Mol. Des. 1992, 6, 629−666. J. Phys. Chem. A 2014, 118, 152−157.

1668 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

(52) Spörkel, L.; Jankowska, J.; Thiel, W. Photoswitching of (70) Hehre, W. J.; Stewart, R. F.; Pople, J. A. Self-Consistent
Salicylidene Methylamine: A Theoretical Photodynamics Study. J. Molecular-Orbital Methods. I. Use of Gaussian Expansions of Slater-
Phys. Chem. B 2015, 119, 2702−2710. Type Atomic Orbitals. J. Chem. Phys. 1969, 51, 2657−2664.
(53) Dokukina, I.; Marian, C. M.; Weingart, O. New Perspectives on (71) Brandenburg, J. G.; Grimme, S. Dispersion Corrected Hartree-
an Old Issue: A Comparative MS-CASPT2 and OM2-MRCI Study of Fock and Density Functional Theory for Organic Crystal Structure
Polyenes and Protonated Schiff Bases. Photochem. Photobiol. 2017, 93, Prediction. Top. Curr. Chem. 2013, 345, 1−23.
1345−1355. (72) Humeniuk, A.; Mitrić, R. Long-range correction for tight-
(54) Elstner, M.; Porezag, D.; Jungnickel, G.; Elsner, J.; Haugk, M.; binding TD-DFT. J. Chem. Phys. 2015, 143, 134120.
Frauenheim, T.; Suhai, S.; Seifert, G. Self-consistent-charge density- (73) Lutsker, V.; Aradi, B.; Niehaus, T. A. Implementation and
functional tight-binding method for simulations of complex materials benchmark of a long-range corrected functional in the density
properties. Phys. Rev. B: Condens. Matter Mater. Phys. 1998, 58, 7260− functional based tight-binding method. J. Chem. Phys. 2015, 143,
7268. 184107.
(55) Yang, Y.; Yu, H.; York, D.; Cui, Q.; Elstner, M. Extension of the (74) Kranz, J. J.; Elstner, M.; Aradi, B.; Frauenheim, T.; Lutsker, V.;
Self-Consistent-Charge Density-Functional Tight-Binding Method: Garcia, A. D.; Niehaus, T. A. Time-Dependent Extension of the Long-
Third-Order Expansion of the Density Functional Theory Total Range Corrected Density Functional Based Tight-Binding Method. J.
Energy and Introduction of a Modified Effective Coulomb Chem. Theory Comput. 2017, 13, 1737−1747.
Interaction. J. Phys. Chem. A 2007, 111, 10861−10873. (75) Grimme, S.; Hansen, A.; Brandenburg, J. G.; Bannwarth, C.
(56) Gaus, M.; Cui, Q.; Elstner, M. DFTB3: Extension of the Self- Dispersion-Corrected Mean-Field Electronic Structure Methods.
Consistent-Charge Density-Functional Tight-Binding Method (SCC- Chem. Rev. 2016, 116, 5105−5154.
DFTB). J. Chem. Theory Comput. 2011, 7, 931−948. (76) Vydrov, O. A.; Van Voorhis, T. Nonlocal van der Waals density
(57) Niehaus, T. A.; Suhai, S.; Della Sala, F.; Lugli, P.; Elstner, M.; functional: the simpler the better. J. Chem. Phys. 2010, 133, 244103.
Seifert, G.; Frauenheim, T. Tight-binding approach to time-depend- (77) Caldeweyher, E.; Ehlert, S.; Hansen, A.; Neugebauer, H.;
ent density-functional response theory. Phys. Rev. B: Condens. Matter Bannwarth, C.; Grimme, S. A Generally Applicable Atomic-Charge
Mater. Phys. 2001, 63, 085108. Dependent London Dispersion Correction Scheme. ChemRxiv.
(58) Gaus, M.; Goez, A.; Elstner, M. Parametrization and Preprint. DOI: 10.26434/chemrxiv.7430216.v2.
Benchmark of DFTB3 for Organic Molecules. J. Chem. Theory (78) Caldeweyher, E.; Bannwarth, C.; Grimme, S. Extension of the
Comput. 2013, 9, 338−354. D3 dispersion coefficient model. J. Chem. Phys. 2017, 147, 034112.
(59) Gaus, M.; Lu, X.; Elstner, M.; Cui, Q. Parameterization of (79) Nishimoto, K.; Mataga, N. Electronic Structure and Spectra of
DFTB3/3OB for Sulfur and Phosphorus for Chemical and Biological Some Nitrogen Heterocycles. Z. Phys. Chem. 1957, 12, 335−338.
Applications. J. Chem. Theory Comput. 2014, 10, 1518−1537. (80) Ohno, K. Some Remarks on the Pariser-Parr-Pople Method.
(60) Kubillus, M.; Kubař, T.; Gaus, M.; Ř ezác,̌ J.; Elstner, M. Theor. Chim. Act. 1964, 2, 219.
Parameterization of the DFTB3Method for Br, Ca, Cl, F, I, K, and Na (81) Klopman, G. A Semiempirical Treatment of Molecular
in Organic and Biological Systems. J. Chem. Theory Comput. 2015, 11, Structures. II. Molecular Terms and Application to Diatomic
332−342. Molecules. J. Am. Chem. Soc. 1964, 86, 4550.
