• Keine Ergebnisse gefunden

2.2. Single-Particle Energies

2.2.2. Wannier Functions

In the following section we follow the review article on “Maximally localized Wannier functions” by Marzari et al. [23] and the underlying articles by Marzari and Vanderbilt [66] as well as Souza et al. [67].

Due to the translational invariance of solid-state materials density functional theory calculations are typically performed in reciprocal space. Thereby, the Kohn-Sham equation from Eq. (2.43) is solved within a delocalized Bloch basis and eigenenergies are obtained in terms of band energies εnpkq for a finite set of k-points. On the other side, lattice models like the Hubbard or the Anderson impurity models, which are regularly used to study many-body physics, are conventionally formulated in real space using a localized basis. Thus, to infer corresponding model parameters from the ab initio calculations in order to gain simplified yet material-realistic descriptions, we need a tool which is capable of transforming the delocalized reciprocal data to localized real-space quantities. For this purpose Wannier functions are utilized here.

Sophisticated Fourier-like transforms are used to transform the former Bloch states to Wannier functions which are in turn employed to evaluate the needed matrix elements in real space. This change of basis allows in addition to choose a sub-set or “low-energy”

space in order to minimize the basis elements in a well controlled manner.

In the context of single-particle properties we will derive tight-binding (TB) lattice models based onab initio calculations. These models will reproduce the original band structures accurately and can additionally be used to interpolate the former coarse k-grids to arbitrary points in reciprocal space. This “Wannier-interpolation” scheme will be used in this thesis to derive highly resolved band-structures from DFT and GW calculations and to describe Coulomb matrix elements in the whole Brillouin zone.

Therefore, the most import definitions and derivations concerning the construction of Wannier functions fromab initio Bloch states are given in the following.

Maximally Localized Wannier Functions

In Dirac’s braket notation the formal definition of Wannier functions in three dimen-sions is given by the Fourier transform

|Rny “ V p2πq3

ż

IBZ

d3k e´ikRnky, (2.61) which corresponds to the real-space function as wnRprq “ xr|Rny [68]. While the real-space representations of the Bloch functions Φnkprq “ xr|Φnky “ eikrunkprq are delocalized, the Wannier functions wnRprq are strongly localized around R. Unfortu-nately Eq. (2.61) does not uniquely define|Rny since the Bloch functions are subject to a certain gauge freedom concerning the overall phase. Thus, a transformation of the form

|Φ˜nky “enpkqnky (2.62) does not change any physical property (as long as φnpkq is real and periodic in k) but changes the resulting Bloch and Wannier wave functions. Therefore, further require-ments are needed to obtain a “well” defined Wannier basis. To this end, Marzari and Vanderbilt introduced a generalized gauge transformation

|Φ˜nky “ ÿJ

m“1

Umnkmky, (2.63)

which shall be chosen tominimize the real-space spread of the resulting Wannier func-tions by mixing a composite set of J bands [66, 23]. This can also be seen as a

“maximal” degree of localization of the Wannier functions in real space, leading to

“maximally localized Wannier functions” (MLFW), which directly translates to “max-imal” smoothness of the transformed Bloch states |Φ˜nky in reciprocal space5. In their original formulation the number of Wannier and underlying Bloch functions had to be the same (here J) and the subspace of used Bloch functions had to be well sepa-rated (disentangled) from the rest of the band structure. However, these additional constraints can be relaxed by the introduction of an appropriate disentanglement pro-cedure which generates a suitable subspace in advance [67] as it will be discussed in the following.

5Since the Wannier and Bloch functions are connected by a Fourier transform strong localization in real space of the former corresponds to smoothly varying Bloch function in reciprocal space.

In order to minimize the real-space spread Marzari and Vanderbilt introduced a spread functional Ω

Ω“ ÿJ

n“1

“x0n|r2|0ny ´ x0n|r|0ny2

“ ÿJ

n“1

“xr2yn´¯r2n

, (2.64)

which measures the quadratic spreads of each Wannier function in the unit cell around their centers. In the following we thus aim to minimize Ω by varying Umnk . Once the corresponding optimal transformation matrices are found Eq. (2.63) and Eq. (2.61) are used to construct the Wannier functions. In order to do so,Ω“ΩI`Ω˜ is separated into a gauge-invariant ΩI and gauge-dependent part Ω˜ via

Ω“ÿ

n

«

x0n|r2|0ny ´ÿ

Rm

|xRm|r|0ny|2

looooooooooooooooooooooomooooooooooooooooooooooon

I

`ÿ

n

ÿ

Rm‰0n

|xRm|r|0ny|2 looooooooooooomooooooooooooon

˜

. (2.65)

