initial setup of thesis repository
[cluster_expansion_thesis.git] / General.Theory / DFT / DFT.Introduction.tex
blob949bdbc3d6f82c6b48099c9c6b6c41d6c2e44ef4
1 \section{Quantum mechanics\cite{Adamson}}
3 It can be said that most of the properties of a chemical compound is
4 determined by its electronic structure, which can be described
5 by the wavefunction, $\Psi$, of the
6 system. Therefore access to the detailed electronic structure should
7 enable prediction or evaluation of the behaviour of the system under
8 investigation.
10 Energy is the fundamental quantity of a system. Despite this fact,
11 molecular systems are usually described with other features, for
12 example the
13 molecular structure or the dipole moment. These
14 properties are determined by external parameters, which may become
15 variables on which a potential energy surface can be mapped.
16 Therefore, analytical derivatives of the energy with respect to these
17 variables yield the molecular properties.
18 For example, the derivatives of the energy with respect to the nuclear
19 positions give the nuclear forces, which allows minimisation of the
20 energy with respect to nuclear coordinates, providing the molecular
21 structure. The energy is thus the basis for the calculation of all
22 molecular properties, and a method for evaluating this property is of
23 fundamental interest.
25 Computational methods lead in different ways to the energy and
26 electronic structure of the investigated compound. This introduction
27 will describe Hartree-Fock methods and then move to Density Functional
28 Theory (DFT), the method of choice in all calculations of this project.
30 \subsection{The Schrödinger equation; Born interpretation}
32 The Schrödinger equation\cite{Schroedinger} is able to describe all non-relativistic chemical
33 phenomena. The time-dependent Schrödinger equation is
35 \hat{H}\Psi=-i\hbar \frac{\partial \Psi}{\partial t},
36 \label{schroedinger-time}
38 where $\Psi$ is the wavefunction of the molecular system and $\hat{H}$ is
39 the Hamiltonian operator, which describes the
40 energy of the system with the wavefunction. $\Psi$ is a
41 function of the nuclear and electron positions, the electron spins and
42 time. Positions $(\mathbf{r})$\footnote{Bold identifiers are
43 vectors.} and spins $(s)$ are often combined in
44 space-spin coordinates $(\mathbf{x})$. A system can be fully described with the wave function.
46 There are some different interpretations of $\Psi$,
47 but the most common is due to Born\cite{Born}, who stated that:
49 \begin{quote}
50 The probability of finding a system with the
51 wavefunction $\Psi$ in a given state is proportional to $\Psi^*
52 \cdot \Psi$. %\footnote{Wavefunctions with a superscript asterix are
53 %complex conjugate.}
54 \end{quote}
57 \subsection{Wavefunction properties}
59 The wavefunction is subject to some constraints.
61 As a result of the Born interpretation the integral of the system's exact wavefunction
62 over all space-spin
63 coordinates must equal one, because the
64 electron must be located somewhere. This is:
66 \int \Psi^* \cdot \Psi \, \mathrm{d}\tau = 1.
67 \label{normalized}
69 Equation (\ref{normalized}) is called the {\em normalisation} criteria.
71 Wavefunctions are considered to be orthogonal, the {\em
72 orthogonality} criteria. That means:
74 \int \Psi^*_{m} \cdot \Psi_{n}\, \mathrm{d}\tau = 0 \Leftrightarrow m
75 \not= n.
76 \label{orthogonal}
79 Both properties are often summed up under {\em orthonormality} and
80 expressed with the Kronecker delta:
82 \int \Psi^*_{m} \cdot \Psi_{n} \, \mathrm{d}\tau = \delta_{mn} =
83 \left\{ \begin{array}{lcr}
84 0 & \Leftrightarrow & m \not= n \\
85 1 & \Leftrightarrow & m = n
86 \end{array} \right.
87 \label{Kronnecker}
90 Because electrons are fermions, the wavefunction must be
91 {\em antisymmetric} with respect to exchange of two electrons, i. e.
92 the wavefunction has to change sign if two electrons are
93 exchanged. This is the Pauli Exclusion Principle.\cite{Pauli}
95 For most problems, a separation of the wave function into a
96 time-dependent part and a time-independent part is possible. The
97 wavefunction is then of the form:
99 \Psi(t)=\Psi \cdot \mathrm{e}^{\left[\frac{-iEt}{\hbar}\right]}.
101 Considering only the time independent part, the Schrödinger equation (\ref{schroedinger-time})
102 reduces to
104 \hat{H}\Psi = E \Psi,
105 \label{schroedinger-time-independent}
107 an eigenvalue equation, where $E$, the energy, is the eigenvalue for
108 the operator $\hat{H}$ and
109 $\Psi$ is the eigenfunction.
111 \subsection{The expectation value}
113 In general, considering $\hat{A}$ as a time-independent operator for a
114 physical quantity and $A$ the resulting eigenvalue, the equation
115 (\ref{schroedinger-time-independent}) changes to
117 \hat{A}\Psi = A \Psi.
118 \label{equation1}
120 Multiplying equation (\ref{equation1}) with the complex conjugate of
121 $\Psi$, which is $\Psi^*$, and integrating over all coordinates leads
124 \int \Psi^* \hat{A} \Psi \, \mathrm{d}\tau = \int \Psi^* A \Psi \, \mathrm{d}\tau.
126 $A$ is a constant scalar and so can be taken outside the integral
128 \int \Psi^* \hat{A} \Psi \, \mathrm{d}\tau = A\int \Psi^* \Psi \, \mathrm{d}\tau.
130 Dividing by $\int \Psi^* \Psi \, \mathrm{d}\tau$ gives the {\em
131 expectation
132 value} of the physical quantity (which can be considered to be the
133 average value of the physical quantity connected with $\hat{A}$ that
134 can be observed):
136 A=\frac{\int \Psi^* \hat{A} \Psi \, \mathrm{d}\tau}{\int \Psi^* \Psi
137 \, \mathrm{d}\tau}.
138 \label{expecting-value}
140 For normalised wavefunctions, $\int \Psi^* \Psi \, \mathrm{d}\tau=1$,
141 and so equation (\ref{expecting-value}) simplifies to $A=\int \Psi^* \hat{A} \Psi \, \mathrm{d}\tau$ or
142 written in {\em Bra-Ket} notation;
144 A= \langle \Psi | \hat{A} | \Psi \rangle =\int \Psi^* \hat{A} \Psi \, \mathrm{d}\tau
146 One can say $A$ is a {\em functional} of $\Psi$, written as
147 $A[\Psi]$, where $\Psi$ is itself a function of space-spin coordinates.