(61) Gaus, M.; Jin, H.; Demapan, D.; Christensen, A. S.; Goyal, P.; (82) Grimme, S. A General Quantum Mechanically Derived Force
Elstner, M.; Cui, Q. DFTB3 Parametrization for Copper: The Field (QMDFF) for Molecules and Condensed Phase Simulations. J.
Importance of Orbital Angular Momentum Dependence of Hubbard Chem. Theory Comput. 2014, 10, 4497−4514.
Parameters. J. Chem. Theory Comput. 2015, 11, 4205−4219. (83) Mermin, N. D. Thermal Properties of the Inhomogeneous
(62) Vujović, M.; Huynh, M.; Steiner, S.; Garcia-Fernandez, P.; Electron Gas. Phys. Rev. 1965, 137, A1441−A1443.
Elstner, M.; Cui, Q.; Gruden, M. Exploring the applicability of density (84) Roothaan, C. C. J. New Developments in Molecular Orbital
functional tight binding to transition metal ions. Parameterization for Theory. Rev. Mod. Phys. 1951, 23, 69−89.
nickel with the spin-polarized DFTB3 model. J. Comput. Chem. 2019, (85) Hall, G. G. Molecular Orbital Theory of Chemical Valency.
40, 400. VIII. A Method of Calculating Ionization Potentials. Proc. R. Soc.
(63) Bursch, M.; Hansen, A.; Grimme, S. Fast and Reasonable London, Ser. A 1951, 205, 541−552.
Geometry Optimization of Lanthanoid Complexes with an Extended (86) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and
Tight Binding Quantum Chemical Method. Inorg. Chem. 2017, 56, accurate ab initio parametrization of density functional dispersion
12485−12491. correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010,
(64) Struch, N.; Bannwarth, C.; Ronson, T. K.; Lorenz, Y.; Mienert, 132, 154104.
B.; Wagner, N.; Engeser, M.; Bill, E.; Puttreddy, R.; Rissanen, K.; (87) Pyykkö, P.; Atsumi, M. Molecular Single-Bond Covalent Radii
Beck, J.; Grimme, S.; Nitschke, J. R.; Lützen, A. An Octanuclear for Elements 1−118. Chem. - Eur. J. 2009, 15, 186−197.
Metallosupramolecular Cage Designed To Exhibit Spin-Crossover (88) Mantina, M.; Valero, R.; Cramer, C. J.; Truhlar, D. G. In CRC
Behavior. Angew. Chem., Int. Ed. 2017, 56, 4930−4935. Handbook of Chemistry and Physics, 91st ed.; Haynes, W. M., Ed.;
(65) Seibert, J.; Bannwarth, C.; Grimme, S. Biomolecular Structure CRC Press: Boca Raton, FL, 2010; pp 9-49−9-50.
Information from High-Speed Quantum Mechanical Electronic (89) Köhler, C.; Seifert, G.; Frauenheim, T. Density functional based
Spectra Calculation. J. Am. Chem. Soc. 2017, 139, 11682−11685. calculations for Fen (n ≤ 32). Chem. Phys. 2005, 309, 23−31.
(66) Pracht, P.; Bauer, C. A.; Grimme, S. Automated and efficient (90) Grimme, S. A simplified Tamm-Dancoff density functional
quantum chemical determination and energetic ranking of molecular approach for the electronic excitation spectra of very large molecules.
protonation sites. J. Comput. Chem. 2017, 38, 2618−2631. J. Chem. Phys. 2013, 138, 244104.
(67) Grimme, S.; Bannwarth, C.; Dohm, S.; Hansen, A.; Pisarek, J.; (91) Finnis, M. W.; Paxton, A. T.; Methfessel, M.; van Schilfgaarde,
Pracht, P.; Seibert, J.; Neese, F. Fully Automated Quantum- M. Crystal Structures of Zirconia from First Principles and Self-
Chemistry-Based Computation of Spin-Spin-Coupled Nuclear Mag- Consistent Tight Binding. Phys. Rev. Lett. 1998, 81, 5149−5152.
netic Resonance Spectra. Angew. Chem., Int. Ed. 2017, 56, 14763− (92) Bodrog, Z.; Aradi, B. Possible improvements to the self-
14769. consistent-charges density-functional tight-binding method within the
(68) Asgeirsson, V.; Bauer, C. A.; Grimme, S. Quantum chemical second order. Phys. Status Solidi B 2012, 249, 259−269.
calculation of electron ionization mass spectra for general organic and (93) Boleininger, M.; Guilbert, A. A.; Horsfield, A. P. Gaussian
inorganic molecules. Chem. Sci. 2017, 8, 4879−4895. polarizable-ion tight binding. J. Chem. Phys. 2016, 145, 144103.
(69) Grimme, S.; Bannwarth, C.; Caldeweyher, E.; Pisarek, J.; (94) Köster, A. M.; Leboeuf, M.; Salahub, D. R. In Molecular
Hansen, A. A general intermolecular force field based on tight-binding Electrostatic Potentials; Murray, J. S., Sen, K., Eds.; Theoretical and
quantum chemical calculations. J. Chem. Phys. 2017, 147, 161708. Computational Chemistry; Elsevier: 1996; Vol. 3, pp 105−142.

