cond-mat0309259/pap.tex
1: \documentclass[showpacs,aps,prb,twocolumn]{revtex4}
2: \usepackage{epsfig}
3: \usepackage{amsmath}
4: 
5: \def\rr{{\bf r}}
6: \def\xx{{\bf x}}
7: \def\kk{{\bf k}}
8: \def\pp{{\bf P}}
9: \def\jj{{\bf J}}
10: \def\aa{{\bf a}}
11: \def\RR{{\bf R}}
12: \def\bb{{\bf b}}
13: \def\ee{\boldsymbol{\cal E}}
14: \def\hh{\hat{H}}
15: \def\tk{\hat{T}_\kk}
16: \def\ket#1{\vert#1\rangle}
17: \def\bra#1{\langle#1\vert}
18: 
19: \begin{document}
20: 
21: \title{Dynamics of Berry-phase polarization in time-dependent electric fields}
22: 
23: \author{Ivo Souza}
24: \author{ Jorge \'I\~niguez}
25: \altaffiliation{Present address: NIST Center for Neutron Research and 
26: Department of Materials and Nuclear Engineering of the University of Maryland.}
27: \author{David Vanderbilt}
28: 
29: \affiliation{Department of Physics and Astronomy, Rutgers
30: University, Piscataway, New Jersey 08854-8019, USA}
31: 
32: \begin{abstract}
33: 
34: We consider the flow of polarization current $\jj=d\pp/dt$
35: produced by a homogeneous electric field $\ee(t)$ or by rapidly varying 
36: some other parameter in the Hamiltonian of a solid. 
37: For an initially insulating system and a collisionless time evolution,
38: the dynamic polarization $\pp(t)$
39: is given by a nonadiabatic version of the King-Smith--Vanderbilt 
40: geometric-phase formula.
41: This leads to a computationally convenient form for the Schr\"odinger equation
42: where the electric field is described by a linear scalar potential
43: handled on a discrete mesh in reciprocal space. 
44: Stationary solutions in sufficiently weak
45: static fields are local minima of the energy functional of Nunes and Gonze.
46: Such solutions only exist below a critical field that depends inversely
47: on the density of $k$ points. For higher fields they become long-lived
48: resonances, which can be accessed dynamically by gradually increasing $\ee$. 
49: As an illustration the dielectric function 
50: in the presence of a dc bias field is computed for a tight-binding model from
51: the polarization response to a step-function discontinuity in $\ee(t)$,
52: displaying the Franz-Keldysh effect. 
53: 
54: \end{abstract}
55: \date{\today}
56: \pacs{71.15.Qe, 78.20.Bh}
57: 
58: \maketitle
59: 
60: % ----------------------------------------------------------------------------
61: 
62: \section{INTRODUCTION}
63: \label{sec:intro}
64: 
65: A very successful theoretical and computational framework was developed by
66: King-Smith and Vanderbilt\cite{ksv93} for dealing within periodic boundary 
67: conditions with 
68: the macroscopic dielectric polarization of an 
69: insulator. The central result of the theory of bulk polarization (TBP)
70: is an expression for the electronic contribution
71: $\pp$ which takes the form of a Berry's phase\cite{berry84} 
72: of the valence-band Bloch
73: wave functions transported across the Brillouin zone (BZ). Alternatively, it 
74: can be recast in real space as the vector sum of the centers of charge of the 
75: valence-band Wannier functions. 
76: Practical prescriptions were devised for computing both the Berry's 
77: phase\cite{ksv93} and the Wannier functions,\cite{mv97} 
78: which have become standard features of first-principles electronic
79: structure codes. 
80: 
81: The measurable quantity accessed by the TBP is the change $\Delta \pp$ in 
82: macroscopic polarization induced by changing some parameter $\lambda$ in 
83: the electronic Hamiltonian $\hh(t)=\hh[\lambda(t)]$. The following 
84: assumptions were explicitly made in the original derivation\cite{ksv93}. 
85: (i) Adiabaticity: the change in $\lambda(t)$ is slow enough that
86: the electrons remain in the instantaneous ground 
87: state of $\hh(t)$, apart from small deviations 
88: proportional to $d\lambda/dt$ described by first-order adiabatic 
89: perturbation theory; (ii) the ground state of $\hh(t)$ remains insulating at 
90: all times, separated from excited states by finite energy gaps; (iii) $\hh(t)$
91: is lattice-periodic.