150 \subsection{The Hamiltonian operator}
152 If we consider a system of $M$ nuclei of charge $Z_{A}$ and $N$
153 electrons without an external field, the Hamiltonian, in atomic
154 units, is given by
156 \hat{H}=\hat{T}+\hat{V}
158 where,
160 \hat{T}= \underbrace{-\frac{1}{2} \sum\limits_{A}^{M}
161 \nabla^2_{A}}_{\mbox{\tiny{for the nuclei}}} \underbrace{ -
162 \frac{1}{2} \sum\limits_{i}^{N} \nabla^2_{i}}_{\mbox{\tiny{for the
163 electrons}}}
165 for the kinetic energy of nuclei and electrons, and
167 \begin{array}{lll}
168 \hat{V}& = & \underbrace{-\sum\limits_{A}^{M} \sum\limits_{i}^{N}
169 \frac{Z_{A}}{\left|\mathbf{R}_{A}-\mathbf{r}_{i}\right|}}_{\mbox{\tiny{nuclear-electron (ne) attraction}}}\\
170 &&\\
171 & & \underbrace{+\sum\limits_{A}^{M} \sum\limits_{B>A}^{M}
172 \frac{Z_{A}Z_{B}}{\left|\mathbf{R}_{A}-\mathbf{R}_{B}\right|}}_{\mbox{\tiny{nuclear-nuclear (nn) repulsion}}}\\
173 &&\\
174 & & \underbrace{-\sum\limits_{i}^{N} \sum\limits_{j>i}^{N}
175 \frac{1}{\left|\mathbf{r}_{i}-\mathbf{r}_{j}\right|}}_{\mbox{\tiny{electron-electron (ee) repulsion}}}
176 \end{array}
178 for the potential energy of attractive and repulsive interactions of
179 nuclei and electrons.
181 Thus the total energy of the system, $E_{\mbox{\tiny{tot}}}$, is
182 the sum of all partial energies.
184 \subsection{Born-Oppenheimer approximation}
185 It is common to treat quantum mechanical problems from the point of
186 view of the Born-Oppenheimer approximation.\cite{Born-Oppenheimer}
188 The masses of the nuclei are much greater than the masses of the
189 electrons, hence the electrons can respond almost instantaneously to
190 any change in the nuclear positions. Thus, to a good approximation, we
191 can think of the electrons moving in a field of fixed nuclei. As a
192 consequence, the terms introducing the nuclear kinetic energy can neglected and
193 the nuclear-nuclear repulsion can then be considered as constant. Removing these terms leads to the electronic Hamiltonian,
194 the electronic wavefunction and the electronic Schrödinger equation.
195 Only these will be considered in the following treatment.
197 The resulting energy will be the electronic energy. In order to regain the
198 total energy, the nuclear-nuclear repulsion terms must be added.
200 This reduces the Hamilatonian to:
202 \hat{H}=-\frac{1}{2} \sum\limits_{i}^{N} \nabla^2_{i}-\sum\limits_{A}^{M} \sum\limits_{i}^{N}
203 \frac{Z_{A}}{\left|\mathbf{R}_{A}-\mathbf{r}_{i}\right|} + -\sum\limits_{i}^{N} \sum\limits_{j>i}^{N}
204 \frac{1}{\left|\mathbf{r}_{i}-\mathbf{r}_{j}\right|}
206 Introducing the external potential $v(\mathbf{r}_{i})$
208 v(\mathbf{r}_{i})= - \sum\limits_{A}^{N}\frac{Z_{A}}{\left|\mathbf{R}_{A}-\mathbf{r}_{i}\right|}
210 and the shorthand form
212 r_{ij}= \sum\limits_{i}^{N}\left|\mathbf{r}_{i}-\mathbf{r}_{j}\right|
214 follows
216 \hat{H}={-\frac{1}{2} \sum\limits_{A}^{M}\nabla^2_{A}} +
217 \sum\limits_{i}^{N} v(\mathbf{r}_{i}) +
218 \sum\limits_{j>i}^{N}\frac{1}{r_{ij}}.
219 \label{Hamiltonian-simplyfied}
223 \subsection{The Variation Method}
225 As stated above, the expectation value of an operator is given
226 by equation (\ref{expecting-value}) and the operator for the molecular
227 energy is the Hamiltonian, $\hat{H}$. Consider an approximate
228 wavefunction $\Psi_{\mbox{\tiny{app}}}$ and the associated expectation
229 value of the energy
230 $E_{\mbox{\tiny{app}}}$, which can be obtained with
232 E_{\mbox{\tiny{app}}}=\frac{\int \Psi^*_{\mbox{\tiny{app}}} \hat{H}
233 \Psi^{}_{\mbox{\tiny{app}}}\, \mathrm{d}\tau}
234 {\int \Psi^*_{\mbox{\tiny{app}}} \Psi^{}_{\mbox{\tiny{app}}}\, \mathrm{d}\tau}.
236 The varation theorem states that the energy calculated from a trial $\Psi_{\mbox{\tiny{app}}}$ is an upper bond to the
237 true ground state energy. Minimisation
238 with respect to all allowed $N$-electron wavefunctions will give the
239 true ground state energy and the exact wavefunction, which is minimization of the functional
240 $E_{\mbox{\tiny{app}}}[\Psi_{\mbox{\tiny{app}}}]$, known as
241 the {\em Rayleigh-Ritz} minimal principle.
242 Therefore
244 E_{\mbox{\tiny{app}}} = E_{0} \Leftrightarrow
245 \Psi_{\mbox{\tiny{app}}} = \Psi_{0} \qquad .
247 This minimisation with respect to all allowed $\Psi$s is not practical. What is usually
248 done is to develop the trial
249 wavefunction as a linear combination of well-defined wavefunctions of
250 component parts of the system.
252 If $\Psi_{\mbox{\tiny{app}}}$ is developed
253 as a linear combination of wavefunctions,
255 \Psi_{\mbox{\tiny{app, i}}}=\sum\limits_{i} c_{i}\Psi_{i},
257 $E_{i}$ is the energy for the $i$th eigenstate of $\hat{H}$ resulting in the combination
258 of the $i$ wavefunctions. As stated above, every trial wavefunction yields an energy
259 greater than or equal the true ground state energy. As a resuslt, the process of energy minimisation
260 is reduced to a process of finding an optimum set of coefficients with this ansatz. This
261 task can be solved via matrix diagonalisation.
