Tight Density Functional Theory PHD Thesis
Tight Density Functional Theory PHD Thesis
Tight Density Functional Theory PHD Thesis
PhD thesis
Zoltan Bodrog
Universitat Bremen
2012
Promotionskommission
Erstgutachter
Zweitgutachter
Pr
ufer
Pr
ufer
Vertreter der wissenschaftl. Mitarbeiter
Vertreter der Studenten
Thomas Frauenheim
Gotthard Seifert
Thomas Heine
Peter Maa
Contents
0.1
0.2
0.3
0.4
Abstract . . . . . .
Kurzfassung . . . .
Summarium operis
Osszefoglal
as . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3
4
5
6
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
7
7
7
8
10
13
13
16
22
25
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
27
27
27
29
30
34
34
35
35
36
37
41
41
44
55
55
58
63
63
66
67
72
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
75
75
77
77
78
78
78
83
83
84
85
87
Summary
89
Acknowledgements
91
Bibliography
92
0.1
Abstract
0.2
Kurzfassung
0.3
Summarium operis
Methodus nomine Density-Functional Tight-Binding (DFTB) in chemia quantica est approximatio calculationum in theoria functionalis densitatis (DFT)
algorithmo KohnSham factarum, quae fundamenta in integralibus pro geminis
atomorum calculandis et basi minimali utendo habet. Contextum construendi
causa eis, qui non maxime periti in DFT, methodis similibus atque in DFTB
sint, nec non ad nomenclaturam sectionibus sequentibus scriptam introducendam habetur introductio in prima thesi, quae altitudines DFT minime, DFTB
autem maxime tractat. Introductio haec est scripta etiam eis, qui in DFT iam
periti aut DFT tangere non volentes methodum DFTB ipsam celerrime intellegere volunt.
In media thesi tractatur pars energiae in DFTB nomine energia repulsiva
appellata atque investigantur producenda parametra. Producenda parametra
est optimanda pars energiae quasi-classica, ad DFTB specifica, quae summa
potentialium effectivorum pro geminis atomorum, vulgo potentialium repulsivorum, calculatur et ad energiam methodo KohnSham calculatam additur. Haec
potentialia repulsiva omnibus geminis typorum atomorum adiustenda fuerunt
tamen labor immensis vires humanas maxime dissipans. Situatio haec obstitit
novis calculationibus methodo DFTB cum typis atomorum prior non praesentibus. Automaton problema hoc solvens propono et describo, quod in cursu
quaestionis doctoralis materiam huius dissertationis quoque producentis creatum est. Complexae in descriptione habentur etiam partes automati meliorem
faciendi, e.g. implementatio obiectivorum ad adiustanda parametra novorum
energeticalium et perspectiva optimandorum parametrorum electronicorum. In
hoc capite sunt etiam exempla nonnulla parametrisatorem automaticum multum bene functum illustrantia scripta (quorum parametra inter halogenia et
elementa organica optimata brevi tempore publicanda sunt [23]). Hoc caput
contenta maxime ab articulo has res describenti [5] ducit.
Altera solutio problematis parametrorum est calculatio potentialium repulsivorum arte ab initio facta, cuius formulae perfecte derivatae in ultimo capite
sunt scripta. Expectantur repulsiva sic calculata utilia esse in situationibus,
quibus calculationes celeriter currere oportet, errores autem magis tolerandi
sunt. Repulsiva hoc modo calculata sunt etiam publicanda [6].
In parte dissertationis tertia mutationes ad methodum DFTB ipsam et
partem electronicam meliorem fiendam propono. Has mutationes partem SelfConsistent-Charges (onera electrica sibi consistentia) methodi DFTB, implementantem calculationes sibi consistentes iterativas ad modum KohnSham
faciendas, sive praecisius approximationem earum oneribus electricis punctis
formatis utendam, afficiunt. SCC-DFTB possit melior fieri a melioribus interactionibus inter illas fluctuationes onerum electricorum. Ad primum continet
hoc thema expansionem multipolis constructam loco onerum punctis similium.
Ad secundum habentur hic interactionum functiones semiempirice calculatae,
loco interpolationis multum heuristicae. Plurimae harum theseon sunt etiam
in [4] publicatae.
0.4
Osszefoglal
as
Chapter 1
An introduction to the
DFTB method
1.1
1.1.1
| = H|
(1.1)
stationary Schr
odinger equation for the ground state of the system is far from
trivial when is a multi-particle wavefunction with N particles. The spatial
uses
representation of in this case uses 3N spatial coordinates, while that of H
6N , and because of electron-electron interactions, all of these are inseparably
connected in the integro-differential Schrodinger equation.
However, quantum chemistry, as a prevalent, killer application of quantum
mechanics, quickly came out with the most successful simplification scheme of
the above problem, the density-based calculation of electronic configurations.
This means changing the Schrodinger equation or the equivalent
E() = |H|
= min., | = 1
(1.2)
variational problem to another variational problem, but the new problem searches
for the ground-state electronic density, not the ground-state wavefunction.
E() = min.,
(x)dx = N
(1.3)
x
In the presence of a Vext external electric potential, the above energy func1
1
E() = . . . (x1 , . . . , xN )
+
+
Vext (xi )
2
|x
xj |
i
i
i
i=j
Note that the last term in E() is trivially equal to Vext (x)(x)dx, and therefore the above equation is an implicit proof that the existence and validity of
density-based treatment equals to the existence of the F () = F () functional
which comprises kinetic energy as well as all terms of interaction between the N
electrons. This F () functional is called the universal part of energy functional,
since it is the same for every system containing N electrons.
The first attempt at such a scheme was the ThomasFermi method [15, 35]
with its simple kinetic energy functional and total neglection of exchange and
correlational energies (a term approximating exchange energy was later added
by Dirac [9] and the kinetic energy part was improved somewhat by Weizsacker
[37]), that is a very simplistic universal functional. This approximate densitybased method is far from fulfilling the (1.4) equivalence precisely, and serves
barely more than as a historical example today.
The very success of density-based quantum-chemical methods had begun
with the theoretically strict and exact formulation of density-functional theory
based on the HohenbergKohn theorems that establish the validity of handling
electronic density instead of wavefunction as the basic variable of calculation.
Validity is ensured by the existence of a F () universal functional that makes
the (1.4) equivalence exact. However, the actual form of this F () functional is
unknown, and thus it can only be approximated in practical calculations. Based
on the HohenbergKohn theorems, Kohn and Sham founded the most efficient
calculation method up to date that calculates density as that of non-interacting
electron-like quasiparticles.
1.1.2
The first HohenbergKohn theorem establishes that different problems necessarily have different ground state electron densities, or equivalently, no two
different external potentials can lead to the same ground-state electronic density with the same number of electrons. The proof is simple: if both the V1 (x)
and V2 (x) external potentials give (x) as ground-state density and 1 , 2 are
the two ground-state wavefunctions for the two potentials, respectively, then
(1.6)
1.1.3
(1.11)
The invention of Kohn and Sham about calculating the F () universal functional was calculating it with a method that is highly accurate compared to
earlier attempts, but circumvents the direct construction of an F () functional.
Instead, the density is calculated as a
(x) =
occ.
i(x)i(x)
(1.12)
occ.
i(x) i(x)dx.
2
(1.13)
In the KohnSham machinery, the number of used orbitals equals to the number
of electrons, i.e. there are no unoccupied KohnSham orbitals in the original
scheme. Formulae can be upgraded, however, to the non-zero-temperature case
occ.
by changing
to
ni where ni are the appropriate occupation numbers.
i
To provide the collection of i KohnSham orbitals with a Schrodinger-like system of equations to calculate them, one places them into the noninteracting
reference system of the original problem. This reference system has an external
potential composed of the original external potential and the effective potential
of electrons, called the KohnSham potential:
VKS (x) = Vext (x) + VC (x) + Vxc (x)
where VC is the Coulomb potential of electrons
(y)
VC (x) =
dy,
|x y|
(1.14)
(1.15)
occ
KS |i +
0=
i|H
i i|i
i
(1.16)
(1.17)
(1.19)
J() Exc ()
(1.20)
The exchange-correlational part of KohnSham potential is the most challenging part of the whole construction. While all the other parts are quantummechanically motivated and well underpinned, any computationally reasonable
Vxc () exchange-correlational potential is far from exact and physically correct.
Nonetheless, approximations to it proved to be quite precise in quantum chemical calculations, although they are determined with surprisingly simple constructions. The simplest one is the local-density approximation (LDA), which
formulates the exchange-correlational energy as
Exc = ((x))(x)dx,
(1.21)
x
where P is the one-electron density operator of the KohnSham system, and TKS
is the KohnSham kinetic energy operator. Notethat the kineticenergy operator is a constant with respect to P , this is why Tr P P Tr(TKS P ) = Tr(TKS P ).
This Legendre-transformation-like construction, as a thermodynamical analogy,
makes going over from the external potential to the ground-state density as
the basic determinative variable of the problem analogous to changing the N
particle number to the
E
=
(1.24)
N
12
1
(x)V (x)dx (x)Vxc (x)dx + Exc ().
(1.25)
E = EKS
2
x
1.2
1.2.1
(1.26)
where the VKS (x) KohnSham effective potential behind the VKS operator is the
(1.14) sum of the VC (x) Coulomb potential generated by the (x) charge density,
the Vxc (x) functional derivative of exchange-correlation energy with respect to
(x), and the Vext (x) external (non-electronic) potential of the atomic nuclei:
(y)
Exc [(x)]
VKS (x) =
dy +
+ Vext (x).
(1.27)
|x y|
(x)
y
Expressing all the |i KohnSham electronic orbitals in the atomic orbital basis,
|i =
ci, |
(1.28)
where | runs over all atomic orbitals around all atomic centres in the entire
basis. The stationary Schrodinger equation in (1.26) thus numerically becomes
a generalized eigenvalue problem:
H ci, = i
S ci,
(1.29)
KS 1 VC Vxc
+ Exc + Enuc
(1.31)
E = Tr P H
2
= Tr Pvalence VC Vxc
2
nuclei
A=B
1
UZ Z (rAB )
2 A B
(1.33)
(we drop the valence index in the sequel and unindexed electronic densities will
denote those of the valence electrons from here on). The UZA ZB (rAB ) potentials
which depend on the type of atom pair AB and the rAB distance between
them are fitted preliminarily on fit systems against ab initio or experimental
energetics, in the process of parametrization.
Third, the KohnSham electronic energy part of total DFTB energy is also
simplified. When it is expressed with the basis wavefunctions
KS ] =
EKS = Tr[P H
KS |i =
ni i|H
(1.34)
i,,
|VKS | =
| VB |
(1.35)
B
V ( atoms
A )) since exchange-correlational energy expressions are not linear. In the latter
A
case, however, the construction of pairwise potentials in (1.36) will handle the problem.
15
(1.36)
and they are tabulated as functions of atomic distances. These integral tables
are used by DFTB to calculate the KohnSham energy without actually calculating any integrals run-time. Building A + B-type Hamiltonian terms also
handles the problem of density superposition (see footnote 2). In this case,
V (A + B ) is used instead of VA + VB , and for the Hamiltonian matrix
atomselement
in question the difference between |V (A + B )| and |V ( A
A )|
can be approximated with the sum
|C (x)dx
(1.37)
| Vxc ()
d(x)
=A (x)+B (x)
x
C=A,C=B
if [] = [],
0
otherwise.
(1.38)
1.2.2
SCC-DFTB
To handle the energetics of charge fluctuations, the DFTB method was extended
to an iteratively solved self-consistent DFT-like method, called SCC-DFTB [12].
The iteration is done by constructing a new KohnSham Hamiltonian with the
new (x) charge density from the last step instead of the initial (0) (x) and
calculating KohnSham energies, LCAO KohnSham electronic states as well
as a more precise charge density in the next step, and so on. This iteration is
made until the KohnSham states converge. The actual realisation is detailed
below.
We write electronic charge density as a sum (x) = (0) (x) + (x) of
the starting charge density (the mere superposition of pseudoatomic electron
densities without interatomic interactions) and the fluctuations, and likewise
the density operator as P = P (0) + P .