1669 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

(95) Axilrod, B. M.; Teller, E. Interaction of the van der Waals Type istry, kinetics and noncovalent interactions. Phys. Chem. Chem. Phys.
Between Three Atoms. J. Chem. Phys. 1943, 11, 299−300. 2017, 19, 32184−32215.
(96) Muto, Y. Force between nonpolar molecules. Proc. Phys. Math. (118) Marshall, M. S.; Burns, L. A.; Sherrill, C. D. Basis set
Soc. Jpn. 1943, 17, 629. convergence of the coupled-cluster correction, δCCSD(T)
MP2 : Best practices
(97) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping for benchmarking non-covalent interactions and the attendant
Function in Dispersion Corrected Density Functional Theory. J. revision of the S22, NBC10, HBC6, and HSG databases. J. Chem.
Comput. Chem. 2011, 32, 1456−1465. Phys. 2011, 135, 194102.
(98) Ghosh, D. C.; Islam, N. Semiempirical evaluation of the global (119) Kozuch, S.; Martin, J. M. L. Halogen Bonds: Benchmarks and
hardness of the atoms of 103 elements of the periodic table using the Theoretical Analysis. J. Chem. Theory Comput. 2013, 9, 1918−1931.
most probable radii as their size descriptors. Int. J. Quantum Chem. (120) Setiawan, D.; Kraka, E.; Cremer, D. Strength of the Pnicogen
2009, 110, 1206−1213. Bond in Complexes Involving Group Va Elements N, P, and As. J.
(99) Levenberg, K. A Method for the Solution of Certain Non-linear Phys. Chem. A 2015, 119, 1642−1656.
Problems in Least Squares. Q. Appl. Math. 1944, 2, 164−168. (121) Christensen, A. S.; Elstner, M.; Cui, Q. Improving
(100) Marquardt, D. An Algorithm for Least-Squares Estimation of intermolecular interactions in DFTB3 using extended polarization
Nonlinear Parameters. J. Soc. Ind. Appl. Math. 1963, 11, 431−441. from chemical-potential equalization. J. Chem. Phys. 2015, 143,
(101) TURBOMOLE, V7.0 2015, a development of University of 084123.
Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989−2007; (122) Sure, R.; Grimme, S. Comprehensive Benchmark of
TURBOMOLE GmbH: since 2007; available from http://www. Association (Free) Energies of Realistic Host-Guest Complexes. J.
turbomole.com. Chem. Theory Comput. 2015, 11, 3785−3801.
(102) Ahlrichs, R.; Bär, M.; Häser, M.; Horn, H.; Kölmel, C. (123) Riplinger, C.; Sandhoefer, B.; Hansen, A.; Neese, F. Natural
Electronic Structure Calculations on Workstation Computers: The triple excitations in local coupled cluster calculations with pair natural
Program System Turbomole. Chem. Phys. Lett. 1989, 162, 165−169. orbitals. J. Chem. Phys. 2013, 139, 134101.
(103) Furche, F.; Ahlrichs, R.; Hättig, C.; Klopper, W.; Sierka, M.; (124) Gruzman, D.; Karton, A.; Martin, J. M. L. Performance of Ab
Weigend, F. Turbomole. WIREs Comput. Mol. Sci. 2014, 4, 91−100. Initio and Density Functional Methods for Conformational Equilibria
(104) Vahtras, O.; Almlö f, J.; Feyereisen, M. W. Integral of CnH2n+2 Alkane Isomers (n = 4 − 8). J. Phys. Chem. A 2009, 113,
approximations for LCAO-SCF calculations. Chem. Phys. Lett. 1993, 11974−11983.
213, 514−518. (125) Kesharwani, M. K.; Karton, A.; Martin, J. M. L. Benchmark ab
(105) Eichkorn, K.; Weigend, F.; Treutler, O.; Ahlrichs, R. Auxiliary Initio Conformational Energies for the Proteinogenic Amino Acids
basis sets for main row atoms and transition metals and their use to through Explicitly Correlated Methods. Assessment of Density
approximate Coulomb potentials. Theor. Chem. Acc. 1997, 97, 119− Functional Methods. J. Chem. Theory Comput. 2016, 12, 444−454.
(126) Ř eha, D.; Valdés, H.; Vondrásě k, J.; Hobza, P.; Abu-Riziq, A.;
124.
(106) Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Crews, B.; de Vries, M. S. Structure and IR Spectrum of Phenylalanyl-
Glycyl-Glycine Tripetide in the Gas-Phase: IR/UV Experiments, Ab
Phys. Chem. Chem. Phys. 2006, 8, 1057−1065.
Initio Quantum Chemical Calculations, and Molecular Dynamic
(107) Frauenheim, T. DFTB+ (Density Functional based Tight