Since the whole procedure should be carried out in momentum space, Ω has to be formulated in reciprocal-space coordinates. The corresponding details shall not be discussed here, but can be found in Refs. [66] and [23]. In the end ΩI and Ω˜ are reformulated to

I “ 1 N

ÿ

k,b

wb

˜

J´ÿ

mn

ˇˇMmnk,bˇ ˇ2

¸

(2.66)

and

Ω˜ “ 1 Nk

ÿ

k,b

wb

ÿ

m‰n

ˇˇMmnk,bˇ ˇ2` 1

Nk

ÿ

k,b

wb

ÿ

n

`´Im lnMmnk,b´b¨¯rn˘2

, (2.67)

where Nk is the number of points in the k-mesh, J is the amount of involved Bloch states,bare connection vectors of neighbouringk-points,wb are corresponding weight-ing factors (see Ref. [66]) and

¯

rn“ ´ 1 Nk

ÿ

k,b

wbbIm lnMnnk,b (2.68) is the expectation value of r. In addition we used the matrices Mk,b which give the overlaps between Bloch states at neighbouring k-points

Mmnk,b“ xumk|un,k`by. (2.69)

The Role of ΩI: Disentangling the Bloch Bands ΩI can be rewritten with the help of the projector Pk “ř

n|unky xunk|and its comple-ment Qk “1´Pk leading to

I “ 1 N

ÿ

k,b

wbTrrPkQk+bs. (2.70) SinceUkare unitary matrices any gauge transformation according to Eq. (2.63) (acting on the Bloch functions|unky) will not change the projectorsP andQand ΩI therefore stays the same as well. Hence, ΩI is indeed gauge independent. In addition, by rewriting ΩI with the help of the P and Q we obtain a simple interpretation: The product of the projection operators PkQk+b will vanish as soon as Pk “ Pk+b which corresponds to a situation in which the difference between the subspaces atkand k+b vanishes. ΩI therefore measures the smoothness of the changes between neighbouring subspacesSpkqandSpk+bqor the “changes of characters” fromSpkqtoSpk+bq. Thus, in a situation in which N Wannier functions shall be constructed out of a group of JkěN entangled (partially degenerated) bands these changes of characters should be as small or as smooth as possible. To this end, ΩI has to be minimized to find a set of optimal subspaces Spkq Ă Fpkq for each k from the Jk-dimensional Hilbert space Fpkqof the entangled bands6. As shown by Souzaet al.in Ref. [67] the corresponding optimal projector Pk is obtained from the stationarity conditionδΩIptunkuq “0 which can be reformulated to an iterative and self-consistent scheme to solve the eigenvalue problem [67]:

«ÿ

b

wbPk+bpi´1q

|upiqnky “λpiqnk|upiqnky. (2.71) Thereby, the projector Pkpiq “řN

n“1|upiqnky xupiqnk| is updated in each iteration step using the eigenvectors |upiqnky corresponding to N highest eigenvalues λpiqnk. As soon as self-consistency is reached, an optimally smooth subspace Spkq and a minimized ΩI is obtained.

A particular advantage of this ΩI minimization procedure is the fact that certain constraints can be introduced. For instance, one might require that the Wannier functions will exactly reproduce the Bloch states eigenenergies in a certain “inner”

energy window. Thereby those parts of the original band structure which are most important for the subsequent considerations can be fixed7.

The Role of Ω: Maximal Localization˜

So far, we have seen thatΩI is in the sense of Eq. (2.63) gauge independent but can be used to disentangle sets of Bloch states. This means, remembering the initial intention

6IfJkN SpkqandFpkqare the same and there is no need to perform the disentanglement.

7See Ref. [67] for more details.

and definition of the complete spread functional Ω, that Ω˜ is in fact describing the degree of localization of the Wannier functions. Hence,Ω˜ must be minimized by finding adequate rotation matrices Uk (acting on the optimal smooth subspace Spkq only) to gain the maximally localized Wannier functions mentioned at the very beginning. Since Ω˜ is completely defined by the overlap matricesM, the minimization can be achieved by updating it according to

Mpi`1qk,b “Upiqk: Mpiqk,b Upiqk+b, (2.72) using

Upi`1qk “Upiqk exp`

∆Wpiqk ˘

, (2.73)

where∆W are properly chosen updates to U which only depend onM as discussed in more detail in Ref. [66]. Hence, with the initialMp0q as calculated from the underlying ab initio calculation and an initial Umnp0q “δmn the minimization or localization can be performed without any other ingredient. Having reached the maximal degree of localization the very last update ofUpiqk can finally be used to calculate the maximally localized Wannier functions using Eq. (2.63) acting on the “smooth” subspace Spkq and Eq. (2.61) afterwards. Thereby, it is important to note that eventually fixed eigenenergies of states within the smooth subset Spkq will not be changed due to the unitary transformation using Upiqk .