As the charge density is in a very close linear relation with the density op KS [(x)] KohnSham Hamiltonian
erator, we will take the dependence of the H
16
(1.39)
1
E = Tr P T + VC (P ) + Vext
+ Exc (P ) + Ecores + Enuc
(1.41)
2
(where Ecores is the Tr[Pcores . . .] term in the (1.33) definition of repulsive energy), the second derivative of E with respect to P comes only from the Coulomb
and the exchange-correlation energy, the other parts either do not depend on P
or are only linear in it.
2 E
2 EC
2 Exc
=
+
P P
P P
P P
(1.42)
where EC = 21 Tr[P VC ].
We investigate the Coulomb part of the above derivative first. For this, (x),
the electronic charge density can be calculated as
(x) = x| P |x
(1.43)
ni |ii| |x =
ni |i(x)|2 ,
(1.44)
x|
ni |ii| |x
i,,
(1.45)
i,,
where i(x), (x) and (x) are position-space wavefunction values of the i Kohn
Sham orbital and the , basis atomic orbitals, respectively.
17
VC = VC (x) |x x| dx
x
(y)
|x x| dydx
|x y|
x,y
y| P |y
|x x| dydx.
|x y|
(1.46)
x,y
y| P |y x| P |x
1
1
dxdy.
EC = Tr(P VC ) =
2
2
|x y|
(1.47)
x,y
1
2
i,j,,,,
1
(y)(y)(x)dxdy.
(x)
|x y|
(1.48)
x,y
1
2
A,B
i,j,,
{A},{B}
1
(y)(y)(x)dxdy = 1
(x)
IAB
|x y|
2
(1.49)
A,B
x,y
defining the pairwise IAB subsums. With different but similar rearrangements,
we can also define other pairwise subsums: a constraint {A}, {B} in the
summation yields JAB , {A}, {B} yields KAB and {A}, {B}
yields LAB . With them
1
1
1
1
IAB =
JAB =
KAB =
LAB .
(1.50)
EC =
2
2
2
2
A,B
A,B
So
EC =
A,B
A,B
1
{IAB + JAB + KAB + LAB } .
8
(1.51)
A,B
The above summation can be taken apart to an on-site and an off-site part:
(on)
EC = EC
(off)
+ EC
1
{IAA + JAA + KAA + LAA }
8
A
1
+
{IAB + JAB + KAB + LAB } .
8
A=B
18
(1.52)
1
=
8
A=B
i,j,,
{A},{B}
1
S S + . . .
|xA xB |
(1.53)
where xA is the position of the Ath atomic centre and S is the overlap matrix
(the point-like limit of IAB is presented in the equation, the limit of JAB +
KAB + LAB is symbolised by the dots). Defining
qA =
{A},,i
1
ci, ci, S + ci, ci, S )
ni (
2
(1.54)
as the Mulliken charge of atom A, the (1.53) form of off-site Coulomb energy
part in this point-like approximation easily becomes
(off)
EC
1 qA qB
2
rAB
(1.55)
A=B
where
AB = (A , B , rAB )
(1.57)
1
and AB rAB
0 with rAB . The actual shape of AB at short distances
is determined by the A and B parameters [13] that depend only on the types
of atoms A and B, respectively.
ni (
ci, ci, || + ci, ci, ||) .
{A},,i
19
(1.58)
qA =
x| PA |x dx =
A (x)dx
(1.59)
A PA = P and
A A = ). With these partial atomic density operators we
write the on-site part
(on)
EC
1
=
2
A x,y
y| PA |y x| PA |x
dxdy
|x y|
=
1
2
A x,y
A (y)A (x)
dxdy.
|x y|
(1.60)
Thus we see that, just like the off-site part, the on-site part of Coulomb energy
is quadratic in PA or A .
The above on-site part of the Coulomb energy cannot be further simplified
and an approximative formula, unlike the off-site case, cannot be derived. To
circumvent this, one determines its second derivatives with respect to atomic
charges semiempirically. If the energy of a lone atom A with qA electrons is EA ,
then according to Janaks theorem [19], its first derivative is
EA
= A
qA
(1.61)
where A is the energy of the highest occupied atomic orbit of the free atom.
Furthermore,
2 EA
A
=
= UA
(1.62)
2
qA
qA
that is nothing else than the Hubbard parameter or chemical hardness of element
ZA . If we set the parameters in (1.57) so that
AA = UA ,
(1.63)
1
1
1
2
AB qA qB +
UA qA
=
AB qA qB .
2
2
2
A=B
(1.64)
A,B
Note that the above semi-empirical on-site energy derivative is based on the
total energy, not only the Coulomb energy of electrons. This means that not
only the second-order change of Coulomb energy in (1.60) is covered by it, but
both relevant parts of (1.42) are included (Coulomb and xc). In addition to
this, the analytical interpolation [13] that is realized by the s between the
on-site chemical hardness and the long-range off-site Coulomb interaction can
be regarded also as a coarse description of the gradually vanishing second-order
xc energy between two neighbouring atoms.
20
1
AB qA qB .
2
(1.65)
A,B
Then
KS
H
E
P
AB
A,B
qA
P
qB =
AB
A,B
qA
qB
P
(1.66)
where the last step makes matrix elements of the operator derivative explicit.
Since
ni ci, ci, ,
(1.67)
P =
i
1
2
P S + P S
{A},
1
2
(P S + P S )
(1.68)
{A},
(the last equality coming from the self-adjointness of P ). With all this
1
2 S if {A},
1S
if {A},
qA
= 2
S
P
if , {A},
0
otherwise,
(1.69)
and finally
KS
H
1
= S
([]B + []B )qB
2
(1.70)
where [] and [] denote the respective atomic centres of the orbitals. Because
the above Hamiltonian is linear in the Mulliken charges, its change with respect
to the non-interacting atoms is
1
KS
H = H
= S
[]B + []B qB .
(1.71)
2
l
ql
(1.72)
for every atom type (atom type indices were suppressed for brevity, l and l
denote angular momenta of subshells). With the Hubbard matrix, the secondorder on-site energy change for every atom can be calculated as
1
(on)
ql Ull ql
(1.73)
Eone atom =
2
l,l
taking into account that electron fluctuation is not restricted to only one atomic
subshell.
We are aware of only one publicly available DFTB implementation using
different Hubbard parameters for different subshell interactions [2]. But even
there, the off-diagonal Hubbard matrix elements are approximated with means
of diagonal elements instead
Ull
Ull + Ul l
,
2
(1.74)
1.2.3
Spin-spin interaction
In the collinear spin model of spin-dependent DFTB [25], the density matrix of
electrons depends on the z-aligned spin values too. Thus traces in the spatial
representation are summations not only along spatial coordinates, but along
spin indices too. This makes a second-order correction to energy like this:
1
2 E
2 E =
(x, )(y, )dxdy
(1.75)
2
(x, )(y, )
, x,y
22
2 E
2 E
(x)(y) +
(x)(y)
(x)(y)
(x)(y)
x,y
2 E
2 E
(x)(y) +
(x)(y)dxdy
(x)(y)
(x)(y)
through (x,)
and (x,)
.
= 12 (x)
+ (x)
= 12 (x)
(x)
+
(1.76)
E
After realizing that (x)(y)
must be zero due to symmetry reasons3 , the
second-order energy correction stemming from magnetization (spin polarization)
of electrons can be simplified. The second-order energy change thus becomes
1
2 E
2 E
2
E=
(x)(y) +
(x)(y)dxdy (1.77)
2
(x)(y)
(x)(y)
x,y
in which the first term is the spin-independent second-order correction constructed with the energy derivatives in (1.42). Its reformulation within the
DFTB formalism was treated in 1.2.2. The second parts simplification and inclusion into the DFTB total energy expression goes in the same way, but due to
the more localized nature of spin-spin interaction, the sum of interacting spin
populations is strictly restricted to on-site terms. The DFTB spin-spin term
similar to (1.64) is
1
WA p2A
(1.78)
2
A
(1.79)
A,l,l
23
l,
1 l,
.
(1.80)
WA,ll =
2 nl
nl
With the above DFTB spin-spin interaction energy, e.g. magnetization properties of iron clusters can be computed quite accurately with DFTB [21]. The
need of a DFTB code calculating these magnetic properties accurately lead to
a subshellwise calculation of Mulliken populations and the respective almost
complete subshellwise implementation of atomic hardness mentioned in 1.2.2.
The simple collinear spin model can be generalized to non-collinear spins,
it involves extending wavefunctions to two-component non-relativistic spinors,
and substituting (1.79) with
1
WA,ll pA,l pA,l
2
(1.81)
A,l,l
where Mulliken charges first become 2 2 matrices with spinor indices (these
come from the orbital-times-orbital part of Mulliken density operator) and a
decomposition of them to a q + px x + py y + pz z linear combination of Pauli
matrices leads to the Mulliken charge itself along with spin populations in three
spatial dimensions. For an accurate derivation see [24].
Spin-orbit coupling
In the non-collinear treatment of spin effects in the DFTB+ code [2], the spin 2 2 (spinor)
orbit coupling Hamiltonian can be easily written as the L
is the angular momentum operator and is the vector of Pauli
matrix where L
matrices. For a bit deeper insight, see also [24].
Possibly, the non-collinear spin-orbit coupling scheme can also be simplified
to a collinear one by collapsing the interaction part into the z direction. This
leaves
m q p
(1.82)
24
1.2.4
where A and B both run over the atoms in the system and ZA ZB indicates the
type of atom pair AB. Naturally, UZA ZB (r) = UZB ZA (r).
The parametrization process optimizes these UZA ZB (r) pair potentials to
cover the difference between the reference energies of certain fit systems and the
corresponding electronic DFTB energy. The reference energies may be taken
from experimental data or ab initio calculations. We prefer the latter as it
allows a versatile reference data generation. The best parametrization can be
viewed as the one where the set of pair potentials minimizes the error:
2
R=
(Eref EDFTB ) =
A=B
(1.84)
Due to the pair approximation of Erep and the otherwise approximative
nature of DFTB, parametrizations lack universal transferability, but as the cases
of successful parametrizations show, the validity of a good parameter set can
extend to a wide range of problems.
Hand-made repulsive potentials
In its usual course, parametrization for a bond type (e.g. the carbon-carbon
bond) begins with stretching one bond of that kind in an appropriate molecule,
as the simplest case, and creating a UCC (r) curve based on the energy difference
between the DFT reference and DFTB:
UCC (r) = EDFT (r) EKS (r) + const
(1.85)
with r being the length of the stretched bond. The constant term covers the
limit of EDFT (r) EKS (r) at r (this limit contains e. g. the repulsive
contributions from non-varying bonds, that may not be known in detail at all) in
order to ensure a zero limit for UCC (r), r . Of course, one chooses stretched
molecules so that stretching affects only one bond (or maybe several bonds, but
in a totally equivalent way), all the other pairs of atoms with changing distances
remain outside the ranges of their respective repulsives.
One can construct a reasonable curve for a given atom pair type by merging
curve sections created for different molecules which represent different chemical
25
26
Chapter 2
Improvements to handling
repulsive energy
2.1
2.1.1
ZZ , fZZ , (rAB ),
(2.1)
where ZZ is the type of atomic pair AB. Substituting this into the pair potential structure of Erep from equation (1.83), the total repulsive energy for a
given system becomes a linear combination
Erep =
ZZ , XZZ ,
Z,Z ,
27
(2.2)
1
2
fZZ , (rAB ).
(2.3)
ZA ZB =ZZ
A=B
The sum runs over all possible atom pairs where the pair AB belongs to pair
type ZZ .
Using the above, the best ZZ , coefficients may be easily approximated by
a least-squares fit to energy values of several different distortions of a chemical
system as a function of the changing X values. Due to the linearity of the energy
as a function of XZZ , s, this fitting is a multidimensional linear regression.
Running over a sequence of distortions denoted by s of the same system (we
will call this sequence a fit path and the distortions fit steps), the least-squares
fit minimizes the overall error
R=
all
steps
2
(s)
(s)
(s)
Erep
(Eref EKS )
all
steps
(s)
ZZ , XZZ ,
(s)
(Eref
(s)
EKS )
(2.4)
Z,Z ,
where EKS is the KohnSham energy of DFTB, that is the KohnSham energy
of valence electrons. With expression (2.2) of the total repulsive energy, the
stationary condition
R[ZZ , ] = 0
(2.5)
ZZ ,
of the above error leads to a matrix expression of the coefficients ZZ , :
1
A = XXT
XE.