Simulations. Chem. - Eur. J. 2005, 11, 6803−6817.
Binding); Universität Bremen: Bremen, Germany, 2008; http://www.
(127) Goerigk, L.; Karton, A.; Martin, J. M. L.; Radom, L. Accurate
dftb.org (August 9, 2014).
quantum chemical energies for tetrapeptide conformations: why MP2
(108) Stewart, J. J. P. MOPAC2016; Stewart Computational
data with an insufficient basis set should be handled with caution.
Chemistry: Colorado Springs, CO, USA, 2016; http://
Phys. Chem. Chem. Phys. 2013, 15, 7028−7031.
OpenMOPAC.net (August 16, 2016). (128) Csonka, G. I.; French, A. D.; Johnson, G. P.; Stortz, C. A.
(109) See http://www.thch.uni-bonn.de/.
Evaluation of Density Functionals and Basis Sets for Carbohydrates. J.
(110) Grimme, S.; Steinmetz, M. Effects of London Dispersion
Chem. Theory Comput. 2009, 5, 679−692.
Correction in Density Functional Theory on the Structures of (129) Kozuch, S.; Bachrach, S. M.; Martin, J. M. Conformational
Organic Molecules in the Gas Phase. Phys. Chem. Chem. Phys. 2013, Equilibria in Butane-1,4-diol: A Benchmark of a Prototypical System
15, 16031−16042. with Strong Intramolecular H-bonds. J. Phys. Chem. A 2014, 118,
(111) Risthaus, T.; Steinmetz, M.; Grimme, S. Implementation of 293−303.
Nuclear Gradients of Range-Separated Hybrid Density Functionals (130) Kozuch, S.; Martin, J. M. L. Spin-component-scaled double
and Benchmarking on Rotational Constants for Organic Molecules. J. hybrids: An extensive search for the best fifth-rung functionals
Comput. Chem. 2014, 35, 1509−1516. blending DFT and perturbation theory. J. Comput. Chem. 2013, 34,
(112) Bühl, M.; Kabrede, H. Geometries of Transition-Metal 2327−2344.
Complexes from Density-Functional Theory. J. Chem. Theory Comput. (131) Kruse, H.; Mladek, A.; Gkionis, K.; Hansen, A.; Grimme, S.;
2006, 2, 1282−1290. Sponer, J. Quantum Chemical Benchmark Study on 46 RNA
(113) Jurečka, P.; Š poner, J.; Cerny, J.; Hobza, P. Benchmark Backbone Families Using a Dinucleotide Unit. J. Chem. Theory
database of accurate (MP2 and CCSD(T) complete basis set limit) Comput. 2015, 11, 4972−4991.
interaction energies of small model complexes, DNA base pairs, and (132) Marianski, M.; Supady, A.; Ingram, T.; Schneider, M.; Baldauf,
amino acid pairs. Phys. Chem. Chem. Phys. 2006, 8, 1985−1993. C. Assessing the Accuracy of Across-the-Scale Methods for Predicting
(114) Ř ezác,̌ J.; Riley, K. E.; Hobza, P. S66: A Well-balanced Carbohydrate Conformational Energies for the Examples of Glucose
Database of Benchmark Interaction Energies Relevant to Biomolec- and α-Maltose. J. Chem. Theory Comput. 2016, 12, 6157−6168.
ular Structures. J. Chem. Theory Comput. 2011, 7, 2427. (133) Dohm, S.; Hansen, A.; Steinmetz, M.; Grimme, S.; Checinski,
(115) Gráfová, L.; Pitoňaḱ , M.; Ř ezác,̌ J.; Hobza, P. Comparative M. P. Comprehensive Thermochemical Benchmark Set of Realistic
Study of Selected Wave Function and Density Functional Methods Closed-Shell Metal Organic Reactions. J. Chem. Theory Comput. 2018,
for Noncovalent Interaction Energy Calculations Using the Extended 14, 2596−2608.
S22 Data Set. J. Chem. Theory Comput. 2010, 6, 2365−2376. (134) Grimme, S. Supramolecular binding thermodynamics by
(116) Ř ezác,̌ J.; Riley, K. E.; Hobza, P. Benchmark Calculations of dispersion corrected density functional theory. Chem. - Eur. J. 2012,
Noncovalent Interactions of Halogenated Molecules. J. Chem. Theory 18, 9955−9964.
Comput. 2012, 8, 4285−4292. (135) Bytautas, L.; Ruedenberg, K. Correlation energy extrapolation
(117) Goerigk, L.; Hansen, A.; Bauer, C.; Ehrlich, S.; Najibi, A.; by intrinsic scaling. IV. Accurate binding energies of the homonuclear
Grimme, S. A look at the density functional theory zoo with the diatomic molecules carbon, nitrogen, oxygen, and fluorine. J. Chem.
advanced GMTKN55 database for general main group thermochem- Phys. 2005, 122, 154110.