263 %$E_{0}\ge E_{1} \ge E_{2} \ge \cdots E[\Psi]$, is always greater than
264 %or equal to $E_{0}$. Thus the
265 %problem is reduced, one has to find the optimum sets of coefficients, $c_{i}$, which
266 %can be achieved via matrix diagonalisation.
270 % this last paragraph has to be reviewed once again due to some
271 % unclearness in its meaning and purpose
274 \subsection{Molecular wavefunction, atomic orbitals and basis sets}
276 The wavefunction $\Psi$, which is used here, is the molecular
277 wavefuntion, describing the complete electronic structure of a given
278 molecule. For a molecular wavefunction $\Psi_{\mbox{\tiny{app}}}$, the
279 most obvious choice of component wavefunctions,
280 $\Psi_{\mbox{\tiny{i}}}$, are those of the atoms that make up the
281 molecule.
283 As a consequence, the molecular wavefunction, $\Psi$, can be conceived as a combination
284 of one-electron wavefunctions, $\chi_{i}(\mathbf{x})$, which may be
285 regarded as atomic orbitals. This is called the molecular orbital
286 approximation.
288 The $\chi_{i}$ consists of a spatial
289 orbital $\psi_{i}(\mathbf{r})$ and one of the two spin functions
290 $\alpha(s)$, spin-up, or $\beta(s)$, spin-down, thus we get
291 $\chi_{i}(\mathbf{x})=\psi_{i}(\mathbf{r})\alpha(s_{i})$ or
292 $\chi_{i}(\mathbf{x})=\psi_{i}(\mathbf{r})\beta(s_{i})$.
294 %The orbital approximation states that you can consider all electrons
295 %as occupying 1-electron orbitals similar to those of
297 %in this paragraph we have to introduce the concepts of:
298 %- one electron molecular wave funtions , so calles orbitals,
299 %- separation of these functions into on spatial and one spin
300 %dependent part
301 %- introduce the concept of MO theory..
302 %.. which in my opinion has nothing to do with atomic wave functions,
303 %depending of the atoms building up the molecule!!
306 \subsubsection{Basis sets\cite{Computational-Chemistry}}
308 Because the exact mathematical description of the spatial orbitals
309 are even almost unknown for more complex molecules than hydrogen, they are usually developed in a set of
310 known functions, the {\em basis set}.
312 \psi_{i}=\sum\limits_{\mu}c_{\mu i}\phi_{\mu}
314 This is not an approximation as long as the basis set is complete, i. e.
315 it contains an infinite number of functions. This is not possible in
316 practice, so the
317 introduction of the basis set necessarily introduces some error.
319 The most convenient way to define a basis set for any nuclear
320 configuration is to define a particular set of functions for each
321 nucleus, depending only its charge.
323 In principle, any kind of function, or combination of functions can
324 be used. Their behaviour should agree with the physics of the
325 problem, e. g.
326 \begin{itemize}
327 \item{For bound atomic and molecular systems the function should
328 tend towards zero as the distance between nuclei and electron
329 becomes large.}
330 \item{The energy should converge reasonably rapidly as more basis
331 functions are added.}
332 \end{itemize}
334 The best compromise between the flexibility of the basis set, which
335 increases with the number of basis functions, and the computational
336 cost has to be made. The basis set should be kept as small as possible
337 to minimize the computational cost.
339 There are two main types in use today. The Slater-Type Atomic Orbitals (STOs) introduced in
340 1930 by Slater\cite{Slater-Orbitals} and the Gaussian-Type Atomic
341 Orbitals (GTOs), introduced in 1950 by Boys\cite{Gaussian-Orbitals}.
343 \paragraph{STOs} are of the form
345 \phi_{a}(\mathbf{r})=(x-A_{x})^{a_{x}}(y-A_{y})^{a_{y}}(z-A_{z})^{a_{z}}\cdot \mathrm{e}^{-\alpha |\mathbf{r}-\mathbf{A}|}
347 with a center $\mathbf{A}(A_{x},A_{y},A_{z})$, angular momentum
348 $\mathbf{a}(a_{x},a_{y},a_{z})$, a nuclei dependent exponent $\alpha$
349 and exponential radial parts, which takes care of the decay for large
350 distances. Like the real wavefunctions they have cusps at the nuclei.
351 Their major disadvantage is that integrals over STOs are expensive to
352 compute.
354 To overcome this difficulty, Boys suggested, in 1950, the construction
355 of basis functions with Gaussian-Type Atomic
356 Orbitals, GTOs.
358 \paragraph{GTOs} are of the form
360 \phi_{a}(\mathbf{r})=(x-A_{x})^{a_{x}}(y-A_{y})^{a_{y}}(z-A_{z})^{a_{z}}\cdot \mathrm{e}^{-\alpha |\mathbf{r}-\mathbf{A}|^2}.
362 GTOs decay too fast and have incorrect nuclear cusps, but the speed
363 with which integrals over GTOs can be computed compensates for this.
364 As a result, if the same accuracy as with STOs is required, they can be
365 approximated by a sum of Gaussian-Type orbitals. This leads to the
366 STO-$n$G basis sets, where one STO is represented by so-called
367 {\em contraction} of $n$ Gaussian Type Orbitals. A contraction is a linear
368 combination of Gaussian Orbitals where the coefficients are not
369 changed during the calculation.
371 Having specified the type of functions to be used and the locations,
372 it must be decided what number of functions are to be used. The
373 smallest number of functions possible is a {\em minimal basis set},
374 where only one function for each electron of the neutral atoms are
375 used. Improvements in accuracy and flexibility can be achieved through
376 using double the amount of basis functions needed, a
377 {\em Double Zeta} basis set, indicated by double-$\zeta$. The next
378 step is the use of three times more functions than the number of
379 electrons, triple-$\zeta$. A variation of the basis set which only increases
380 the number of valence orbitals produces a {\em split valence basis set}.
382 This is reasonable, because core electrons are not significantly
383 influenced by the bonding environment, especially for atoms with a
384 large number of electrons. Functions with a higher angular momentum
385 are denoted {\em polarisation functions}.