(2.6)
The matrices E, X and A are constructed from the above energies, X structural
quantities and s in the following way:
(1)
(1)
Eref EKS
(2)
(2)
E EKS
E=
ref
..
.
(2)
XCH,1
. . . . . . . . . . .
(1)
(2.7a)
(2)
X
HH,1
(1)
XHH,2
..
.
X=
X (1)
CH,1
(1)
XCH,2
..
.
XHH,1
(2)
XHH,2
28
(2.7b)
HH,1
HH,2
..
.
A=
CH,1
CH,2
..
.
(2.7c)
2.1.2
R=
all
paths
p
2
(ps)
(ps)
(ps)
Erep
(Eref EKS )
(2.8)
sp
with p enumerating the paths and Erep written as a function of s and Xes
within each path in the same way as the one-path case. It should be noted
here that the number of ZZ , parameters does not depend on the number of
fit paths (nor on the number of steps in the individual fit paths). Its value is
determined only by the choice of the basis functions to describe the repulsive
interactions in question.
The E column vector of target repulsive energies for the multiple-path fitting
1 Displacements of a carbon atom in a carbohydrogen molecule alter the CC and CH
bonds (but not the HH bonds), and thus enables fitting to these repulsive potentials only.
29
E
..
E(2)
E=
= (21)
(21)
..
EKS
Eref
.
E (22) E (22)
ref
KS
..
.
(2.9a)
X=
X(1)
X(2)
..
.
(2.9b)
This all gives back matrix equation (2.6) on A for the multiple-path fit.
2.1.3
ZA ZB , f ZA ZB , (rAB )
B=A,
30
rAB
u.
rAB
(2.10)
(i,u)
FA,u =
ZZ , XZZ ,
(2.11)
ZZ ,
rAB
(A,u)
XZZ , =
f ZZ , (rAB )
u
rAB
(2.12)
ZA ZB =ZZ
B=A
(s)
(s+1)
(s)
(s+1)
(s)
Erep
= Erep
Erep
=
ZZ , XZZ , XZZ ,
(2.13)
ZZ ,
HA,uv = uHv =
um
B=A,mn
2 fZA ZB , (rAB )
vn
xA,m xA,n
B=A,mn,
1 f
1 f
1 2 f
(u r)(v r) +
(u v).
ZA ZB ,
2
2
3
r r
r r
r r
ZZ ZA ZB =ZZ
B=A,
2 UZA ZB (rAB )
vn
xA,m xA,n
ZA ZB , um
(2.14)
(with rAB = (xB,1 xA,1 , xB,2 xA,2 , xB,3 xA,3 ) at the beginning and f =
fZA ZB , (rAB ), r = rAB , r = rAB in the last step). So, with
(A,uv)
XZZ ,
ZA ZB =ZZ
B=A
1 f
1 2 f
3
2
2
r r
r r
(u r)(v r) +
1 f
(u v), (2.15)
r r
(A,uv)
HA,uv =
ZZ , XZZ , .
(2.16)
ZZ ,
With the E vector composed of (Href HKS )A,uv s and the X matrix composed
of the above Xes, the fitting of the Hessian can be included as an additional
path into the fitting scheme.
When u and v are on the Ath and B th atoms, respectively,
HAB,uv = uHv = um
2 UZA ZB (rAB )
2 UZA ZB (rAB )
vn = um
vn .
xA,m xB,n
xA,m xA,n
(2.17)
Therefore a similar construction applies to off-site Hessian parts, but with the
opposite sign and without the summation over B.
As a linear combination of the above u and v atomic virtual displacements,
every collective distortion of a molecule can be constructed. This knowledge
can be used to fit to Hessians of DFT reference algorithms that give no detailed
Hessian matrix but only vibrational modes and frequencies in their output. If
e is a (3N -component) collective eigenmode of the molecular Hessian with
frequency,
2 Me = He = (HKS + Hrep )e
(2.18)
where M is the diagonal mass matrix. The vector of equations contained in
2 Me HKS e = Hrep e
32
(2.19)
(s+1)
(s)
(s+1;i,u)
(s;i,u)
Frep,A,u Frep,A,u =
ZZ , XZZ ,
XZZ ,
(2.20)
ZZ ,
rep
1 E
1
=
V mn
V
=
FAB,m rAB,n
Aa cell
B
1
V
Aa cell
B
1
V
1 U (rAB )
rAB,m rAB,n
rAB rAB
ZA ZB , f (rAB )
Aa cell
B
rAB,m rAB,n
rAB
(2.21)
(uv)
uv =
mn um vn =
ZZ , XZZ ,
(2.22)
ZZ ,
mn
2 An important issue is whether we must compare DFT-equilibrium Hessians to DFTBequilibrium Hessians or we must compare Hessians of the very same geometry (practically
the DFT equilibrium geometry, as DFT calculators tend to compute (real) eigenvalues and
eigenmodes instead of outputting raw Hessian matrices). From a theoretical point of view,
the latter comparison is more valid while from a semiempirical (practical) point of view, the
former comparison is much more justified.
33
XZZ , =
f (rAB )
ZA ZB =ZZ
Aa cell
B
(rAB u)(rAB v)
.
rAB
(2.23)
So with the above Xes, uv can be another valid target of our repulsive fitting
algorithm.
2.1.4
In the formalism described so far, every fit step contributes to the R overall
error with the same weight. As this may not always be the desired behaviour,
we allow each step in each path to have an individual weight for its contribution
to the total error. If the fit is done for multiple physical properties (e.g. energies
and forces) each property can also be weighted differently.
The weighting issues come to play mainly in two areas. First, one typically
would overweight near-equilibrium geometries to assure a higher precision at
near-equilibrium bond lengths at the cost of less precise description for strongly
distorted geometries. Furthermore, weighting becomes a key issue when multiple physical properties are invoked into the fit, since the numerical values of the
differences in the various properties (energy, force, etc.) must be converted to
the same scale. This requires some experimenting but it offers the possibility of
balancing the performance of repulsives for various physical quantities. For example, heavy weights for forces are usually necessary when the fitted repulsives
give poor results with geometry optimization otherwise.
2.1.5
We have experimented with several different f (r) basis functions for the repulsive fitting. The splines used in most of the current DFTB implementations
turned out to be inappropriate for a fitting procedure as they tend to give
very oscillatory behaviour. As a straightforward alternative, we decided for the
(2.24) cut-off polynomials first. They were used in the earliest DFTB implementation [31] and still retain popularity with making parametrization by hand,
since most of the parametrizations up to now have been able to be done with
them. The zeroth and first-degree terms are omitted from such a polynomial to
ensure a smooth decay at its cutoff distance r0 :
(r r0 )
if r < r0 and 2,
f (r) =
(2.24)
0
otherwise.
This representation showed to be successful at the relatively easy hydrocarbon
parametrization, among other examples.
To emulate spline-like behaviour in our scheme, we also tested bases containing the above cut-off polynomials, but having no universal cutoff value (these
bases can be regarded as sums of multiple single-cutoff bases). Bases with two
cutoff values are very efficient at improving polynomial repulsives, while more
cutoff values bring up the oscillatory nature of splines. Less successful but
still noteworthy examples of spline-like bases are wavelet bases, which we also
34
2.1.6
Besides the automatized fitting process itself, there are three subprocesses of the
parametrization workflow in which our program substantially lowers the human
contribution.
The path building methods mentioned so far and some others are implemented to be executed automatically. They include bond stretchings,
displacing atoms, uniform volume changes, linear interpolations between
two configurations, and using predefined paths (e.g. MD trajectories or
reaction paths).
Instead of using fixed sets of metaparameters (input parameters determining the parametrization itself) for the fitting process, batch fits can
use intervals of them, e.g. instead of setting the longest chlorine-oxygen
polynomial cutoff to a fixed value of 2.5
angstroms, we can make a probethrough from 2.3 up to 2.8 by steps of 0.1
A. Scanning over all of these
values in all of these intervals in every combination spares a lot of try-andfail cycles for the user. At the end of the batch run, the set with lowest
total error on the targets (as defined in (2.8)) is picked as the fittest solution.
A module for defining test systems is built into our program too. It
tests the energetical and geometrical performance of fitted repulsives on
the specified test systems. This way it can give a first-glance feedback
about the performance of the fitted repulsive set on systems that were not
necessarily fit systems.
2.1.7
Batch fitting of parameters outlined in the previous section is not only able
to run over large sets of metaparameters, but it is also capable of brute-force
choosing from a set of different electronic parameters. This means changing
electronic parameters (the pseudoatomic compression radii) defined in 1.2.1 step
by step, just like metaparameters of the fitting process, and then picking the
one that allowed the best fit. The best fit means the same as in the previous
section: the smallest accumulated error of repulsive targets defined in (2.8).
35
2.2
Since the molecular reference ab initio calculations in the hand-made sets were
mostly done using the Gaussian [17] code, we also used it as a reference for
molecular systems. For the periodic systems, however, we have found the Siesta
[34] code far more stable (less prone to convergence failures) in our automatic
fitting environment, where distorted systems far from the equilibrium must be
calculated very often. Apart from stability issues, this choice is also a good
cross-calculator and cross-methodology (even between different xc functionals
in DFT references) consistency check of our algorithm and in general for the
DFTB parametrization philosophy. As it will be seen from the results, this
mixing of DFT references did not pose any problem.
The DFT calculations of the last application example, the half halogenorganic set were not made with the above ad hoc machinery, but with the
NWChem [36] code, the B3LYP exchange-correlation functional and a cc-pVTZ
basis. This setting was in line with the other part of the halogen-organic project,
carried by other authors.
The DFTB calculations were carried out using the DFTB+ package [2].