Initial Projections

For the minimization ofΩI we need an initial guess forSpkqor, correspondingly,|up0qnky. To this end we can utilize N localized trial functions gαprq onto which the initial Jk

Bloch states are projected

projαk y “

Jk

ÿ

m“1

mky xΦmk|gαy, (2.74) where gαprq should be functions which have the same angular momentum and are localized at the sites of the expected Wannier functions. For example, in the program package Wannier90[69], which will be frequently used throughout the thesis, orbitals of the hydrogen atom are used for this purpose. In practice the matrix

Ak “ xΦmk|gαy (2.75)

is calculated in advance and used afterwards in Eq. (2.74). Additionally it is used to obtain the overlap matrix Sαβk “ xΦprojαkprojβk y “ pAk:Akqαβ in order to orthogonalize the resulting projected Bloch states

orthβk y “ ÿN

α“1

´ Sk´1{2¯

αβprojαk yp2.74q

Jk

ÿ

m“1

´

ASk´1{2¯

mky. (2.76)

This orthogonalized subset of Bloch states|Φorthαk ycan be used afterwards as the starting point of a minimization procedure of ΩI to obtain an optimal (smooth) subspaceSpkq as discussed before.

Alternative Approaches

The combination of projection and orthogonalization as discussed above is already a proper gauge transformation as defined in Eq. (2.63) for the case of entangled bands.

The resulting Bloch states can be used in Eq. (2.61) to get localized Wannier functions.

Without further refinements of the gauge the resulting functions are called “one shot” or

“first guess” Wannier functions (FGWF) which preserve the trial functions’ symmetries [70]. The latter must not be fulfilled anymore after the minimization of Ω.˜

Following the ideas of Souza, Marzari and Vanderbilt, it seems to be useful to divide the whole process of deriving MLWF into two pieces, namely minimizing ΩI and Ω˜ separately in the case of disentangled bands in a periodic crystal. Nevertheless, there is no good argument against a direct minimization of the complete Ω. And indeed, as shown by Thygesen et al. an adequate procedure to directly minimize Ω leads to very similar results in the case of periodic crystals and can even be used for molecules without translational symmetries in which some Wannier orbitals are occupied only partially [71, 72].

Furthermore, other approaches to derive Wannier functions from ab initio calcula-tions can be found in the literature, which do not rely on the direct localization of the Wannier functions. For instance, there are variational approaches like given by Kohn [73] in which energy functionals are varied with respect to localized trial functions [74].

Additionally, there are methods which rely on the construction of Slater-Koster-like model Hamiltonians which can be used subsequently to interpolate theab initio band structure [75].

However, in the following the method by Souza, Marzari and Vanderbilt will be employed, since their approach results in very accurate descriptions of the systems under considerations, can be controlled precisely and the before-mentioned program package Wannier90 is already linked to all of the here employed DFT packages.

Single-Particle Matrix Elements and Basis Transformations

From the Ω minimization procedure the “rotated” Hamiltonian for the subspace Spkq HSpkq(rot)pkq “ Uk:HSpkqpkq Uk (2.77) is known on the initialk-grid. Via a standard discrete Fourier transform the real-space representation can be readily derived

Hαβ(rot)pRq “ xRα|H|0βy “ 1 Nk

ÿ

k

e´ikRHαβ(rot)pkq “tαβpRq, (2.78)

whereNk is the number of involvedk-points [76, 23]. As already indicated, the matrix elements of the real-space Hamiltonian are normally called hopping matrix elements and are denoted bytαβ. As soon as these hopping matrix elements are known,H(rot)pkq can be reconstructed at arbitrary k1-points via

Hαβ(rot)pk1q “ÿ

R

eik1RtαβpRq. (2.79) In order to derive the resulting band structure the eigenvalues ε˜nk1 of H(rot)pk1q have to be calculated via diagonalization

H(rot)pk1q |k1ny “ε˜nk1|k1ny (2.80) which yields the eigenenergies and the eigencoefficients cαnpk1q which can be used to transform arbitrary matrix elements from the orbital basis|k1αyto the eigen- or band-basis |k1ny

|k1ny “ÿ

α

cαnpk1q |k1αy. (2.81)