386 %Usually the next orbital function of the last occupied valence orbital
387 %is used to improve the
388 %description of the electronic distribution in a certain direction, e.g. p-functions to polarize s-functions,
389 %f-functions to polarize d-functions, etc.
391 If basis sets of the form {\em k-nlm}G{\em w} are referred to, then some
392 special kinds of split-valence basis sets, designed by {\em Pople et.
393 al}\cite{Pople-basis-sets1, Pople-basis-sets2, Pople-basis-sets3,
394 Pople-basis-sets4} have been used. The {\em k} indicates how many
395 primitive GTOs have been contracted to represent the core orbitals,
396 the number of values after the dash indicates the grade of splitting
397 of the valence orbitals, {\em nlm} is thus a triple split valence
398 basis set. Thus the valence basis is represented with three functions,
399 each of which is constructed from the number of Gaussian Type Orbitals given.
400 Values after the G indicates polarisation functions.
404 \subsection{The Hartree approximation, Hartree-Fock Theory, Slater
405 Determinant}
407 If you want to construct a molecular wavefunction built up from
408 one-electron wavefuntions a very simple approach is to construct your
409 molecular wavefuntion as a combination of {\em independent}
410 one-electron wavefunctions,
412 The simplest way to construct a molecular wavefunction of one-electron
413 wavefunction is the independent-particle, or Hartree-approximation,
414 where the total wavefunction is built up as a product of orthonormal
415 one-electron wavefunctions\cite{Hartree1, Hartree2, Hartree3}. This
416 approximation assumes that each
417 electron moves independently in its own orbital, only influenced by
418 the {\em average} field generated by all the other electrons and explicit electron
419 interaction is neglected.
421 The
422 Hartree wavefunction for a system of $N$ electrons is
424 \Psi = \chi_{1}(\mathbf{x}_{1})\chi_{2}(\mathbf{x}_{2}) \ldots
425 \chi_{N-1}(\mathbf{x}_{N-1})\chi_{N}(\mathbf{x}_{N}).
428 This wavefunction is an invalid one, because it is not antisymmetric with
429 respect to electron interchange.
431 This was first mentioned by Fock in
432 1930\cite{Fock} and one solution he suggested
433 was to add and substract all possible permutations of the Hartree
434 product wavefunction, forming the Hartree-Fock (HF) wavefunction.
435 Slater found later that the result is the determinant of a matrix,
436 called the Slater determinant\cite{Slater1, Slater2}
438 \Psi_{\mbox{\tiny{SD}}} = \frac{1}{\sqrt{N!}} \left|
439 \begin{array}{cccc}
440 \chi_{1}(\mathbf{x}_{1}) & \chi_{2}(\mathbf{x}_{1}) & \cdots &
441 \chi_{N}(\mathbf{x}_{1}) \\
442 \chi_{1}(\mathbf{x}_{2}) & \chi_{2}(\mathbf{x}_{2}) & \cdots &
443 \chi_{N}(\mathbf{x}_{2})\\
444 \vdots & \vdots & & \vdots \\
445 \chi_{1}(\mathbf{x}_{2})& \chi_{1}(\mathbf{x}_{N})& \cdots &
446 \chi_{N}(\mathbf{x}_{N})
447 \end{array} \right|.
449 The prefactor $1/\sqrt{N!}$ normalizes the HF wavefunction.
451 Usually the trial wavefunction is considered as
452 a {\em single} Slater determinant, which implies that
453 electron-interaction is neglected (the electron-electron repulsion is
454 only included as an average effect). If additional approximations are
455 made, this leads to {\em semi-empirical} methods. If additional
456 determinants are included, this leads to more exact {\em post-HF} methods.
458 \subsubsection{Hartree-Fock equations}
460 Hartree-Fock calculations start with an initial guess of the
461 wavefunction and then successively improve this wavefunction with
462 respect to the energy following the variational principle and
463 optimizing the set of coefficients.
466 Considering the Born-Oppenheimer approximation the Hamiltonian is
467 given as
469 \hat{H}=\hat{T}_{\mbox{\tiny{e}}} + \hat{V}_{\mbox{\tiny{ne}}} +
470 \hat{V}_{\mbox{\tiny{ee}}}\ ,
472 grouping according to the number of electron indices (e) leads to
473 \begin{eqnarray}
474 \hat{h}_{i}&=& \hat{T}_{\mbox{\tiny{e}}}+\hat{V}_{\mbox{\tiny{ne}}} \nonumber \\
475 \hat{g}_{ij}&=& \hat{V}_{\mbox{\tiny{ee}}}\ .
476 \end{eqnarray}
478 The one-electron operator $\hat{h}_{i}$ describes the motion of
479 electron $i$ in the field of all nuclei and the electron-electron
480 repulsion is described with the two-electron operator $\hat{g}_{ij}$.
482 To evaluate the energy of a single Slater determinant, we apply these
483 operators to the determinant. For each of the operators we only get
484 a non-zero value if there is no contribution of overlap integrals, i. e.
485 for the one-electron operator, if both one electron wavefunctions of
486 the determinant are of the same electron. That yields for the
487 expectation value
488 \begin{eqnarray}
489 & {} & \bra \chi_{i}(\mathbf{x}_{i}) \chi_{i+1}(\mathbf{x}_{i+1}) \ldots
490 \chi_{N}(\mathbf{x}_{N}) |\hat{h}_{i}| \chi_{i}(\mathbf{x}_{i})
491 \chi_{i+1}(\mathbf{x}_{i+1}) \ldots
492 \chi_{N}(\mathbf{x}_{N}) \ket \nonumber \\
493 & = & \bra \chi_{i}(\mathbf{x}_{i})|\hat{h}_{i}|\chi_{i}(\mathbf{x}_{i}) \ket
494 \bra \chi_{i+1}(\mathbf{x}_{i+1})|\chi_{i+1}(\mathbf{x}_{i+1}) \ket \ldots
495 \bra \chi_{N}(\mathbf{x}_{N})|\chi_{N}(\mathbf{x}_{N}) \ket \\
496 & = & \sum\limits_{i}^{N} h_{i} \ . \nonumber
497 \end{eqnarray}
498 For the two electron operator, the value is non-zero only if
499 the same orbital combinations contribute and every possible permutation
500 in which only two orbitals contribute. For
501 the first case, if identical orbitals contribute, it leads to
502 \begin{eqnarray}
503 & {} & \bra \chi_{i}(\mathbf{x}_{i}) \bra \chi_{j}(\mathbf{x}_{j}) \chi_{i+1}(\mathbf{x}_{i+1}) \ldots
504 \chi_{N}(\mathbf{x}_{N}) |\hat{g}_{ij}| \chi_{i}(\mathbf{x}_{i})
505 \chi_{j}(\mathbf{x}_{j})
506 \chi_{i+1}(\mathbf{x}_{i+1}) \ldots
507 \chi_{N}(\mathbf{x}_{N}) \ket \nonumber \\
508 & = & \bra \chi_{i}(\mathbf{x}_{i}) \bra
509 \chi_{j}(\mathbf{x}_{j})|\hat{g}_{ij}|\chi_{i}(\mathbf{x}_{i}) \bra \chi_{j}(\mathbf{x}_{j})\ket
510 \bra \chi_{i+1}(\mathbf{x}_{i+1})|\chi_{i+1}(\mathbf{x}_{i+1}) \ket \ldots
511 \bra \chi_{N}(\mathbf{x}_{N})|\chi_{N}(\mathbf{x}_{N}) \ket \nonumber \\
512 & = & \sum\limits_{i}^{N} \sum\limits_{j> i}^{N} J_{ij}\ .