36
2.2.1
Carbohydrogens
37
reference
mio
hom
inhom
0
1.093
7.3
1.089
-2.5
1.094
-0.1 (52.9)
1.080
0
1.531
1.096
17.7
1.501
1.098
-1.0
1.535
1.102
0.1 (94.8)
1.516
1.088
0
1.331
1.087
14.6
1.327
1.094
-2.4
1.327
1.099
-3.6 (68.6)
1.326
1.084
0
1.205
1.067
21.7
1.203
1.075
10.2
1.200
1.080
-3.5 (53.1)
1.204
1.066
0
1.397
1.087
52.9
1.396
1.098
-1.7
1.405
1.104
0.8 (170.8)
1.397
1.090
38
butane
E
(1, 2) CC
(2, 3) CC
(1) CH
isobutane
E
CC
CH
diamond
CC
cyclobutane
E
CC
CH
isobutene
E
CC
C=C
CH (in CH3 )
CH (in CH2 )
bicyclobutane
E
CC (edge)
CC (middle)
CH (in CH2 )
CH (in CH)
cyclobutene
E
CC
C=C
CH (in CH2 )
CH (in CH)
cyclopropane
E
CC
CH
propane
E
CC
CH (end)
CH (middle)
cyclopropene
E
CC
C=C
CH (opposite to C=C)
CH (neighbour to C=C)
0
1.547
1.536
1.097
38.7
1.519
1.518
1.097
2.8
1.555
1.552
1.102
1.6 (179.8)
1.537
1.534
1.088
0
1.535
1.097
38.0
1.518
1.098
1.9
1.552
1.102
1.5 (178.7)
1.534
1.088
1.555
1.540
1.575
1.558
0
1.557
1.095
40.0
1.539
1.102
7.3
1.569
1.107
13.4 (169.2)
1.534
1.094
0
1.509
1.337
1.099
1.087
36.2
1.493
1.341
1.100
1.093
0.5
1.524
1.34
1.104
1.099
-2.7 (153.0)
1.505
1.339
1.090
1.084
0
1.510
1.900
1.112
1.095
26.9
1.464
2.003
1.195
1.066
-2.1
1.549
2.112
1.161
1.021
10.3 (143.4)
1.486
1.980
1.158
1.065
0
1.573
1.519
1.097
1.087
29.5
1.569
1.524
1.104
1.097
-1.8
1.597
1.548
1.109
1.103
7.4 (140.6)
1.538
1.493
1.097
1.089
0
1.509
1.087
18.9
1.489
1.096
-9.6
1.523
1.100
-0.2 (113.8)
1.502
1.087
0
1.532
1.097
1.099
27.7
1.509
1.098
1.107
0.2
1.544
1.102
1.110
0.0 (136.6)
1.525
1.088
1.097
0
1.508
1.295
1.095
1.080
12.2
1.495
1.319
1.107
1.090
-13.8
1.528
1.319
1.109
1.095
-7.7 (83.7)
1.508
1.318
1.096
1.081
39
spiropentane
E
CC (radial)
CC (outer)
CH
methylene-cyclopropane
E
C=C
CC (radial)
CC (outer)
CH (in CH2 )
CH (on ring)
propadiene
E
C=C
CH
1,3-butadiene
E
CC
C=C
CH (middle)
CH (end)
2-butyne
E
CC
CC
CH
propyne
E
CC
CC
CH (in CH3 )
CH (in CH)
propene
E
CC
C=C
CH (in CH3 )
CH (in CH)
CH (in CH2 )
0
1.485
1.530
1.088
29.4
1.479
1.508
1.097
-18.9
1.508
1.547
1.102
-2.0 (173)
1.488
1.524
1.088
0
1.322
1.470
1.540
1.088
1.089
25.9
1.328
1.465
1.512
1.095
1.098
-10.6
1.327
1.491
1.551
1.101
1.102
-4.5 (128.7)
1.327
1.472
1.529
1.086
1.089
0
1.307
1.088
21.2
1.312
1.096
-2.6
1.312
1.102
-7.8 (83.6)
1.312
1.087
0
1.439
1.392
1.089
1.086
49.9
1.436
1.372
1.098
1.104
16.2
1.457
1.373
1.103
1.085
10.0 (143.2)
1.441
1.370
1.089
1.095
0
1.462
1.209
1.097
38.8
1.455
1.209
1.100
6.5
1.477
1.205
1.105
-1.2 (132.0)
1.461
1.209
1.091
0
1.460
1.207
1.097
1.066
30.3
1.453
1.206
1.100
1.074
8.3
1.475
1.203
1.104
1.079
1.2 (92.6)
1.459
1.207
1.090
1.066
0
1.502
1.333
1.098
1.091
1.087
24.9
1.485
1.334
1.100
1.102
1.093
-1.5
1.517
1.334
1.105
1.106
1.098
-3.7 (110.3)
1.497
1.333
1.091
1.092
1.084
1
Erep ({Rnuclei }) =
UZA ZB (rAB ) +
UZA
(2.25)
2
A=B
40
2.2.2
Zinc-oxides
2.2.3
Titanium-organics
Our next test of the fitting automaton was producing a titanium-oxygen set and
extending it to a titanium-organic set. For this parametrization a good-quality
hand-made set (tiorg) has been recently created by Dolgonos et al. [10]. We
41
Table 2.2: Reference data and its comparison with previous hand-made
parametrization [27] (znorg) and the automatically created one (auto) for
Zn and ZnO crystals. The values given here refer to the equilibrium structure
of each system. E denotes the atomization energy difference with respect to
the reference in kcal/mol. Atom pairs denote distances in
A.
property
reference znorg auto
Zn hcp
E (per Zn2 )
0
115.5
94.1
ZnZn (#1)
2.523
2.796 2.433
ZnZn (#2)
2.886
2.864 2.931
ZnZn (#3)
3.831
4.051 3.788
ZnZn (#4)
4.591
4.872 4.524
ZnO zincblende
E (per ZnO)
0
22.3
-1.1
ZnO
2.005
2.015 2.011
ZnZn
3.274
3.290 3.281
ZnO wurtzite
E (per Zn2 O2 )
0
46.5
-0.7
ZnO
2.017
2.015 2.018
ZnO
2.037
2.014 2.004
used the same reference structures and ab initio reference data (various molecular systems calculated with the B3LYP functional and with mixed SDD+ basis
set) augmented with crystalline reference systems. For the reference calculations
of the periodic systems, the PBE functional, double- plus polarized basis functions and norm-conserving TroullierMartins pseudopotentials had been used.
K-point sampling was set to an 8 8 8 MonkhorstPack scheme with both
Siesta and DFTB in this fit session.
In order to fit repulsive potentials for the Ti-Ti and Ti-O interactions, we
used a fit set including a titanium dimer (with a very low weight), a TiO2
molecule, a planar Ti2 O2 molecule, a tribridged Ti2 O4 , the bulk hcp titanium,
and the bulk anatase and rutile forms of TiO2 . The molecular fit paths were
created by stretching bonds and displacing titanium atoms while the crystalline
paths were created by uniformly changing the volume of the crystal lattices and
by using crystals with displaced titanium atoms. We used both energy and force
targets (generally weighted 1:2) in the fit.
In a fit session of a few days, we were able to produce a set of Ti-Ti and
Ti-O repulsive potentials which reproduce energy and geometrical data in the
same quality as the reference hand-made set. A detailed comparison is given in
2
Table 2.3. These results were obtained using ea1 r and ea2 r -type exponential
functions as basis functions for the fit because this analytical basis gave very
good results quickly with the Ti and Ti-O chemistry.
After creating the repulsives for the TiTi and TiO interactions, we extended the set to a complete Ti-organic set, still using exponential basis functions for expressing the repulsive potentials. The extension turned out to be
more difficult than expected, mainly due to the sudden cutoff in the hand-made
Ti-Ti repulsive giving a very stable 3
A TiTi distance in hcp titanium, titanium nitride and titanium carbide. This feature (shown in Figure 2.1) can be
42
Table 2.3: Titanium-oxygen compound reference data and its comparison with
previous hand-made parametrization [10] (tiorg) and the automatically created one (auto). The values given here refer to the equilibrium structure of
the various systems. E means atomization energy difference with respect to
the reference in kcal/mol. Atom pairs denote neighbour distances in
A, triples
denote angles in degrees (double distance values show artificially broken symmetries of DFTB-optimized lattices). Italicized system names denote fit systems.
property
reference tiorg auto
TiO
E
0
55.0
40.2
TiO
1.586
1.592 1.586
Ti2 O2 planar
E
0
87.7
69.4
TiTi
2.198
2.355 2.092
TiO
1.857
1.891 1.866
Ti2 O2 non-planar
E
0
68.0
57.6
TiTi
2.127
2.249 2.133
TiO
1.838
1.888 1.826
Ti2 O4 #1 (dibridged with end O atoms in cis position)
E
0
49.4
81.3
TiTi
2.716
2.800 2.635
bridging TiO
1.848
1.887 1.812
end TiO
1.622
1.606 1.589
OTiTi (ending O)
126.1
124.7 123.7
Ti2 O4 #2 (dibridged with end O atoms in trans position)
E
0
145.7 82.8
TiTi
2.709
2.726 2.709
bridging TiO
1.840
1.831 1.806
end TiO
1.625
1.608 1.590
OTiTi (ending O)
123.7
122.3 122.2
Ti2 O4 #3 (tribridged with an O atom at one end)
E
0
123.3 66.4
TiTi
2.394
2.540 2.399
bridging TiO (opposite to end O)
1.763
1.801 1.742
end TiO
1.628
1.606 1.586
Ti hcp
E (per Ti)
0
-40.6
5.4
TiTi
2.900
2.993 2.915
TiO2 anatase
E (per TiO2 )
0
22.5
-4.5
shortest TiTi
3.028
2.996 3.082
1.921
1.957
shortest TiO
1.933
1.995
1.958
TiO2 rutile
E (per TiO2 )
0
19.4
1.1
shortest TiTi
3.559
3.613 3.605
1.914
1.976
shortest TiO
1.974
1.992
1.995
43
Figure 2.1: Comparison of the tiorg (dashed) and the automatically generated
(solid line) Ti-Ti repulsives in the area of the sharp cutoff of the former.
0.2
0.1
2.5
3.5
distance []
hardly reconstructed with analytical sets. Although this peculiar shape is the
numerically most convenient way to confine the range of the Ti-Ti repulsive well
below the second-neighbour distance as well as it gives good results for various
systems, it may be an interesting question for future investigations whether it
is a precise representation of the underlying physics.
Similar to the case of the titanium-oxygen fit, we used the same molecules
(TiH4 , Ti(CH3 )4 , Ti(NH2 )4 ) as during the hand-made parametrization [10] extended by crystalline fit systems (TiN and TiC). The fit paths were similar
constructions to the titanium-oxygen case: every molecule had two fit paths
with the relevant bonds stretched and the titanium atom displaced, while the
crystalline systems had a volume change path and a path with a titanium atom
displaced. We handled the relative difficulty of Ti-C and Ti-N fitting compared
to Ti-Ti and Ti-O by lowering the relative weights of the energy targets in the
Ti-C and Ti-N case. As it can be seen from the results (Table 2.4), this resulted
in fairly good geometries at the expense of less accuracy in energies (one-body
terms as a tool for resolving the conflict between energy and geometry accuracy
were not used here).
2.2.4
Halogen-organics
44
Table 2.4: Reference data and its comparison with previous hand-made
parametrization [10] (tiorg) and the automatically created one (auto) for
various titanium compounds. The values given here refer to the equilibrium
structure of each system. E denotes the atomization energy difference with
respect to the reference in kcal/mol. Eb indicates the binding energy between the central Ti atom and the ligands compared to the reference value in
kcal/mol. Atom pairs denote neighbour distances in
A, triples denote angles in
degree. Italicized system names refer to fit systems.
property reference tiorg auto
Ti(CH3 )4
Eb
0
64.6 180.2
TiC
2.072
2.096 2.025
Ti(CH3 )2
Eb
0
-38.8
93.4
TiC
2.038
2.096 2.025
CTiC
113.7
110.2 109.9
crystalline TiC
E
0
111
91.7
TiC
2.141
2.159 2.170
TiTi
3.024
3.047 3.067
Ti(NH2 )4
Eb
0
30.6 287.4
TiN
1.899
1.902 1.853
H3 Ti(NH2 )
Eb
0
12.0
76.2
TiN
1.846
1.898 1.837
HN=Ti=NH
Eb
0
15.5 156.1
TiN
1.707
1.703 1.671
NTiN
114.8
114.7 113.7
crystalline TiN
E
0
196.6 192.2
TiN
2.094
2.159 2.115
TiTi
2.958
3.043 2.982
Ti2 H2 (dibridged planar)
E
0
123.5
131
TiTi
1.985
1.967 2.011
TiH
1.868
1.827 1.899
TiHTi
64.2
65.2
63.9
45
one-body
natural
0.0525
1.8266
2.1182
0.8363
0.6911
0.5817
rep [eV]
reduced
0.0105
0.3653
0.4236
0.1673
0.1382
0.1163
The cutoff values of polynomials building the repulsives were partly allowed
to vary during fitting. The optimized values of these metaparameters are shown
in Table 2.6. The quality of parameters is illustrated by tables 2.72.16.
5 It would be worth investigating that fitting to a B3LYP reference is much more difficult
than fitting to PBE or PBE0 (with which these fits were made more successfully, but they were
rejected for a compatibility reason with mios B3LYP reference). The most straightforward
explanation is that DFTB itself uses integral tables produced with PBE, but there may or may
not be other reasons in the background too. Cf. this with the similar note at the beginning
of 2.2.1.
46
Table 2.6: Cutoff values in the repulsives (note that close multiple values come
from the parts of blended parameter sets, while more distant values come from
multiple-cutoff fits).
pair
ClH
ClO
ClN
ClCl
BrH
BrO
BrN
BrBr
IH
IO
IN
II
cutoff(s) [
A]
1.9, 2.0
1.5, 1.95, 2.8
2.8
2.8
1.9, 2.1
1.8, 2.1, 3.0
3.0
3.0
2.4
1.9, 2.0, 2.3, 3.2
3.1
3.4
47
48
49
50
Table 2.10: Results with molecules containing chlorine and nitrogen. E: atomization energy in eV (E : calculated with one-body terms on the DFTB
side), doubles are bond lengths (distinguished with bond quality if needed) in
Table 2.12: Results with molecules containing iodine and nitrogen. E: atomization energy in eV (E : calculated with one-body terms on the DFTB
side), doubles are bond lengths (distinguished with bond quality if needed) in
NCl3
HClO4
molecule
Cl2
HCl
HClO
ref.