1670 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671
Journal of Chemical Theory and Computation Article

(136) Maniero, A. M.; Acioli, P. H. Full configuration interaction


pseudopotential determination of the ground-state potential energy
curves of Li2 and LiH. Int. J. Quantum Chem. 2005, 103, 711−717.
(137) Karton, A.; Goerigk, L. Accurate reaction barrier heights of
pericyclic reactions: Surprisingly large deviations for the CBS-QB3
composite method and their consequences in DFT benchmark
studies. J. Comput. Chem. 2015, 36, 622−632.
(138) Goerigk, L.; Sharma, R. The INV24 test set: how well do
quantum-chemical methods describe inversion and racemization
barriers? Can. J. Chem. 2016, 94, 1133−1143.
(139) Karton, A.; O’Reilly, R. J.; Chan, B.; Radom, L. Determination
of Barrier Heights for Proton Exchange in Small Water, Ammonia,
and Hydrogen Fluoride Clusters with G4(MP2)-Type, MPn, and
SCS-MPn Procedures−A Caveat. J. Chem. Theory Comput. 2012, 8,
3128−3136.
(140) Karton, A.; O’Reilly, R. J.; Radom, L. Assessment of
Theoretical Procedures for Calculating Barrier Heights for a Diverse
Set of Water-Catalyzed Proton-Transfer Reactions. J. Phys. Chem. A
2012, 116, 4211−4221.
(141) Bryantsev, V. S.; Diallo, M. S.; van Duin, A. C. T.; Goddard,
W. A. Evaluation of B3LYP, X3LYP, and M06-Class Density
Functionals for Predicting the Binding Energies of Neutral,
Protonated, and Deprotonated Water Clusters. J. Chem. Theory
Comput. 2009, 5, 1016−1026.
(142) Hait, D.; Head-Gordon, M. How Accurate Is Density
Functional Theory at Predicting Dipole Moments? An Assessment
Using a New Database of 200 Benchmark Values. J. Chem. Theory
Comput. 2018, 14, 1969−1981.
(143) xtb standalone code (version 5.8.1). Please contact xtb@
thch.uni-bonn.de for the program.

■ NOTE ADDED AFTER ASAP PUBLICATION


This paper was published on February 11, 2019, with errors in
the Abstract Graphic and the Supporting Information
paragraph. The corrected version was reposted on February
12, 2019.

1671 DOI: 10.1021/acs.jctc.8b01176


J. Chem. Theory Comput. 2019, 15, 1652−1671

You might also like