513 \end{eqnarray}
514 This is called {\em Coulomb} integral. It represents a classical
515 repulsion between two charge distributions described by the $i$th and
516 $j$th wavefunction.
518 For the second case the integral is
519 \begin{eqnarray}
520 & {} & \bra \chi_{i}(\mathbf{x}_{i}) \bra \chi_{j}(\mathbf{x}_{j}) \chi_{i+1}(\mathbf{x}_{i+1}) \ldots
521 \chi_{N}(\mathbf{x}_{N}) |\hat{g}_{ij}| \chi_{j}(\mathbf{x}_{j})
522 \chi_{i}(\mathbf{x}_{i})
523 \chi_{i+1}(\mathbf{x}_{i+1}) \ldots
524 \chi_{N}(\mathbf{x}_{N}) \ket \nonumber \\
525 & = & \bra \chi_{i}(\mathbf{x}_{i}) \bra
526 \chi_{j}(\mathbf{x}_{j})|\hat{g}_{ij}|\chi_{j}(\mathbf{x}_{j}) \bra
527 \chi_{i}(\mathbf{x}_{i})\ket
528 \bra \chi_{i+1}(\mathbf{x}_{i+1})|\chi_{i+1}(\mathbf{x}_{i+1}) \ket \ldots
529 \bra \chi_{N}(\mathbf{x}_{N})|\chi_{N}(\mathbf{x}_{N}) \ket \nonumber \\
530 & = & \sum\limits_{i}^{N} \sum\limits_{j> i}^{N} K_{ij} \ .
531 \end{eqnarray}
532 It is termed {\em exchange} integral and has no classical analogue.
534 The energy of a single Slater determinant can then be written as
535 \begin{eqnarray}
536 E_{\mbox{\tiny{SD}}} &=& \sum\limits_{i}^{N} h_{i} +
537 \sum\limits_{i}^{N} \sum\limits_{j> i}^{N} J_{ij} -
538 \sum\limits_{i}^{N} \sum\limits_{j> i}^{N} K_{ij} \nonumber \\
539 &=& \sum\limits_{i}^{N} h_{i} + \sum\limits_{i}^{N} \sum\limits_{j\ge
540 i}^{N} \left( J_{ij} - K_{ij} \right) \\
541 &=& \sum\limits_{i}^{N} h_{i} + \frac{1}{2}\sum\limits_{i}^{N}
542 \sum\limits_{j}^{N} \left( J_{ij} - K_{ij} \right) \ . \nonumber
543 \end{eqnarray}
545 One remarkable point is that for $i=j$, the Coulomb operator
546 generates a spurious self-interaction of the electron, which is exactly
547 compensated by the coresponding exchange operator, because
548 $J_{ii}=K_{ii}$.
550 For Hartree-Fock calculation, the energy is expressed in terms of operators, the identifiers are the
551 same as for the integrals. Minimizing the energy with respect to the
552 orbitals is then carried out following the variation principle. This
553 must be carried out in a way that the orbitals remain orthonormal. This
554 constraint optimization is usually done self-consistently with the method of {\em
555 Lagrange} multipliers, as small changes in the orbitals should not
556 change the Lagrange function. The {\em Fock} operator is then
557 defined, which is an effective one-electron operator and associated
558 with the variation of the energy. The result is a set of orbitals satisfying the {\em
559 Hartree-Fock equations}. A normal Hartree-Fock
560 calculation starts with a guess of the set of orbitals, and then
561 iteratively minimises the energy with respect to the orbitals.
562 Obviously a good starting guess is necessary to gain a fast convergence.
564 The principle of {\em self-consistent} calculations is that based on
565 the results of the first set, a second set is generated, then a third
566 one based on the results of the second one and so on. This is carried
567 on until the energy variation is beneath a predefined value and then
568 is called {\em self-consistent} or stationary. Subsequent iterations would give a
569 variation of the energy under this threshold. The {\em Hartree-Fock}
570 limit is the lowest energy can be reached
571 with HF calculations in the limit of the chosen basis set.
573 Explicit
574 electron-electron interaction is neglected. Electrons of the same
575 spin are partially but not completely, correlated. No attemp is made
576 to include electron-correlation of opposite-spin electrons.
578 Because of this neglect, the energy of HF calculations,
579 $E_{\mbox{\tiny{HF}}}$ is always above the exact non-relativistic
580 value, $E_{\mbox{\tiny{exact}}}$, the difference is defined as the
581 {\em correlation energy},
582 $E_{\mbox{\tiny{corr}}}=E_{\mbox{\tiny{exact}}}-E_{\mbox{\tiny{HF}}}$.\cite{Loewdin}
583 This is the main deficiency of HF theory.
585 \subsubsection{Post-HF methods}
587 The major aim of post-HF methods is to include electron correlation.
588 There are different approaches to describe the interactions of the
589 electrons correctly,
590 \begin{itemize}
591 \item{via Configuration Interaction (CI), in which a molecular
592 wavefunction is constructed as a linear combination of wavefunctions
593 of the HF determinant and determinants corresponiding to
594 excitation of electron.