92: The first two assumptions are related in that the size of the energy gap sets 
93: the scale for deviations from adiabaticity.
94: 
95: A spatially homogeneous electric field necessarily violates either
96: (i) or (iii): if the field is introduced via a 
97: vector-potential  
98: ${\bf A}(t)=-c\int^t \boldsymbol{\cal E}(t')dt'$, $\hat{H}(t)$ remains 
99: lattice-periodic but changes nonadiabatically, even for a static 
100: field; if instead a scalar potential term
101: $e\boldsymbol{\cal E}(t)\cdot\hat{\bf r}$ is used, $\hat{H}(t)$ is no longer 
102: lattice-periodic.
103: Nevertheless, the TBP has been successfully applied to situations where 
104: electric fields are 
105: present,\cite{nunes94,dalcorso94,dalcorso96,nunes01,souza02,umari02,sai02}
106: but a
107: rigorous justification for doing so
108: is still lacking. 
109: 
110: In this paper we reexamine the TBP and find that it can be generalized as 
111: follows. Assumption (i) can be dropped altogether.
112: Assumption (ii)
113: is only 
114: invoked at $t=0$; the ensuing nonadiabatic dynamics
115: may admix considerable amounts of excited states into the occupied subspace.
116: Finally assumption (iii) can be relaxed
117: to allow for a linear scalar potential to be present in addition to the
118: periodic crystal potential. 
119: 
120: These generalizations extend the scope of the TBP to 
121: {\it nonadiabatic} polarization currents induced by time-dependent electric 
122: fields, or by other rapid changes in $\hat{H}(t)$ (e.g., the initial
123: nonthermal ionic motion that accompanies photoexcitation of the electrons by 
124: an intense laser pulse.\cite{mazur02})
125: The dynamical equations for the electrons 
126: that come out of this generalized TBP
127: are derived and applied in the context of a tight-binding model.
128: These equations are semiclassical (the electrons are treated
129: quantum-mechanically, whereas the electric field is treated classically)
130: and nonperturbative (electric fields of finite magnitude are allowed).
131: 
132: 
133: We begin by considering in Section~\ref{sec:dynamics_dm} some general 
134: properties of the 
135: coherent dynamics of Bloch electrons that are initially in an insulating state.
136: They are used in Section~\ref{sec:nonadiabatic_pol} to discuss
137: the macroscopic current $\jj(t)$, which is expressed as the rate of 
138: change of a dynamic polarization $\pp(t)$ given by nonadiabatic versions of 
139: the King-Smith--Vanderbilt expressions.
140: In Section~\ref{sec:dynamical} we derive from this generalized TBP a 
141: numerically convenient form for the time-dependent Schr\"odinger equation 
142: (TDSE) in the scalar-potential gauge, discretized on a mesh of $k$ points. 
143: Stable stationary solutions in static fields are discussed in 
144: Section~\ref{sec:stationary}. They exist only 
145: below a critical field ${\cal E}_c$ which decreases with
146: increasing $k$-point density, and are local minima
147: of the energy 
148: functional of Nunes and Gonze.\cite{nunes01,souza02,umari02} A 
149: prescription is given for computing them using an iterative diagonalization 
150: scheme.
151: In Section~\ref{sec:results} we show numerically on a tight-binding
152: model how the regime above the critical 
153: field can be accessed dynamically, by gradually increasing the electric field
154: beyond the critical value. We also compute 
155: the dielectric function of the same model in the presence of a static bias 
156: field, displaying the Franz-Keldysh effect.
157: 
158: % ----------------------------------------------------------------------------
159: 
160: \section{General properties of the dynamics}
161: \label{sec:dynamics_dm}
162: 
163: \subsection{Lattice-periodicity}
164: \label{sec:periodicity}
165: 
166: Here we expound in more detail an argument, sketched in 
167: Ref.~\onlinecite{souza02}, that makes use of the one-particle density matrix
168: to handle the presence of electric fields in a well-controlled
169: fashion.\cite{lazzeri03}
170: %
171: We say that the one-particle density matrix
172: %
173: $n(\rr,\rr')=\langle \rr|\hat{n}|\rr'\rangle$ is lattice-periodic if
174: %
175: \begin{equation}
176: \label{eq:dm_lattice_per}
177: n(\rr,\rr')=n(\rr+\RR,\rr'+\RR),
178: \end{equation}
179: %
180: where $\RR$ is a lattice vector. In particular, this implies periodicity of the
181: charge density. Suppose that (\ref{eq:dm_lattice_per}) is true at $t=0$ (e.g.,
182: the electrons are in the ground state of the crystal Hamiltonian
183: $\hat{H}^{0}(t=0)$).