539.8
2945.5
3774.4
1267.8
740.8
3710.9
1297.2
1211.6
1184.5
1003.9
692.8
550.9
544.4
520.6
395.5
387.6
169.1
614.0
614.0
547.5
349.6
252.6
252.6
DFTB
547.6
2929.6
3665.3
1364.6
1038.4
3649.5
1350.0
1334.3
1175.2
1107.4
771.2
526.0
524.6
496.9
394.7
363.2
96.2
757.1
744.3
633.7
383.6
289.2
280.6
NBr3
HBrO4
molecule
Br2
HBr
HBrO
ref.
469.4 (325.3)
2330.0 (2649.0)
3812.7
1614.4
743.5
3724.8
1241.8
1030.4
1022.2
895.0
678.1
488.8
471.8
404.2
305.0
302.9
212.4
516.0
516.0
418.0
139.1
DFTB
298.3
2615.6
3690.7
1282.1
898.8
3683.3
1092.9
1045.3
1033.4
964.4
708.4
366.8
361.5
343.9
292.3
272.5
101.7
667.8
652.8
493.3
240.7
170.9
166.4
NI3
HIO4
molecule
I2
HI
HIO
ref.
218.1
2314.5
3803.2
1135.7
591.6
3732.9
1039.8
886.1
881.0
825.6
607.2
288.2
287.9
270.1
240.3
229.7
511.3
511.3
360.2
156.2
104.7
104.7
DFTB
202.4
2287.6
3715.5
1203.7
774.7
3737.5
1041.0
930.7
917.2
873.4
633.3
285.5
275.3
253.0
217.5
198.6
114.9
620.0
613.7
395.0
179.3
122.3
118.6
Table 2.16: Vibrational modes of molecules calculated with the reference method and the new parameters (frequencies in cm1 ). Data in
parentheses are experimental values from [1].
54
2.3
2.3.1
The representation of Erep in (1.33) as a sum of quasiclassical multibody potentials is not only a matter of convenience. Its validity is strictly underpinned
by the fact that Erep does not depend on molecular charge fluctuations in the
leading order. This means that Erep is a functional of the superposed (0) electronic density of atoms (or the corresponding P (0) density operator) up to a
satisfactory precision, therefore the repulsive energy of a certain chemical system is influenced by only its geometry (i.e. how (0) is composed from atomic
densities), not its particular electronic state. This enables using transferable
quasiclassical repulsive potentials.
Erep (P ) in general. Generally, the E total energy of a chemical system can
be approximated with its Taylor expansion by P , which we investigate here up
to the second order:
2 E
1
E
(0)
P .
(2.26)
+ Tr P
E = E + Tr P
P 0
2
P2 0
The KohnSham energy of the same system6 in DFT is then given by
E
EKS = Tr P
.
P
(2.27)
In the first-order machinery of non-SCC DFTB, the above KohnSham energy is approximated by
(1)
EKS
E
(0)
(0)
(1)
= Tr[P HKS ] = Tr (P + P)
P 0
E
(0) E
= Tr P
+ Tr P
,
P 0
P 0
(2.28)
(1)
and this is supplemented by the repulsive energy to give E = EKS +Erep . Finally
(1)
Erep
=E
(0)
1
2 E
(0) E
+
Tr
P
Tr P
P
.
P 0
2
P2 0
(2.29)
55
2 E
E
(0)
P
= Tr (P + P)
+
P 0
P2
0
E
(0) E
+ Tr P
= Tr P
P 0
P 0
2
2 E
(0) E
+ Tr P
P
+
Tr
P
P
(2.30)
P2 0
P2 0
by the (1.34) definition of the KohnSham energy. The repulsive energy in this
case would be
(2)
(2)
Erep
= E EKS
1
2 E
(0)
(0) E
(0) E
= E Tr P
Tr
P
P
P
,
Tr
P
P 0
P2 0
2
P2 0
(2.31)
E
2 E
HKS =
+
P
(2.32)
P 0
P2 0
KohnSham Hamiltonian to calculate electronic states, but total energy is calculated as its first-order DFTB approximation plus the second-order term in its
Taylor series:
1
2 E
(1)
(2)
E = EKS + Erep + Tr P
P .
(2.33)
2
P2 0
This construction propagates the usage of charge-fluctuation-independent firstorder repulsive potentials into SCC-DFTB. Furthermore, if we take a closer look
at what Erep in SCC-DFTB is
1
2 E
(1)
(2)
(0) E
Erep = E EKS Tr P
P = Tr P
,
(2.34)
2
P2 0
P 0
we notice that the repulsive potential of SCC-DFTB is independent of P up
to the second order.
In the following, we calculate the repulsive energy part in detail when the
total energy is of the well-known structure: E = EC + Exc , the sum of Coulomb
and xc energy (we omit kinetic energy
it is x|P |xx|A|xdx
x(x)x|A|xdx.
56
((x)VC (x))dx
1
1
(0)
(0) (x)VC (x)dx +
(0) (x)VC (x)dx
=
2
2
1
1
(0)
+
(x)VC (x)dx +
(x)VC (x)dx,
2
2
(2.35)
if we expand charge density in the first order as = (0) + . Also using that
the Coulomb potential itself is a linear function of charge density (i.e. the total
Coulomb energy is quadratic in ), we get that
(0)
(0)
(x)VC (x)dx = (x)VC (x)dx
(2.36)
and therefore
1
1
(0)
(0)
(0)
(x)VC (x)dx + (x)VC (x)dx +
(x)VC (x)dx.
EC =
2
2
(2.37)
(1)
Then the Coulomb part of EKS is
EC
(1)
(0)
EKS,C = (x)
dx
=
((0) (x) + (x))VC (x)dx.
(2.38)
(x) 0
It implies a repulsive energy of the form
(1)
(1)
Erep,C = EC EKS,C
1
1
(0)
(0) (x)VC (x)dx +
(x)
(y)dxdy,
=
2
|x y|
(2.39)
Exc =
(x)((x))dx
x
=
x
d
d
1 d2
(0)
dx,
d 0
d 0
2 d2 0
(2.40)
d
=
. In a semilocal
where (0) = ((0) ). In the above mentioned LDA case, d
d
xc
GGA case of xc functional, the part denoted by d (a part of the E
operator
P
derivative) is a differential operator in this integral representation.
57
Exc
(0)
= (x)
dx = ((0) (x) + (x))Vxc
(x)dx
(x) 0
Exc
Exc
(0)
= (x)
dx + (x)
dx
(x) 0
(x) 0
(0)
(0)
(0) d
(0)
(0) d
=
+
+ +
(x) dx.
d 0
d 0
(2.41)
(0)
1
(1)
(1)
(0) d
dx,
Erep,xc = Exc EKS,xc = Wxc
2
d 0
(2.42)
where Wxc is the second derivative of Exc with respect to . It may be written
2
d
(0) d
Wxc = 2
+
,
(2.43)
d 0
d2 0
where differentials of are partial derivatives in LDA, and differential operators
in GGA, just like above. Again here, the investigated non-SCC DFTB repulsive
energy part depends on in the second order only, and even the second-order
term is dropped from repulsive energy in SCC-DFTB.
2.3.2
58
Figure 2.2: The Cl-Cl repulsive potential fitted without any overbinding strategy
and the one with one-body energies (the upper and lower strong dashed lines,
respectively). Long-range Eref EKS differences (dashed thin line originally,
solid line shifted to zero dissociation) and the pairwise overbinding potential
(dotted) that is equivalent to the applied one-body terms for this dimer are also
shown. (Note, however, that one-body energies and pairwise overbinding potentials cannot be equivalent for several distinct stoichiometries; an equivalence
is valid only for one stoichiometry, e.g. Cl2 in this case. Note also the difference between the lower, dissociation-adjusted fitted repulsive and the reference,
resulting from the absence of middle-range repulsives.)
E [a.u.]
0.2
0.1
-0.1
r []
and long-range energies (i.e. a proper energy fit) but having a slope that bridges
this gap over a much longer range (a proper force fit) is an unsolvable conflict.
As it is also evident from Figure 2.2, covering repulsive energy with short-range
potentials is simply not enough. This is the problem that will be studied briefly
in this section.
Taking a closer look at Figure 2.2, one can see that discrepancy on this dimer
system with respect to the short-range pair potential picture is composed of two
main parts. First, the repulsive energy dies out in a much longer range than we
implicitly expect with our short-range potentials. Second, there is a remanent
discrepancy of repulsive energy in the dissociation limit (r ). While the
former difference in pair potential ranges is rather easy to understand (but not
to solve!), one may be shocked about the fact that dissociated molecules are
not converging energetically to sets composed of lone model atoms in DFTB
and the reference method, which means that for a molecule of stoichiometry
Zl Zm
. . . Zn ,
E (dissoc) (Zl Zm
. . . Zn ) = lE(Z) + mE(Z ) + nE(Z )
(2.44)
Erep =
atoms
UZA
atoms
1
UZA ZB (rAB ).
+
2
(2.45)
A=B
Of the two problems arising from the short-range nature of fitted repulsives,
one may try to handle the middle-range problems with middle-range repulsives,
and the dissociation problem with one-body and pairwise long-range terms. At
least for the first look.
However, infinite-range (dissociation) limit of energy correction cannot be
treated with pairwise terms. Energy being an extensive quantity, it must be an
additive function of stoichiometry when separated systems are added together
10 The argumentation of this paragraph is rather sloppy, preparing the unprecise conclusion
that one-body repulsive energy terms are the virtually only correct solution to the dissociation
energy problem. Carefully stated, the remanent dissociation energy discrepancy is present in
the practical, not the actual dissociation limit, the state where atom pairs have reached distances from each other outside the ranges of respective repulsive potentials (this is a practical
r state from the point of view of repulsive fitting although all r values are finite, of
course.) Therefore in fact, using one-body energies is the only mathematically acceptable way
of handling this problem if one accepts that any additional information coming from atomic
pairs being outside of their respective designated repulsive pair potentials must not influence
the physics of DFTB systems. On the contrary, the basic problem of traditional overbinding
strategy lies in carrying a hidden knowledge about which atom pairs can be, will be or have
ever been in certain types of bonds in their history even beyond the distance of their respective
repulsive ranges. This footnote has been added to the text as a means of clarifying a possible
logical flaw after submitting the thesis for supervision.
60
Figure 2.3: The Cl-Cl dissociation curve calculated with the non-overbinding
and the one-body-corrected repulsive of Figure 2.2 (the dotted and the dashed
strong lines, respectively). The thin solid line is the DFT reference, and the
dashed one is what it seems to be without dissociation adjustment. (Note the
middle-range difference between the reference and the dashed strong line also
here.)
E [a.u.]
0.1
-0.1
-0.2
r []
E (dissoc) (Zl Zm
. . . Zn + Zp Zr . . . Zs )
= E (dissoc) (Zl Zm
. . . Zn ) + E (dissoc) (Zp Zr . . . Zs )
(2.46)
It can be easily derived from the above formula, that only one-body repulsive energy terms fulfill the additivity criterion and can make the two sides of (2.44) if
not equal, consistent between reference and DFTB. Therefore, a correct, consistent replacement of traditional overbinding will use one-body repulsive11 energy
parts.
On top of one-body regime, the non-dissociation (short and middle-range)
repulsive part in question must be fitted as a sum of pairwise profiles. The
fact that pairwise repulsives are not only a short-range phaenomenon can also
be seen in Figure 2.2 and Figure 2.3. Unfortunately, the attempts made at
fitting middle-range repulsive pair potentials are highly limited by a rather
practical circumstance: DFTB and most DFT methods tend not to converge
when atoms leave each others region of influence. The chlorine-chlorine dimer
used to draw Figure 2.2 and Figure 2.3 was a lucky example to present a longrange behaviour of pair energy, but the usual course of these calculations leads
to heavy convergence problems even with simple molecules of chlorine and other
11 Although it is meaningless to call a one-body energy repulsive, we keep this terminology
as a name for elements of the parametrized, quasi-classical part of DFTB total energy.
61
Erep =
ZZ , XZZ , +
UZ NZ
(2.47)
ZZ ,
where UZ is the one-body repulsive for atom type Z and NZ is the number
of such atoms in the fit system in question. As it can be seen from the above
composition, one-body terms behave just like two-body terms in the fit if UZ s
go along s and NZ s along Xes. Thus the (2.7) X and A matrices of fitting
are modified so:
(1)
(2)
XHH,1 XHH,1
(1)
(2)
XHH,2 XHH,2
.
