26.11.2014 Views

Maximally localized Wannier functions: Theory and applications

Maximally localized Wannier functions: Theory and applications

Maximally localized Wannier functions: Theory and applications

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

47<br />

the conductor to the leads (Datta, 1995; Fisher <strong>and</strong> Lee,<br />

1981),<br />

T (E) = Tr(Γ L G r CΓ R G a C), (125)<br />

where G {r,a}<br />

C<br />

are the retarded (r) <strong>and</strong> advanced (a)<br />

Green’s <strong>functions</strong> of the conductor, <strong>and</strong> Γ {L,R} are <strong>functions</strong><br />

that describe the coupling of the conductor to the<br />

left (L) <strong>and</strong> right (R) leads. Since G a = (G r ) † , we consider<br />

G r only <strong>and</strong> drop the superscript.<br />

Expressing the Hamiltonian H of the system in terms<br />

of a <strong>localized</strong>, real-space basis set enables it to be partitioned<br />

without ambiguity into sub-matrices that correspond<br />

to the individual subsystems. A concept that<br />

is particularly useful is that of a principal layer (Lee<br />

<strong>and</strong> Joannopoulos, 1981) (PL), which is a section of<br />

lead that is sufficiently long such that ⟨χ n i |H|χm j ⟩ ≃ 0 if<br />

|m − n| ≥ 2, where H is the Hamiltonian operator of the<br />

entire system <strong>and</strong> |χ n i ⟩ is the ith basis function in the n th<br />

PL. Truncating the matrix elements of the Hamiltonian<br />

in this way incurs a small error which is systematically<br />

controlled by increasing the size of the PL. The Hamiltonian<br />

matrix in this basis then takes the block diagonal<br />

form (see also Fig. 34)<br />

⎛<br />

. .. . . . . . . ..<br />

⎞<br />

· · · H¯0¯0<br />

L H¯1¯0<br />

L 0 0 0 · · ·<br />

· · · H¯1¯0†<br />

L<br />

H¯0¯0<br />

L h LC 0 0 · · ·<br />

H =<br />

· · · 0 h † LC H C h CR 0 · · ·<br />

, (126)<br />

· · · 0 0 h † CR<br />

⎜<br />

H00 R H01 R · · ·<br />

⎝ · · · 0 0 0 H 01†<br />

R<br />

HR 00 · · ·<br />

⎟<br />

⎠<br />

. .. . . . . . . ..<br />

where H C represents the Hamiltonian matrix of the conductor<br />

region, H¯0¯0<br />

L <strong>and</strong> H00 R are those of each PL of the<br />

left <strong>and</strong> right leads, respectively, H¯1¯0<br />

L <strong>and</strong> H01 R are couplings<br />

between adjacent PLs of lead, <strong>and</strong> h LC <strong>and</strong> h CR<br />

give the coupling between the conductor <strong>and</strong> the leads.<br />

In order to compute the Green’s function of the conductor<br />

one starts from the equation satisfied by the<br />

Green’s function G of the whole system,<br />

(ϵ − H)G = I (127)<br />

where I is the identity matrix, <strong>and</strong> ϵ = E + iη, where η<br />

is an arbitrarily small, real constant.<br />

From Eqs. (127) <strong>and</strong> (126), it can be shown that the<br />

Green’s function of the conductor is then given by (Datta,<br />

1995)<br />

G C = (ϵ − H C − Σ L − Σ R ) −1 , (128)<br />

where we define Σ L = h † LC g Lh LC <strong>and</strong> Σ R = h RC g R h † RC ,<br />

the self-energy terms due to the semi-infinite leads, <strong>and</strong><br />

g {L,R} = (ϵ − H {L,R} ) −1 , the surface Green’s <strong>functions</strong><br />

of the leads.<br />

The self-energy terms can be viewed as effective Hamiltonians<br />

that arise from the coupling of the conductor<br />

with the leads. Once the Green’s <strong>functions</strong> of the leads<br />

are known, the coupling <strong>functions</strong> Γ {L,R} can be easily<br />

obtained as (Datta, 1995)<br />

Γ {L,R} = i[Σ r {L,R} − Σa {L,R}], (129)<br />

where Σ a {L,R} = (Σr {L,R} )† .<br />

As for the surface Green’s <strong>functions</strong> g {L,R} of the semiinfinite<br />

leads, it is well-known that any solid (or surface)<br />

can be viewed as an infinite (semi-infinite in the case of<br />

surfaces) stack of principal layers with nearest-neighbor<br />

interactions (Lee <strong>and</strong> Joannopoulos, 1981). This corresponds<br />

to transforming the original system into a linear<br />

chain of PLs. Within this approach, the matrix elements<br />

of Eq. (127) between layer orbitals will yield a<br />

set of coupled equations for the Green’s <strong>functions</strong> which<br />

can be solved using an efficient iterative scheme due to<br />

Lopez-Sancho et al. (1984, 1985). Knowledge of the finite<br />

Hamiltonian sub-matrices in Eq. (126), therefore, is<br />

sufficient to calculate the conductance of the open leadconductor-lead<br />

system given by Eq. (124).<br />

There are a number of possibilities for the choice of<br />

<strong>localized</strong> basis |χ⟩. Early work used model tight-binding<br />

Hamiltonians (Anantram <strong>and</strong> Govindan, 1998; Chico<br />

et al., 1996; Nardelli, 1999; Saito et al., 1996), but the increasing<br />

sophistication of computational methods meant<br />

that more realistic first-principles approaches could be<br />

adopted (Buongiorno Nardelli et al., 2001; Fattebert <strong>and</strong><br />

Buongiorno Nardelli, 2003). <strong>Maximally</strong>-<strong>localized</strong> <strong>Wannier</strong><br />

<strong>functions</strong> were first used in this context by Calzolari<br />

et al. (2004), who studied Al <strong>and</strong> C chains <strong>and</strong> a (5,0)<br />

carbon nanotube with a single Si substitutional defect,<br />

<strong>and</strong> by Lee et al. (2005), who studied covalent functionalizations<br />

of metallic nanotubes - capabilities now encoded<br />

in the open-source packages <strong>Wannier</strong>90 (Mostofi<br />

et al., 2008) <strong>and</strong> WanT (Ferretti et al., 2005a). This was<br />

quickly followed by a number of <strong>applications</strong> to ever more<br />

realistic systems, studying transport through molecular<br />

junctions (Strange et al., 2008; Thygesen <strong>and</strong> Jacobsen,<br />

2005), decorated carbon nanotubes <strong>and</strong> nanoribbons<br />

(Cantele et al., 2009; Lee <strong>and</strong> Marzari, 2006; Li<br />

et al., 2011; Rasuli et al., 2010), organic monolayers (Bonferroni<br />

et al., 2008), <strong>and</strong> silicon nanowires (Shelley <strong>and</strong><br />

Mostofi, 2011), as well as more methodological work on<br />

improving the description of electron correlations within<br />

this formalism (Bonferroni et al., 2008; Calzolari et al.,<br />

2007; Ferretti et al., 2005b,c).<br />

The formulation described above relies on a <strong>localized</strong><br />

description of the electronic-structure problem, <strong>and</strong> it<br />

should be noted that several approaches to calculating<br />

electronic transport properties have been developed using<br />

<strong>localized</strong> basis sets rather than MLWFs, ranging from<br />

Gaussians (Hod et al., 2006) to numerical atomic orbitals<br />

(Br<strong>and</strong>byge et al., 2002; Markussen et al., 2006;<br />

Rocha et al., 2008).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!