595 If the expansion of the molecular wavefunction as a linear
596 combination is complete, every possible excited state is
597 included, the exact correlation energy within the
598 basis set approximation is achieved.}
600 % \item{via Coupled Cluster (CC) Theory, in which the coefficients
601 % of the additional
602 % wavefunctions are constructed of by the exponentiated operator of
603 % the excitations. In
604 % this way, higher excitations are only partially included and their
605 % coefficients are determined by lower order excitations. This
606 % reveals typically $95\%$ of the correlation energy.}
607 \item{via M{\o}ller-Plesset Pertubation Theory, in which the
608 exact Hamiltonian is treated as a {\em small} perturbation of the
609 HF-Hamiltonian.}
610 \end{itemize}
612 All these approaches are based on wavefunction-expansions of the many
613 electron wavefunction $\Psi(\mathbf{x}_{1},\ldots,\mathbf{x}_{N})$. This is
614 combined with an increase in the number of integrals to solve
615 according to the $3N$-coordinate dependency of the wavefunction. If
616 spin is included, this becomes $4N$.
619 \section{Density Functional Theory\cite{Parr-Yang}}
621 \subsection{The electron density}
623 For a given state with wave function $\Psi$ the correlated electron density
624 is given by
626 \rho(\mathbf{r})=N \int \Psi^{*} \cdot \Psi \, \mathrm{d}s_{1}
627 \mathrm{d}\mathbf{x}_{2}\mathrm{d}\mathbf{x}_{3}\ldots\mathrm{d}\mathbf{x}_{N}
628 \label{electron-density}
630 which is the square sum over all occupied orbitals
632 \rho(\mathbf{r})= \sum\limits_{i}^{\mbox{\tiny{occ.}}}
633 \psi_{i}(\mathbf{r})\cdot \psi_{i}^{*}(\mathbf{r})
635 and it is
637 \int \rho(\mathbf{r}) \, \mathrm{d}\mathbf{r} = N \ .
640 If the exact electron density is known, then the cusps in
641 $\rho(\mathbf{r})$ will provide the position of the nuclei. The slope
642 of $\rho(\mathbf{r})$ at the nucleus $A$ must obey
644 \left. \frac{\partial}{\partial
645 \mathbf{r}_{A}}\bar{\rho}(\mathbf{r}_{A}) \right|_{\mathbf{r}_{A}} =
646 -2Z_{A}\cdot \bar{\rho}(0)
648 giving the charge at the nucleus, $Z_{A}$ ($\bar{\rho}$ denotes the
649 spherical average of the density). As the Hamiltonian is defined
650 completely by the nuclear charges and positions, it is fully known.
651 Therefore the wavefunction and energy can be found and the system can be
652 completely described by the electron density. A Hamiltonian like
653 equation (\ref{Hamiltonian-simplyfied}) is completely determined by
654 the external potential $v(\mathrm{r})$.
657 \subsection{The first Hohenberg-Kohn theorem\cite{HK1}}
658 The theoream says:
660 For non-degenerate ground-states
661 \begin{quote}
662 the external potential $v(\mathrm{r})$ is
663 determined by the electron density, $\rho(\mathbf{r})$.
664 \end{quote}
666 More precisely, there is a one-to-one correspondence between the electron
667 density of a system and its energy and vice versa. This theorem can be proven with the
668 variational principle.\cite{Kohn-Rev.Mod.Phys.-1999}
670 %Consider a system with the potential $v_{a}(\mathbf{r})$ and the corresponding
671 %electron density $\rho$. Consider another system with the same
672 %electron density $\rho$ and a different potential $v_{b}(\mathbf{r})$.
673 %A Hamiltonian $\hat{H}_{a}$ and a wavefunction
674 %$\Psi_{a}$ are connected
675 %with potential $a$. Because potential $b$ is different from potential $a$, there
676 %must be a different wavefunction $\Psi_{b}$ with the Hamiltonian
677 %$\hat{H}_{b}$. Using the variational principle,
678 %\begin{eqnarray}
679 %E_{a}^0 < \bra \Psi_{b}|\hat{H}_{a}|\Psi_{b} \ket & = & \bra
680 %\Psi_{b}|\hat{H}_{b}|\Psi_{b} \ket + \bra
681 %\Psi_{b}|\hat{H}_{a}-\hat{H}_{b}|\Psi_{b} \ket \nonumber \\
682 %&=& E_{b}^0 + \int \rho(\mathbf{r})
683 %[v_{a}(\mathbf{r})-v_{b}(\mathbf{r})] \, \mathrm{d}\mathbf{r}.
684 %\end{eqnarray}
685 %Similarly
687 %E_{b}^0= E_{a}^0 + \int \rho(\mathbf{r})
688 %[v_{b}(\mathbf{r})-v_{a}(\mathbf{r})] \, \mathrm{d}\mathbf{r},
690 %which leads to the contradiction
692 %E_{a}^0 + E_{b}^0 < E_{b}^0 + E_{a}^0 \ .
695 Hence the external potential is determined by the density and we can
696 represent the energy as a functional of the density. In comparison
697 with the wave mechanics approach, the energy may be divided into
698 several parts: kinetic energy, $T[\rho]$ and attraction between the
699 nuclei and electrons, $V_{\mbox{\tiny{ne}}}[\rho]$.
701 The last may be written in
702 terms of the external potential and the electron density as
704 V_{\mbox{\tiny{ne}}}[\rho] = \int \rho(\mathbf{r})v(\mathbf{r})\ \mathrm{d}\mathbf{r}.
706 Further, the electron-electron repulsion -- divided into Coloumb part, $J[\rho]$,
707 and an exchange part, $K[\rho]$:
709 E[\rho]= T[\rho] + V_{\mbox{\tiny{ne}}}[\rho] + J[\rho] + K[\rho]\ .
712 The theorem has since been extended to non-degenerate ground
713 states.\cite{Dreizler}
715 \subsection{The second Hohenberg-Kohn theorem\cite{HK1}}
717 The second Hohenberg-Kohn theorem adopts the variational principle
718 into density functional theory.
719 \begin{quote}
720 For a trial density $\tilde{\rho}(\mathbf{r})$, such that
721 $\tilde{\rho}(\mathbf{r})\ge 0$ and $ \int
722 \tilde{\rho}(\mathbf{r}) \, \mathrm{d}\mathbf{r} = N$, an energy
723 $E[\tilde{\rho}] \ge E^0$ arises.