..
(2)
(1)
XCH,1 XCH,1
(1)
X=
(2.48a)
XCH,2 . . . . . . . . . . .
.
.
.
(1)
(2)
NH
NH
(1)
(2)
NC
NC
..
.
HH,1
HH,2
..
.
CH,1
A = CH,2
(2.48b)
..
.
UH
UC
..
.
where we remained at the example composition of systems containing H and
C. With the above modifications, the fitting algorithm can fit for one-body
terms in the way presented for two-body terms. Of course, fit targets that are
homogeneous, in the sense that they do not depend on the absolute value of
62
2.4
2.4.1
Theory
As it was seen in (1.33), the repulsive energy is the sum of the so-called counterterms of double-counting terms in KohnSham energy (for valence electrons),
as well as the total energy of core electrons. These together formed
1
Erep = Tr Pvalence VC Vxc
2
KS 1 VC Vxc
+ Exc + Enuc
+ Tr Pcores H
2
63
(2.49)
+
dy
2
|x y|
x
(x)((y) + (y))
+
dy
|x y|
y
+(x)Vext (x)
(2.50)
1 (x)((y) + (y))
dy
2
|x y|
y
1 (x)(y)
dy +
2 |x y|
1 (x)(y)
dy
2 |x y|
(2.51)
A=B
A B
+
r
A B
r
(A + B + A + B )(A + B + A + B )
(A + B )Vxc (A + B + A + B )
ZA ZB
+ A Vext,B + B Vext,A +
rAB
(2.52)
(with spatial arguments of the integrands omitted, and |x y| of (2.51) substituted with r). In order to maintain the zero dissociation limit of repulsive
pair profiles (see 2.3.2), we must subtract its dissociation limit from the pair
repulsive energy. Thus
(A + A )(A + A ) + A Vxc (A + A )
(B + B )(B + B ) + B Vxc (B + B ).
(2.53)
With this definition of UAB it remains true within the precision of negligible
three-center terms (see the argumentation about density superposition at the
end of 1.2.1) that for any chemical system
Erep =
1
UAB (rAB )
2
(2.54)
A=B
(dissoc)
Erep
=
UA =
(A + A )(A + A ) A Vxc (A + A ) .
A
(2.55)
If one pursues absolute energy level compatibility (and comparability) with
all-electron DFT calculations, the constant terms should contain all the oneatom contributions in the above sum. The terms that constitute the atomic
part of the full repulsive energy and must be used instead of the UA s in (2.55)
are
The kinetic energy of core electrons:
T(A) = Tr(Pcore,A T) =
core of A
65
|T |
(2.56)
(2.57)
x y
(2.58)
x y
(A)
(A)
Exc
EKS,xc = (A + A )(A + A ) A Vxc (A + A )
(2.59)
Eext,core = A Vext,A
2.4.2
The specific structure of DFTB machinery gives two more particular aspects
which must be considered at repulsive calculations. First, the calculated repulsive energy uses atomic quantities that need to come from distinct pseudoatomic
calculations. Exchange-correlational potentials and energy densities come from
pseudoatoms compressed with the density compression radius. Densities used
as the source of Coulombic potentials also come from these pseudoatoms. Densities, however, which counterbalance P terms in counterterms, must be compressed with the wavefunction compression radius. Core densities that are not
used to build potentials do not play a role in the Hamiltonian and the Kohn
Sham energy, and they can be attributed a quite arbitrary compression. To
avoid further complication of the structure of repulsive calculations, we make
them behave always parallel to valence densities. Thus the pair repulsive becomes
UAB =
1
A B
1
A B
1
A B
1
A B
+
+
r
2
r
2
r
2
r
+ (A
+ B
+ A + B )(A
+ B
+ A + B )
(A + B )Vxc (A
+ B
+ A + B )
ZA ZB
+ A
Vext,B + B
Vext,A +
rAB
(A
+ A )(A
+ A ) + A Vxc (A
+ A )
(B
+ B )(B
+ B ) + B Vxc (B
+ B ) (2.60)
66
r
2
r
core of A
+ (A + A )(A + A ) A Vxc (A + A ) + A
Vext,A ,
1
|T | +
2
(2.61)
where quantities marked with come from an atomic calculation with the socalled density compression radius (which would be better called potential compression radius), and those marked with come from atomic calculations compressed with the so-called wavefunction compression radius. Later, it may be
needed to bring in more tunable parameters in calculated repulsives. To fulfill this need, we will let the potential and wavefunction compression radii of
repulsive calculation (r2 and r3 ) be other than the same radii of Hamiltonian
construction (r0 and r1 as it was introduced in 1.2.1).
The second technical detail that needs consideration here, is how potential
and density superposition affects calculating repulsive potentials. Clearly, this
question affects the
(A +B +A +B )(A +B +A +B ) (A +B )Vxc (A +B +A +B )
(2.62)
part of repulsives. In density superposition, the formulae to calculate the xc part
of repulsives are literally as above, densities are superposed in the arguments
of xc-density-type quantities ( and Vxc ). In potential superposition, the dimer
xc potential is the Vxc (A + A ) + Vxc (B + B ) superposition of atomic ones,
while the dimer can be two different ways. Either we leave the above (A +
B + A + B ) xc energy density because we do not want to tinker with the
inner structure of , or we can use the
(A + B + A + B ) =
(A + A )(A + A ) + (B + B )(B + B )
(2.63)
A + A + B + B
weighted sum of atomic xc energy densities (which gives back the potentialsuperposed Vxc as its derivative with respect to , as well as it gives back the
dimer xc energy as the sum of atomic xc energies). In our attempts made so far,
we have not tried and experimented with the potential superposition schemes
proposed here.
2.4.3
Results
As a first example of CC repulsive calculation, we took the electronic parameters of the mio set (r0 = 7 a.u., r1 = 2.7 a.u.), and optimized the wavefunction compression radius used for the repulsive, independent of the wavefunction
compression radius used for the Hamiltonian (the potential compression used
for repulsive remained equal with the potential compression used for electronic
tables, r2 = r0 ). This choice was an ad hoc first trial to calculate repulsives
after realizing that letting both compression radii for making repulsives equal to
the electronic tabulation counterpart gives unusable results (this problem may
be an effect of choosing compression radii rather freely in the world of fitted
67
12 Of course, the brute-force optimization process was a little bit altered to use calculated
repulsives instead of fitted ones: the fitted part was a dummy set while the electronic part
contained all the information about calculated repulsives.
68
Figure 2.4: The CC dimer dissociation curve. The solid line is the DFT reference, and the dashed line is the one produced with DFTB and a calculated repulsive where electronic parameters of the mio set were used for electronic tables,
and the potential compression radius for repulsive was equal to the tabulation
one (r2 = r0 ). The wavefunction compression radius for repulsive calculation
was 0.2 a.u. in the top graph (for a better comparison of curve shapes, both
curves were shifted to an arbitrary absolute level determined by the one-body
repulsives used here). The same parameter of the bottom graph was optimized
further to 0.01 a.u. to give better geometries (this latter parametrization was
used for test calculations).
E [a.u.]
-150
-150.5
-151
1.5
2.5
r []
E [a.u.]
1.5
0.5
1.5
r []
69
ref.
DFTB
error
rel. error
-18.266
1.09302
-15.344
1.10835
2.92274
0.01533
-0.1600
0.01403
-30.881
1.53061
1.09615
-26.469
1.58222
1.11910
4.41151
0.05162
0.02295
-0.1429
0.03372
0.02093
-24.437
1.33103
1.08743
-21.201
1.31039
1.11682
3.23639
-0.0206
0.02940
-0.1324
-0.0155
0.02703
-17.485
1.205
1.06653
-16.442
1.15838
1.08725
1.04324
-0.0466
0.02073
-0.0597
-0.0387
0.01944
-59.083
1.39663
1.08699
-53.143
1.41683
1.12142
5.94045
0.02021
0.03444
-0.1005
0.01447
0.03168
-55.967
1.54680
1.53643
1.09667
-48.644
1.60341
1.59862
1.11864
7.32247
0.05661
0.06219
0.02196
-0.1308
0.03660
0.04048
0.02003
-56.235
1.53530
1.09665
-48.835
1.59850
1.11919
7.40019
0.06320
0.02254
-0.1316
0.04116
0.02055
1.555
1.62533
0.08533
0.05541
-49.524
1.55728
1.09484
-43.599
1.62048
1.12521
5.92536
0.06319
0.03037
-0.1196
0.04058
0.02774
-50.066
1.50876
1.33658
1.09898
1.08736
-43.821
1.57347
1.3234
1.12092
1.11584
6.24536
0.06471
-0.0132
0.02194
0.02848
-0.1247
0.04289
-0.0099
0.01996
0.02619
70
bicyclobutane
E
CC (edge)
CC (middle)
CH (in CH2 )
CH (in CH)
cyclobutene
E
CC
C=C
CH (in CH2 )
CH (in CH)
cyclopropane
E
CC
CH
propane
E
CC
CH (end)
CH (middle)
cyclopropene
E
CC
C=C
CH (opposite to C=C)
CH (neighbour to C=C)
spiropentane
E
CC (radial)
CC (outer)
CH
methylene-cyclopropane
E
C=C
CC (radial)
CC (outer)
CH (in CH2 )
CH (on ring)
propadiene
E
C=C
CH
-39.222
1.51018
1.90017
1.11157
1.09512
-34.175
1.57435
2.08801
1.15735
1.12634
5.04709
0.06418
0.18785
0.04578
0.03122
-0.1287
0.04250
0.09886
0.04119
0.02851
-43.155
1.57318
1.51869
1.09709
1.08716
-38.052
1.64547
1.59169
1.12671
1.12103
5.10254
0.07229
0.07299
0.02962
0.03387
-0.1182
0.04595
0.04806
0.02700
0.03115
-36.872
1.50922
1.08659
-31.570
1.57198
1.11653
5.302
0.06276
0.02994
-0.1438
0.04158
0.02755
-43.548
1.53199
1.09717
1.09854
-37.630
1.59033
1.11925
1.12985
5.91752
0.05834
0.02208
0.03131
-0.1359
0.03808
0.02013
0.02850
-29.452
1.50826
1.29488
1.09518
1.07983
-25.360
1.57681
1.30087
1.12367
1.11128
4.09168
0.06855
0.00599
0.02849
0.03145
-0.1389
0.04545
0.00463
0.02601
0.02913
-55.408
1.48491
1.52990
1.08815
-47.655
1.55222
1.60192
1.11817
7.75290
0.06731
0.07202
0.03002
-0.1399
0.04533
0.04707
0.02758
-42.934
1.32232
1.46992
1.53954
1.08776
1.08928
-37.323
1.30854
1.53619
1.60484
1.11987
1.11842
5.61093
-0.0138
0.06627
0.06530
0.03210
0.02914
-0.1307
-0.0104
0.04508
0.04242
0.02951
0.02675
-30.567
1.30686
1.08824
-27.153
1.29116
1.12033
3.41345
-0.0157
0.03209
-0.1117
-0.0120
0.02949
71
1,3-butadiene
E
CC
C=C
CH (middle)
CH (end)
2-butyne
CC
CC
CH
propyne
E
CC
CC
CH (in CH3 )
CH (in CH)
propene
E
CC
C=C
CH (in CH3 )
CH (in CH2 )
CH (in CH)
2.4.4
-39.068
1.43937
1.39158
1.08870
1.08596
-34.610
1.49511
1.36317
1.12067
1.12845
4.45754
0.05574
-0.0284
0.03197
0.04249
-0.1141
0.03872
-0.0204
0.02937
0.03913
1.46169
1.20920
1.09707
1.51404
1.16399
1.12084
0.05235
-0.0452
0.02377
0.03582
-0.0374
0.02167
-30.483
1.46027
1.20715
1.09689
1.06610
-27.785
1.51272
1.16164
1.12004
1.08671
2.69794
0.05245
-0.0455
0.02314
0.02061
-0.0885
0.03592
-0.0377
0.02110
0.01933
-37.260
1.50206
1.33336
1.09846
1.08675
1.09114
-32.486
1.56603
1.31711
1.12126
1.11577
1.12569
4.77352
0.06397
-0.0163
0.02280
0.02902
0.03455
-0.1281
0.04259
-0.0122
0.02076
0.02670
0.03166
The results shown here illustrate that the proposed machinery of calculating
repulsive potentials instead of fitting them is able to produce parametrizations
that give less precise results than the best fitted repulsive sets, but they still
ensure reasonable, qualitatively good results right at the first glance.