184: At that time a homogeneous electric field is turned on, which 
185: may subsequently have an arbitrarily strong and rapid variation;
186: $\hat{H}^{0}(t)$ may also undergo arbitrarily rapid variations (but must
187: remain periodic). The full Hamiltonian in the scalar potential gauge is
188: %
189: \begin{equation}
190: \label{eq:hamiltonian}
191: \hat{H}(t)=\hat{H}^0(t)+\hat{H}^{\ee}(t),
192: \end{equation}
193: %
194: where $\hat{H}^{\ee}(t)=e{\ee}(t)\cdot\hat{\rr}$ 
195: describes the electric field in the dipole approximation,
196: and $-e$ is the electron charge. 
197: 
198: Let us show that in the absence of scattering the lattice-periodicity of 
199: $n(\rr,\rr')$ is preserved at all later times. 
200: It suffices to establish that $\dot{n}(\rr,\rr')=\dot{n}(\rr+\RR,\rr'+\RR)$.
201: The density matrix evolves according to
202: $i\hbar\, d\hat{n}/dt=[\hat{H},\hat{n}]$, or, in the position 
203: representation,
204: %
205: \begin{equation}
206: i\hbar\,\dot{n}(\rr,\rr')=\int [H(\rr,\xx)n(\xx,\rr')-
207: n(\rr,\xx)H(\xx,\rr')]d\xx.
208: \end{equation}
209: %
210: (When 
211: left unspecified, the domain of integration over spatial coordinates is
212: understood to be the entire space.)
213: For clarity we consider the effect of $\hat{H}^0$ and
214: $\hat{H}^{\ee}$ separately. The $\hat{H}^0$ term yields
215: %
216: \begin{eqnarray}
217: i\hbar\,\dot{n}(\rr+\RR,\rr'+\RR)&=&\int 
218: [H^0(\rr+\RR,\xx)n(\xx,\rr'+\RR)
219: \nonumber \\&-&n(\rr+\RR,\xx)H^0(\xx,\rr'+\RR)]d\xx.
220: \end{eqnarray}
221: %
222: Making the change of variables $\xx'=\xx-\RR$ and invoking the
223: lattice-periodicity of $\hat{H}^0$ and $\hat{n}$, we find
224: $\dot{n}(\rr+\RR,\rr'+\RR)=\dot{n}(\rr,\rr')$.
225: Using 
226: $\rr(\rr,\rr')=\langle \rr|\hat{\rr}|\rr'\rangle=\rr\,\delta(\rr-\rr')$ 
227: the contribution from $\hat{H}^{\ee}$  
228: is seen to have the same property:
229: %
230: \begin{eqnarray}
231: i\hbar\, \dot{n}(\rr+\RR,\rr'+\RR)&=&e\ee\cdot(\rr+\RR)\,n(\rr+\RR,\rr'+\RR) 
232: \nonumber \\&-&e\ee\cdot(\rr'+\RR)\,n(\rr+\RR,\rr'+\RR)\nonumber \\
233: &=&e\ee\cdot(\rr-\rr')\,n(\rr,\rr')=i\hbar\, \dot{n}(\rr,\rr').\nonumber\\
234: \end{eqnarray}
235: %
236: Hence $n(\rr,\rr')$ remains lattice-periodic under the action of the full
237: Hamiltonian (\ref{eq:hamiltonian}).
238: This was to be expected, since in the
239: vector potential gauge the Hamiltonian is periodic.\cite{krieger86}
240: The purpose of this exercise was to show explicitly
241: how this result comes about in the scalar-potential gauge,
242: where the
243: nonperiodicity of $\hat{H}$ has been a source of some confusion regarding
244: this issue.