724 \end{quote}
725 This implies, an incorrect density will yield a energy which is higher than
726 the exact ground state energy for every trial density supplied by a
727 trial wavefunction.
730 %Assume a trial wavefunction $\Psi_{\rho_{0}}$, that integrates to the
731 %ground state densisty $\rho_{0}(\mathbf{r})$, and the real ground state
732 %wavefunction $\Psi_{0}$. The varational principle gives
734 %\bra \Psi_{\rho_{0}} | \hat{H} | \Psi_{\rho_{0}} \ket \ge \bra
735 %\Psi_{0} | \hat{H} | \Psi_{0} \ket = E^0 \ .
737 %Splitting the Hamiltonian into the usual parts and taking into account
738 %that the potential energy due to the external potential is a
739 %functional of the density leads to
740 %\begin{eqnarray}
741 %\bra \Psi_{\rho_{0}} | \hat{T} + \hat{J} + \hat{K} | \Psi_{\rho_{0}}
742 %\ket + \int \rho_{0}(\mathbf{r})v(\mathbf{r})\,
743 %\mathrm{d}\mathbf{r} &\ge& \bra \Psi_{\rho_{0}} | \hat{T} + \hat{J} + \hat{K} | \Psi_{\rho_{0}}
744 %\ket + \int \rho_{0}(\mathbf{r})v(\mathbf{r})\,
745 %\mathrm{d}\mathbf{r} \nonumber \\
746 %\bra \Psi_{\rho_{0}} | \hat{T} + \hat{J} + \hat{K} | \Psi_{\rho_{0}}
747 %\ket & \ge & \bra \Psi_{\rho_{0}} | \hat{T} + \hat{J} + \hat{K} | \Psi_{\rho_{0}}
748 %\ket \ .
749 %\end{eqnarray}
750 %Thus the $\Psi_{0}$ is the wavefunction that integrates to $\rho_{0}$
751 %and minimizes the expectation value of $\hat{T} + \hat{J} + \hat{K}$.
752 %Defining a universal functional as
754 %F[\rho]=\min_{\scriptscriptstyle{\Psi \to \rho}} \bra \Psi |
755 %\hat{T} + \hat{J} + \hat{K} | \Psi \ket \ ,
757 %which searches all $\Psi$'s that yields the input density $\rho$,
758 %allows the energy to be expressed as
759 %\begin{eqnarray}
760 %E^0 & = & \min_{\scriptscriptstyle{\rho}} \left[ F[\rho] + \int \rho(\mathbf{r})v(\mathbf{r})\,
761 %\mathrm{d}\mathbf{r} \right] \nonumber \\
762 % &=& \min_{\scriptscriptstyle{\rho}} \ E[\rho] \ ,
763 %\end{eqnarray}
764 %which is a search over all $N$-representable densities. $N$-representable
765 %means that the density must be obtainable from some antisymmetric
766 %wavefunction.
768 The task is now to find an accurate functional $E[\rho]$ and minimize it with respect to
769 the electron densitiy and the corresponding energy.
772 \subsection{The functionals\cite{Computational-Chemistry}}
774 The energy functional can be split according to the wave mechanics
775 approach into the usual parts
777 E[\rho]= T[\rho] + V_{\mbox{\tiny{ne}}}[\rho] + J[\rho] + K[\rho]\ .
779 The functionals for the electron-nucleus attraction and for the
780 Coulomb electron interaction can be expressed in their classical form
781 \begin{eqnarray}
782 V_{\mbox{\tiny{ne}}}[\rho]& = & \int \rho(\mathbf{r})v(\mathbf{r})\,
783 \mathrm{d}\mathbf{r} \\
784 J[\rho] & = & \frac{1}{2} \int \!\!\! \int
785 \frac{\rho(\mathbf{r})\rho(\mathbf{r}')}{|\mathbf{r}-\mathbf{r}'|}\,
786 \mathrm{d}\mathbf{r}\mathrm{d}\mathbf{r}'
787 \end{eqnarray}
789 There are no known ways to express the functionals for the
790 kinetic energy and the exchange energy classically.
794 \subsubsection{Thomas-Fermi-Dirac functionals}
796 An early approach to obtain approximate expressions for the unknown
797 functionals are by {\em Thomas}\cite{Thomas}, {\em Fermi}\cite{Fermi1, Fermi2,
798 Fermi3}, {\em Block}\cite{Block} and {\em Dirac}\cite{Dirac} considering
799 a non-interacting electron gas. For this model the functionals are
800 given as
801 \begin{eqnarray}
802 T[\rho] &=& \frac{3}{10}(3\pi^2)^{2/3} \int
803 \rho^{5/3}(\mathbf{r}) \, \mathrm{d}\mathbf{r} \\
804 K[\rho] &=& \frac{3}{4}\left( \frac{3}{\pi} \right)^{1/3} \int
805 \rho^{4/3}(\mathbf{r}) \, \mathrm{d}\mathbf{r}
806 \end{eqnarray}
808 The assumption of a non-interacting electron gas is not good for
809 atomic and molecular systems. It may be a usable model for solid state
810 calculations, and may give reasonable results there.
812 For molecular systems, for example, this model does not predict any
813 bonding, which is obviously not reasonable and can therefore not
814 compete with the results of wave mechanic approaches.
816 The main problem in Thomas-Fermi models is that the kinetic energy
817 part is poorly represented.
820 \subsubsection{Kohn-Sham Theory}
822 The foundation for the use of DFT in computational chemistry was the
823 introduction of orbitals by {\em Kohn} and {\em
824 Sham}.\cite{Kohn-Sham} With this introduction, the treatment of the main problem in
825 Thomas-Fermi Theory can be significantly improved. The basic idea is
826 to minimize the error in the kinetic energy term by splitting the
827 kinetic energy into two parts, one of which can be calculated exactly,
828 and a small correction term.
830 Consider a system of {\em non-interacting} electrons. For this system,
831 the kinetic energy is described exactly with the functional of the
832 Hartree-Fock-Slater theory, that is
833 \begin{eqnarray}
834 T_{\mbox{\tiny{S}}} & = & \sum \limits_{i}^N \bra \chi_{i} | \hat{T} |
835 \chi_{i} \ket \nonumber \\
836 & = & \sum \limits_{i}^N \bra \chi_{i} | -\frac{1}{2} \nabla^2 |
837 \chi_{i} \ket \ .