The calculated repulsives still contain parameters that allow semiempirical
tuning of them to ensure the least difference with respect to reference methods,
but optimization of them means a huge simplification of the parametrization process compared to fitting pair potentials. First, they are two new real numbers
per atom, instead of the pair potentials, that have infinite degrees of freedom;
and second, they are atomic, not pairwise parameters. Once the transferability
of these new parameters is proved (this is foreshadowed by the fact that CC
repulsives got from dimer optimization were the same as the CC parameters
from the CH optimization), their atomwise nature ensures that extending a
parameter set to a new atomic species means tuning only one set of atomic parameters (two compression radii), not a complete set of new pairwise repulsives
between the new species and all the previous members of the parameter set.
The electronic parameters for calculating HCO systems optimized in our
first usages of the repulsive calculation scheme seemingly leave more space for
further optimization. Besides the performance of the calculated parameters,
that is surprisingly good compared with the simplicity of theory, but still lag
72
73
74
Chapter 3
Enhancement proposals to
SCC-DFTB
3.1
1
(y)(y)(x)dxdy
(x)
(3.1)
|x y|
x,y
(x)(y)
dxdy
(3.2)
|x y|
x,y
m n
1
xm yn |x y| x0 ,y0
m
(x x0 ) (x)dx (y y0 )n (y)dy
x
(3.3)
1
kernel. After rearranging
by simply using its Taylor series instead of the |xy|
the off-site part of the (1.48) sum (quite similarly to the rearrangement that led
to the (1.55) original off-site SCC formula) and using multipole-series expansions
instead of the (3.1) integrals, we get a multipole extension of (1.55). It is the
same as if we took the atomic Mulliken densities defined in (1.59) as and ,
and used the multipole expansion of their interaction integrals. Of them, the
75
{A}, {B}.
(3.4)
Plausibly we choose atomic centres A and B as base points of the Taylor series
in the multipole expansion, and so the first terms of the series taking the place
of (3.4) are
1
|xA xB | S S
(A)
(B)
xA xB
|xA xB |3 S X
(A) (B)
A xB )(xA xB )
|xA xB | 2(x
X X + . . .
|xA xB |5
xB
|xxAAx
3 X S +
B|
(3.5)
(A)
(A)
X = |X XA |x = (x)(x
xA )(x)dx.
(3.6)
x
1 (00)
(10)
AB qA qB + AB dA qB
2
A,B
(01)
(11)
+ AB qA dB + AB dA dB +
where
(mn)
AB
1 m n
n AB ,
m!n! xm
A xB
1
(A)
(A)
dA =
ni ci, ci, X + ci, ci, X .
2
(3.7)
(3.8)
(3.9)
{A},,i
76
3.2. SEMIEMPIRICAL
In the following, we will use the very first terms above the original SCC level,
the monopole-dipole ones. The construction can be extended to higher terms
as well. Up to the desired monopole-dipole level, the SCC Hamiltonian matrix
becomes (cf. (1.71) in the monopole-monopole approximation)
(00)
1
(00)
S
[]B + []B qB
2
B
(01)
1
(01)
+ S
[]B + []B dB
2
B
1 [] (10)
[] (10)
+
X []B + X []B qB .
2
H =
(3.10)
The transformation rules needed to relate the (3.6) position matrix elements
in the molecular coordinate system to the tabulated ones are presented in 3.3.3.
3.2
3.2.1
A
(3.11)
(ZA , ZB , r) =
qB r
where atoms A and B, being of types ZA and ZB , are in a distance of r, A
is the highest occupied electronic energy level of the atom A, and qB is the
total electronic population of atom B. Similarly to the calculation of chemical
hardness in traditional SCC scheme, an excess charge on atom B is created by
varying the population on its highest occupied orbital.
Because in a self-consistent diatomic calculation the energy levels cannot
be attributed to specific atoms, the derivative in (3.11) can be calculated in
a perturbative way. If A is the atomic state behind A and VB is the total
77
A |VB |A
A
=
.
qB r
qB
3.2.2
(3.12)
qB,l
2l + 1
qB,l
r
where l and l are the angular momentum indices of the interacting orbitals of
atoms A and B, respectively, and m is the magnetic quantum number of the
participating A orbitals, over which we make a mean because we want to have
a general l-l interaction profile (angular dependence of SCC interaction which
arises due to uneven filling of different-m orbitals is covered by multipole SCC).
We do not have to average over magnetic quantum numbers of atom B since
VB (x) is calculated with a spherically symmetric atom.
3.2.3
Fully resolved s
|V[] |
=
q
q
(3.14)
3.2.4
Although there is a strong competition between computational costs and calculational precision, that must result in a fair compromise, it is not worthless
2 This section takes motivation from discussions with Adriel Dominguez and Thomas
Niehaus.
78
3.2. SEMIEMPIRICAL
investigating where the further resolution of s can theoretically lead to. Some
aspects of such a further resolution are presented in the following.
Semioperator s
One step beyond the Mulliken-analysis-based improvement of more and more
resolved s, one can begin resolving them at least on one side to get fullyfledged orbital-indexed quantities, that is, operators written in the LCAO space,
not only the diagonal elements described in 3.2.3. This resolution beyond every
limit refines (3.14) on one side to a well-behaved orbital-indexed matrix:
; =
|V[] |
H
=
.
q
q
(3.15)
1
P ; q
2
(3.16)
i,,,
would represent the SCC term, where P is the change of density matrix
(0)
element i ni ci, ci, with respect to P , and
; q
(3.17)
KS
H
=
; q .
(3.18)
H
|VB |
=
.
qB
qB
(3.19)
The number of the above s is only |{A}| |{A}| per distance step per atom
pair type (this means practically |{A}| |{A}| + |{B}| |{B}| entries in an
A-B-type table row).
Besides the (3.15) arrangement where the three indices of came from centres AA; B, we must shortly define all the other possible combinations of centres
3 This number of
; values is valid when cases with [] = [] (i.e. three-center-type
ones) are neglected.
79
(0)
(1)
KS
H
=
;B qB + ;B dB .
(3.20)
3.2. SEMIEMPIRICAL
, {B} because there is no distance parameter in chemical hardness with
respect to which this profile could be derived, as well as there are no nuclei able
to be moved with respect to each other. ;B when {B} but
/ {B} is a
similar complicated case: the problem of handling of higher multipole order s
also applies here because {B}.
There is, however, a possible way to circumvent the calculation of multipole
contributions with semi-operator s. Because of the symmetry between the two
sides of (recall that they are nothing else than a complete set of two-electron
integrals, or in other words, elements of a kernel of a second derivative of energy),
a total sum of all possible multipole contributions above monopole-monopole
level in SCC-DFTB can be approximated (exactly at the monopole-dipole level,
and with gradually increasing relative errors at higher levels) by
2 Eall multipoles = 2(2 Esemi-operator 2 Epoint charges ),
(3.21)
where the first term is calculated without any attempt to handle multipole
corrections, and the last term is a point-like-charges SCC energy calculated
with adequately refined s (in (3.11), (3.13) or (3.14)). This makes handling
multipoles even with semi-operator s unnecessary, because
2 E = 2 Epoint charges + 2 Eall multipoles
= 22 Esemi-operator 2 Epoint charges ,
(3.22)
containing the old SCC energy in addition the the semi-operator one, which can
be calculated far more easily than a multipole extension of semi-operator s.
Two-sided exact operators and their approximations implemented
by other schemes
Naturally, one can proceed to refine both sides of . This is nothing else than
giving back the (1.48) two-electron integrals faithfully in the SCC part. While
this means an extremely populous set of ; parameter tables, one can consider neglecting three-center terms also here, arriving at the still very large
|{A}| |{A}| |{B}| |{B}| number of table entries per atom pair per distance
value (this is only an estimation of magnitude, also here). Of course, this regime
makes multipole calculations away, and it gives a possibility of calculating HF
exchange in a more perfect way.
A useful byproduct of the investigation of operator-like s made so far is
that we are now able to compare them with Mulliken-level s and clarify what
and how is simplified in point-like-charges SCC, and what and how is improved
in the multipole extension of it. The two-sided operator-like s represent the
(1.48) four-orbital integrals without any approximation beyond the basic DFTB
rules, but with the second derivative of xc energy added to the Coulomb kernel,
of course:
2 Exc
1
+
(y)(y)dydx.
(3.23)
; =
(x)(x)
|x y| (x)(y)
x y
The fully-resolved, but not operator-like s in 3.2.3 simplify the above plenty
of integrals to
1
2 Exc
=
(x)(x)
+
(y)(y)dydx,
(3.24)
|x y| (x)(y)
x y
81
1
S ( + + + )S .
4
(3.25)
(3.26)
()
1
(; + ; ) S .
2
(3.27)
1
2 Exc
; =
(x)(x)
+
(y)(y)dydx
|x y| (x)(y)
(3.28)
(3.29)
x y
3.3
3.3.1
(3.31)
qA = TrPA = x| PA |x dx = A (x)dx
x
(3.32)
derivation of off-site SCC Coulomb energy with atomwise Mulliken density operators would also have been a bit shorter in 1.2.2, but it would have hidden
some intuitive details.
To be seen in a broader view, the Mulliken density operators give a decomposition of the P density operator using a complete set of projections in our
LCAO space.
1
)
PA =
( P + P
(3.33)
2
{A}
where
=
|| S 1
(3.34)
and
= 1.
(3.35)
P .
(3.37)
{A}
Orbitalwise Mulliken density operators imply orbitalwise Mulliken charge densities and charges, and they may be useful at the investigation of the totally
resolved s in section 3.2.3.
As an intermediate step between the atomwise and the totally resolved
regime, there can be Mulliken densities defined which are resolved only up to
the angular momenta of orbitals. They are instruments of studying the almosttotally refined SCC in section 3.2.2.
83
3.3.2
Shape
For the original shape, see [13]. As the new shapes that we propose in this
paper are semi-empirical, there must be analytical shapes (similar to the original
ones in [13] or not) fitted to the semi-empirical data in order to be able to carry
out derivations on them, if one decides to use them.
Either in the interpolated or in the semiempirical case, the multipole expansion of SCC energy requires derivatives of basic (00) as higher-order s
(00)
(see (3.8)). Since AB (ZA , ZB , xA , xB ) depends on its xA and xB arguments
as a spherically symmetric function of r = |r| = |xA xB | (for brevity we use
r instead of xA xB and we drop the atom-label arguments in the following
equations), the higher-order s are quite simple, e.g.
(10) (r) =
r (00) (r)
r
r
(3.38)
(r) =
(3.39)
The above formulae can be seen in (3.5) applied to the 1r long-range limit potential in the place of a proper (00) .
As we can see from the above too, higher-order s are matrices according to
their multipole indices. (mn) is a matrix of order m + n, e.g. (10) is a vector
(not mentioning the atomic centre indices of s when they are listed as AB ).
Identities
As every is a AB = (ZA , ZB , rAB ) = (ZA , ZB , |xA xB |), it must hold
that
AB
AB
=
.
(3.40)
xA
xB
From physical reasons (from the fact that nothing must depend on the listing
order of atoms or from the second-derivative nature of s) it is also obvious
that
(mn)
(nm)
AB = BA .
(3.41)
In the multipole expansion of SCC energy, the s were coefficients of a Taylor
series, therefore
(mn)
(m+1,n)
(m + 1)AB
AB
xA
(mn)
AB
xB
(m,n+1)
= (n + 1)AB
(3.42)
and hence
(mn)
AB
= (1)
mn (nm)
AB .
(3.43)
mn (mn)
BA .
(3.44)
AB
= (1)
FB, = qB
(00)
AB
qA
xB,
A=B
dB
(10)
(01)
AB
AB
qA qB
dA .
xB,
xB,
(3.45)
A=B
A=B
Writing all the vector quantities by components and using the identities of s,
()
FB, = qB
(01)
AB, qA
A=B
dB,
(02)
AB, qA
A=B
qB
(02)
AB, dA, .