245: 
246: 
247: \subsection{Wannier-representability}
248: 
249: The previous result on the conservation of lattice-periodicity is valid for 
250: both metals and insulators. In what follows we
251: shall specialize to the case where the system is initially in an insulating
252: state, in which case a stronger statement can be made regarding the nature of
253: the states at $t>0$.
254: 
255: We will assume the absence of spin degeneracy throughout, so that states are 
256: singly-occupied. In terms of the valence Bloch eigenstates of $\hat{H}^0(t=0)$,
257: the initial density matrix is
258: %
259: \begin{equation}
260: n(\rr,\rr';t=0)=\Omega_B^{-1}\sum_{n=1}^M\int d\kk\,\psi_{\kk n}(\rr)
261: \psi_{\kk n}^*(\rr'),
262: \end{equation}
263: %
264: where the integral is over the BZ of volume 
265: $\Omega_B=(2\pi)^3/v$,
266: and $M$ is the number of filled bands. 
267: (Clearly, such a density matrix is lattice-periodic. Its idempotency 
268: %
269: can
270: be checked using Eq.~(\ref{eq:normalization}).)
271: We shall prove that, as the density matrix evolves in time according to
272: %
273: \begin{equation}
274: \label{eq:dyn_dm}
275: i\hbar\,\dot{n}(\rr,\rr';t)=\bra{ \rr}[\hat{H}^0+\hat{H}^{\ee},\hat{n}]
276: \ket{\rr'},
277: \end{equation}
278: % 
279: it can still be expressed in the same form,
280: %
281: \begin{equation}
282: \label{eq:dm_bloch}
283: n(\rr,\rr';t)=\Omega_B^{-1}\sum_{n=1}^M\int d\kk\,\phi_{\kk n}(\rr,t)
284: \phi_{\kk n}^*(\rr',t).
285: \end{equation}
286: %
287: Although at $t>0$ the occupied states $\phi_{\kk n}(\rr,t)$ may depart 
288: significantly from the valence states of $\hat{H}^0(t)$,
289: they remain orthonormal and {\it Bloch-like}:
290: $\phi_{\kk n}(\rr,t)=e^{i\kk\cdot\rr}v_{\kk n}(\rr,t)$, with 
291: $v_{\kk n}(\rr+\RR,t)=v_{\kk n}(\rr,t)$.
292: 
293: The cell-periodic states $v_{\kk n}(\rr,t)$ are the central
294: objects in our formalism. For discussion purposes only, let us 
295: expand them in the set of eigenstates 
296: $u_{\kk m}(\rr,t)=e^{-i\kk\cdot\rr}\psi_{\kk m}(\rr,t)$
297: of the cell-periodic Hamiltonian
298: $\hat{H}_\kk^0(t)=e^{-i\kk\cdot\hat{\rr}}\hat{H}^0(t)e^{i\kk\cdot\hat{\rr}}$:
299: %
300: \begin{equation}
301: \label{eq:linear_comb}
302: \ket{v_{\kk n}(t)}= \sum_{m=1}^{\infty} 
303: c_{\kk,nm}(t)\,\ket{u_{\kk m}(t)}.
304: \end{equation}
305: %
306: Individual eigenstates will in general have 
307: fractional occupations $0\leq n_{\kk m}=\sum_{n=1}^M |c_{\kk,nm}|^2\leq 1$
308: at $t>0$, but the total population 
309: $n_\kk=\sum_{m=1}^\infty n_{\kk m}$ is the same for every $\kk$ and equals 
310: the number of filled bands at $t=0$.
311: This is intuitively clear, since a spatially homogeneous electric field causes
312: vertical transitions in $k$-space which amount to a redistribution of the
313: electron population among states with equal $\kk$; the same is
314: true for the transitions induced by varying the lattice-periodic 
315: $\hat{H}^0(t)$.
316: 
317: We will justify Eq.~(\ref{eq:dm_bloch}) by deriving a dynamics for
318: the $|v_{\kk n}\rangle$ that insures that Eq.~(\ref{eq:dm_bloch})
319: provides a solution to (\ref{eq:dyn_dm}).  Since there is a gauge
320: freedom
321: %
322: \begin{equation}
323: \label{eq:gauge_transf}
324: |v_{\kk n}\rangle\rightarrow\sum_{m=1}^M\,U_{\kk,mn}|v_{\kk m}\rangle
325: \end{equation}
326: %
327: ($U_\kk$ is a ${\bf k}$-dependent unitary $M\times M$ matrix)
328: in the definition of the $|v_{\kk n}\rangle$,\cite{mv97}
329: the evolution equation for them is not unique.  We require only that the
330: $|v_{\kk n}\rangle$ should yield the correct dynamics for the gauge-invariant 
331: density matrix, Eq.~(\ref{eq:dyn_dm}), and we will look for the simplest
332: solution that achieves this goal.