838 \end{eqnarray}
839 The electron density for this system is given exactly by
841 \rho(\mathbf{r})=\sum\limits_{i}^{N} \left| \chi_{i}(\mathbf{r})
842 \right|^{2}
844 and the total energy is given by
846 E[\rho]=T_{\mbox{\tiny{S}}}[\rho] + V_{\mbox{\tiny{ne}}}[\rho] +
847 J[\rho] \ .
849 In this way the major contribution of the kinetic enery is described exactly.
851 Kohn and Sham reformulated the interaction problem so that the
852 difference between the exact kinetic energy and the one calculated
853 for non-interacting orbitals is absorbed into an {\em
854 exchange-correlation} term, including also the exchange part $K[\rho]$.
856 In general, a DFT energy expression can be written as
858 E_{\mbox{\tiny{DFT}}}= T_{\mbox{\tiny{S}}}[\rho] +
859 V_{\mbox{\tiny{ne}}}[\rho] + J[\rho] + E_{\mbox{\tiny{xc}}}[\rho] \ .
860 \label{DFT-energy}
862 The exchange term,$E_{\mbox{\tiny{xc}}}[\rho]$ is defined by equation (\ref{DFT-energy}).
864 This exchange correlation energy can not completely be compared with
865 the terms of Hartree-Fock theory, because the definitions are different.
866 The major problem in this is the proper handling of long and short
867 range terms of the different functionals. But despite that, the problem
868 remains in principle the same. One has to iteratively find the set of
869 orbitals, yielding the density, which minimizes the energy to a
870 certain threshold.
872 Again the method of choice is the use of Lagrange
873 multipliers. The equations are arranged with a one-electron operator,
874 similar to the Fock-operator and the set of coefficients making the
875 eigenvalue equations stationary with respect to the energy must be
876 found. This yields the Kohn-Sham orbitals of the system, which will
877 describe the electron density correctly and, as a result, yield
878 the correct energy. The eigenvalue equations are known as {\em
879 Kohn-Sham equations}.
881 Because the KS-orbitals only have to decribe the correct density as
882 opposed to the HF-orbitals, which have to describe the complete
883 electronic properties, the
884 requirements for their accuracy are not as high as in Hartree-Fock,
885 which allows the use of lower order basis sets and therefore in this
886 treatment some computational cost can be saved.
888 The problem which still remains is to find the correct exchange functional.
889 There is no systematic way to find such a functional. If the exact
890 functional was known, DFT would provide the exact energy.
891 The major advance of DFT is, that the more exact description of the
892 electron behaviour is included in the Hamiltonian as a
893 correction term, as opposed to post-HF methods, where the
894 wavefunction is extended. That leads to a much simpler treatment, as the functional
895 can be defined as most practical and accurate for ones purposes.
897 As already stated, there is no systematic way for defining
898 correct functionals, but there are a number of well known
899 approximations
900 which yield results in the range of high order wave mechanics.
902 \subsubsection{Local Density Approximation (LDA)}
904 A famous and
905 very simple approximation is the {\em Local Density
906 Approximation}. It is assumed that the density locally can be
907 treated as that of an uniform electron gas, which is equivalent to
908 the view
909 that the density is a slowly varying function. Then the exchange
910 functional is defined by
912 E_{\mbox{\tiny{xc}}}^{\mbox{\tiny{LDA}}}[\rho]=\int
913 \epsilon_{\mbox{\tiny{xc}}}(\rho(\mathbf{r})) \, \rho(\mathbf{r}) \,
914 \mathrm{d}\mathbf{r} \ .
915 \label{LDA}
917 Here $\epsilon_{\mbox{\tiny{xc}}}(\rho(\mathbf{r}))$ is the
918 exchange-correlation energy of a uniform interacting electron gas of
919 density $\rho(\mathbf{r})$. This quantity is known to a very high
920 accuracy ($\approx 0.1\%$).\cite{Kohn-JPhysChem-1996}.
922 It is\cite{Kohn-Rev.Mod.Phys.-1999}
924 \epsilon_{\mbox{\tiny{x}}}(\rho(\mathbf{r}))=-\frac{0.458}{r_{\mbox{\tiny{s}}}} \ ,
926 the {\em exchange} part, given in atomic units, and
928 \epsilon_{\mbox{\tiny{c}}}(\rho(\mathbf{r}))=-\frac{0.44}{r_{\mbox{\tiny{s}}} +7.8} \ ,
930 the {\em correlation} part. $r_{\mbox{\tiny{s}}}$ is the sphere
931 containing one electron and given by
932 $r_{\mbox{\tiny{s}}}=\sqrt[3]{3/4\pi}$.
934 LDA leads to extremly useful results for most applications. Experience
936 shown\cite{Kohn-Rev.Mod.Phys.-1999}
937 that LDA gives ionisation energies of atoms and dissociation energies
938 of molecules with a fair accuracy, typically of $10$--$20\%$.
939 Bond lengths and thus the geometry of molecules have typically
940 accuracies of $\approx 1\%$ for main-group molecules.
941 The results are not so good for transition metals.
943 The solution of the Kohn-Sham equations in the LDA
944 is only minimally
945 more difficult than the the solution of the Hartree equation and very much
946 easier than the solution of the Hartree-Fock equations. The accuracy
947 for the exchange part is about $10\%$, while the much smaller correlation
948 part is generally overestimated with by factor of two. The two errors
949 typically cancel partially.
951 Based on the LDA more sophisticated exchange functionals have been
952 developed, for which accuracy is much higher. The principle is the
953 so-called {\em generalised gradient correction} (GGA), where a more
954 accurate behaviour of the electron density and its variation with
955 respect to the position is taken into account. Some recent and well used
956 functionals are {\em hybrid} functionals, where some degree of {\em
957 ``exact''} exchange energy is incorporated using traditional wave
958 mechanics.
960 \subsection{The method of choice}
962 DFT becomes more favourable with a greaer number of electrons in the system. For this reason DFT calculations are often
963 performed for systems including metal centres, or very large systems. For
964 the evaluation of molecular geometries, even the very simple LDA
965 obtains useful results. The more simple the calculations are, the faster
966 they are.
968 For all of these reasons, the calculations presented in this report
969 have been performed with DFT methods in the LDA.