(3.46)
A=B
3.3.3
In current SCC-DFTB implementations, there are two matrices of quantummechanical integrals tabulated for run-time use with DFTB; the Hamiltonian
matrix and the overlap matrix. Using them needs a transformation because the
coordinate system in which the tabulated integrals have been carried out is not
the same as the molecular coordinate system in which the DFTB calculation
takes place (usually, the system of integration is fixed to the line between the
participating two atoms, while the molecular one is quite arbitrary).
The overlap and the position matrix
In this section we recall the usage rules of overlap matrix tables in a formalized
way, and parallel with it, we construct the similar rules for the position operator
matrix.
,
Let us call the molecular basis , , etc. and the integration basis ,
etc.
From the relative orientations of the molecular and the integration coordinate
systems, a linear transformation can be set up between the two:
=
T
(3.47)
in which orbitals of different angular momenta do not mix, i.e. s, p, etc. orbitals
of the molecular basis are linear combinations of integration basis elements of
85
T ST
(3.48)
S =
,
t x
(3.49)
is the transformation between the two, a position matrix element with origin a
in the molecular system is
(a)
X, =
(x)(x
a )(x)dx
T(x)
t (x b ) + b a (x)T
dx
,
,
x
(b)
+ S (b a )
Tt T
X,
(3.50)
,
,
=
T
(3.51)
T
86
T
=
TT
(3.52)
T
V[] [] (x)
dx
[] (x) q
(3.53)
1
(P S + P S ) ,
2
(3.54)
3.4
1
2 Exc
S
j(y)b(y)dydx
Kia,jb =
i(x)a(x)
+
(3.55)
|x y| (x)(y)
x y
4
On the usage and construction of the coupling matrix, see section 3 in [28].
87
S
Kia,jb
=
ci, cj, ca, cb, ; ,
(3.56)
,,,
where |i =
ci, | etc. with the ; of (3.26), the structure of the
S
improved Kia,jb
shall be also clear from (3.26).
It must be noted here, however, that the coupling matrix of spin triplet
excitations cannot be improved by the multipole expansion introduced in 3.1.
As it was written in 1.2.3, spin-spin interaction terms in DFTB are confined
2 Exc
. In addition to this,
to on-site terms, because of the short range of (x)(y)
on-site multipole-multipole interactions are zero in our approximation above
the monopole-monopole level because the (1.62) semiempirical calculation of
on-site second-order energy change (the chemical hardness, and its refinements
introduced later) includes every level of multipole-multipole interactions. This
is also why our new higher-order s tend to zero when r = 0. Nonetheless, if one
wants to make improvements in this part of DFTB, there remains a possibility
of using fitted monopole-dipole, etc. interaction strengths, if the underlying
physics needs it. The more promising way of improving the description of triplet
excitations is including spin-orbit interaction in TD-DFTB. Using some kind of
operator-like s (see 3.2.4) would be a major improvement in this field, but at
rather high cost.
The second aspect of making TD-DFTB better by the multipole SCC which
we treat here is the improvement of its oscillator strength expression. The
oscillator strength of a transition from state i to a is proportional to the square
of a transitional dipole moment
2
2
fia
(3.57)
D(ia) =
i|X |a
is the component
where = x, y, z is the index of spatial coordinates, and X
of position operator. The above matrix element can be easily calculated with the
aid of the previously
stored position
atoms
(A)
where R
atoms
{A}
atoms
,{A}
| =
ci, ca, |X
{B}
(A)
(A)
R
ci, ca, |(X
) + R
|+
atoms
atoms
D(ia) =
atoms
1
2
A
(ia)
+ D
=
2
(A)
(A)
(A)
(A)
ci, ca, (X, + S R
) + ci, ca, (X, + S R
)=
{A},
atoms
(ia)
dA
(ia)
qA R(A)
(3.59)
where R(A) is the vector of the position of atom A, and the first, trivial step
is didactically included in order to symbolise that half of the sum over A, B
and , is made with a A B, relabelling of the summand. The
quantities defined in the last step are called transitional Mulliken charges and
dipole moments between states i and a.
89
Summary
In this dissertation, I treat two areas of improving the DFTB method. The
first area of interest is the repulsive part of DFTB energy. There has been
a parametrizer automaton developed that performs a least-squares fitting of
repulsive potentials as linear combinations of basis profiles through a linear
regression of coefficients. It makes fitting of repulsive profiles to multiple fitsystems incomparably easier than before; that is illustrated by several examples
as successful usages of the fitter. The automaton is also able to tune electronic
parameters. To make the fitting process truly automatic as seen from the user
side, the parametrizer does many auxiliary tasks automatically as well, e.g. fitsystem generation. Among the objectives of fits, not only energies can be used,
but several other energetical properties (mainly but not exclusively, derivatives
of energy). User-input weighting of fit targets ensures the most precise fitting
to the most important properties of most important fitsystems selected by the
user.
As a further question regarding repulsive energy, its structure is also treated
here briefly. It is quite clear now that in addition to the pairwise repulsive
profiles, one-body repulsive energy terms can be used too to tune the dissociation
behaviour of DFTB. In addition to short-range pair repulsives and one-body
terms, there must be middle-range pairwise repulsives also used, but fitting
them is not (yet?) feasible.
As a perfect way of building repulsives, calculation of them instead of fitting
is also proposed. After deriving the formulas needed for repulsive calculation and
implementing them, the first results suggest that this scheme is able to be used
to get at least first-glance repulsives for any set of elements that has not been
parametrized with DFTB before, if one optimizes atomic electronic parameters
successfully. Being able to make transferable parametrizations depending on
atomic, not pairwise parameters gives the prospect of producing and extending
parametrizations including large sets of elements with reasonable effort.
The second area of DFTB improvement that is investigated here is making
its SCC part better. A proposition of a multipole expansion of SCC energy
instead of the currently used point-like charges is made as well as a proposition to calculate the effective interaction profiles between atomic excess charges
semiempirically, instead of the current heuristical interpolation. Fully derived
formulae of the above improvements are ready to implement in SCC. Furthermore, improvements to time-dependent DFTB are also introduced briefly.
90
91
Acknowledgements
It is hardly possible to write down how deeply I am grateful to the colleagues
who made my doctoral research possible and who helped it. First, I thank to
Professor Thomas Frauenheim to give me the opportunity of working at the
Bremen Center of Materials Science as a PhD student. As large parts of the
research has been carried out in working together with Balint Aradi, I must
acknowledge that my results would have been incomparably less without the
help of him.
In addition to the above, cooperations and other discussions canalized a lot of
helpful stimuli and other useful impressions from Gotthard Seifert, Martin Persson, Karl Jalkanen, Grygoriy Dolgonos, Thomas Niehaus, Adriel Dominguez,
Michael Gaus and Tom
as Kubar into my research.
Many thanks shall go to Christof Kohler and Zoltan Hajnal for facilities I
was able to use in Bremen and temporarily in Budapest, respectively.
Last but not least, my thanks go to my family, as well as to my friends in
Bremen.
92
Bibliography
[1] NIST computational chemistry comparison and benchmark database. NIST
Standard Reference Database Number 101, http://cccbdb.nist.gov. Release 15b, August 2011, Editor: Russell D. Johnson III.
[2] B. Aradi, B. Hourahine, and Th. Frauenheim. DFTB+, a sparse-matrixbased implementation of the DFTB method. J. Phys. Chem. A, 111:5678
5684, 2007.
[3] Nathan Argaman and Guy Makov. Density functional theory an introduction. 1999. arXiv:physics/9806013v2.
[4] Zolt
an Bodrog and Balint Aradi. Possible improvements to the selfconsistent-charges density-functional tight-binding method within the second order. Phys. Status Solidi (b), 249:259269, 2012.
[5] Zolt
an Bodrog, B
alint Aradi, and Thomas Frauenheim. Automated repulsive parametrization for the DFTB method. J. Chem. Theory Comput.,
7:26542664, 2011.
[6] Zolt
an Bodrog, B
alint Aradi, and Thomas Frauenheim. Atomic-parameterbased repulsive potentials for the density-functional tight-binding method.
2012. To be published soon.
zek. A new method of approximative calculation of polycentric inte[7] Jir C
grals used in the quantum mechanical study of molecular structure. Mol.
Phys., 6:1931, 1963.
[8] Larry A. Curtiss, Krishnan Raghavachari, Paul C. Redfern, and John A.
Pople. Assessment of Gaussian-2 and density functional theories for the
computation of enthalpies of formation. J. Chem. Phys., 106:10631079,
1997.
[9] P. A. M. Dirac. Note on exchange phenomena in the Thomas atom. Proc.
Cambridge Phil. Soc, 26:376385, 1930.
[10] Grygoriy Dolgonos, Balint Aradi, Ney H. Moreira, and Thomas Frauenheim. An improved self-consistent-charges density-functional tight-binding
(SCC-DFTB) set of parameters for simulation of bulk and molecular systems involving titanium. J. Chem. Theory Comput., 6:266278, 2010.
[11] R. Dovesi, M. Causa, R. Orlando, C. Roetti, and V.R. Saunders. Ab-initio
approach to molecular-crystals a periodic hartree-fock study of crystalline
urea. J. Chem. Phys., 92:74027411, 1990.
93
E
ni
[23] Tom
as Kubar, Zolt
an Bodrog, and Michael Gaus. Density-functional tightbinding parameters for halogen-organic chemistry. 2012. To be published
soon.
[24] Christof K
ohler, Thomas Frauenheim, Ben Hourahine, Gotthard Seifert,
and Michael Sternberg. Treatment of collinear and noncollinear electron
spin within an approximate density functional based method. J. Phys.
Chem. A, 111:56225629, 2007.
[25] Christof K
ohler, Gotthard Seifert, and Thomas Frauenheim. Densityfunctional-based calculations for Fe(n), (n 32). Chem. Phys., 309:2331,
2005.
[26] Andre Mirtschink. Berechnungen von Bindungsenergien zweiatomiger
Molek
ule der Elemente der ersten und zweiten Periode im Rahmen der
DFTB-Methode. Masters thesis, Technische Universitat Dresden, 2010.
[27] Ney H. Moreira, Grygoriy Dolgonos, Balint Aradi, Andreia L. da Rosa,
and Thomas Frauenheim. Toward an accurate density-functional tightbinding description of zinc-containing compounds. J. Chem. Theory Comput., 5:605614, 2009.
[28] Thomas Niehaus. Approximate time-dependent density functional theory.
J. Mol. Struct. THEOCHEM, 914:3849, 2009.
[29] Robert G. Parr and Weitao Yang. Density-Functional Theory of Atoms
and Molecules. Oxford University Press, 1994.
[30] J. P. Perdew, K. Burke, and M. Ernzerhof. Generalized gradient approximation made simple. Phys. Rev. Lett., 77:38653868, 1996.
[31] Dirk Porezag, Thomas Frauenheim, Thomas Kohler, Gotthard Seifert, and
R. Kaschner. Construction of tight-binding-like potentials on the basis of
density-functional theory: Application to carbon. Phys. Rev. B, 51:12947,
1995.
[32] V.R. Saunders, R. Dovesi, C. Roetti, R. Orlando, C. M. Zicovich-Wilson,
N.M. Harrson, K. Doll, B. Civalleri, I. Bush, Ph. DArco, and M. Llunell.
CRYSTAL2003 Users Manual. University of Torino, Torino, 2003.
[33] G. Seifert. Tight-binding density functional theory: An approximate Kohn
Sham DFT scheme. J. Phys. Chem. A, 111:56095613, 2007.
[34] Jose M. Soler, Emilio Artacho, Julian D. Gale, Alberto Garca, Javier Junquera, Pablo Ordejon, and Daniel Sanchez-Portal. The SIESTA method
for ab initio order-N materials simulation. J. Phys., 14:2745, 2002.
[35] L. H. Thomas. The calculation of atomic fields. Proc. Cambridge Phil.
Soc., 23:542548, 1927.
[36] M. Valiev, E.J. Bylaska, N. Govind, K. Kowalski, T.P. Straatsma, H.J.J.
van Dam, D. Wang, J. Nieplocha, E. Apra, T.L. Windus, and W.A. de Jong.
NWChem: a comprehensive and scalable open-source solution for large
scale molecular simulations. Comput. Phys. Commun., 181:14771489,
2010.
95
96