333: 
334: By hypothesis, at time $t$ $\dot{n}(\rr,\rr')$ takes
335: the form
336: %
337: \begin{eqnarray}
338: \label{eq:dm_dot}
339: \dot{n}(\rr,\rr')&=&\Omega_B^{-1}\sum_{n=1}^M\,\int d\kk\, 
340: e^{i\kk\cdot(\rr-\rr')} \,\times \nonumber\\
341: &&\quad
342: [\dot{v}_{\kk n}(\rr)v_{\kk n}^*(\rr')+v_{\kk n}(\rr)\dot{v}_{\kk n}^*(\rr')].
343: \end{eqnarray}
344: %
345: As in the previous subsection, we consider the contributions from $\hat{H}^0$
346: and $\hat{H}^{\ee}$ in Eq.~(\ref{eq:dyn_dm})
347: separately. The former is captured by 
348: $i\hbar|\dot{v}_{\kk n}\rangle=\hat{H}_\kk^0|v_{\kk n}\rangle$.
349: To deal with $\hat{H}^{\ee}$ we resort to manipulations familiar from the
350: crystal-momentum representation (CMR)\cite{blount62} (but with the crucial 
351: difference that in the CMR those manipulations are applied to the 
352: $\ket{u_{\kk n}}$, not to the $\ket{v_{\kk n}}$). We first observe that
353: %
354: \begin{eqnarray}
355: &&\langle \rr|[\hat{\rr},\hat{n}]|\rr'\rangle=(\rr-\rr')\,n(\rr,\rr') 
356: \nonumber \\
357: &=&\Omega_B^{-1}\sum_{n=1}^M\int d\kk\,v_{\kk n}(\rr)v_{\kk n}^*(\rr')
358: (-i\partial_\kk)e^{i\kk\cdot(\rr-\rr')}.
359: \end{eqnarray}
360: %
361: Integrating by parts and noting that in a {\it periodic gauge}
362: ($\phi_{\kk+{\bf G},n}=\phi_{\kk n}$) the boundary term vanishes, we obtain
363: %
364: \begin{eqnarray}
365: &&\langle \rr|[\hat{H}^{\ee},\hat{n}]|\rr'\rangle=
366: \Omega_B^{-1}\sum_{n=1}^M\int d\kk\,
367: e^{i\kk\cdot(\rr-\rr')}ie{\ee}\cdot \nonumber \\
368: &&\qquad\quad\Big[
369: \big(\partial_\kk v_{\kk n}(\rr)\big)v_{\kk n}^*(\rr')
370:   +v_{\kk n}(\rr)\big(\partial_\kk v_{\kk n}^*(\rr')\big) 
371: \Big].
372: \end{eqnarray}
373: %
374: Comparing with Eqs.~(\ref{eq:dyn_dm}) and (\ref{eq:dm_dot}) we arrive at
375: $i\hbar|\dot{v}_{\kk n}\rangle=ie\ee\cdot\partial_\kk|v_{\kk n}\rangle$.
376: The effect of $\hat{H}^{\ee}$ thus
377: takes the form of a $k$-derivative, and
378: the combined effect of $\hat{H}^0$ and $\hat{H}^{\ee}$ is
379: %
380: \begin{equation}
381: \label{eq:tdse_continuum}
382: i\hbar\ket{\dot{v}_{\kk n}}=(\hat{H}^0_\kk+
383: ie{\ee}\cdot\partial_\kk)\ket{{v}_{\kk n}}.
384: \end{equation}
385: %
386: This is our version of the TDSE for Bloch electrons in the scalar 
387: potential gauge, constructed in order that 
388: Eq.~(\ref{eq:dm_bloch}) will satisfy Eq.~(\ref{eq:dyn_dm}). 
389: The time-independent version was introduced as an {\it ansatz}
390: in Ref.~\onlinecite{nunes01}.\cite{foot:mistake}  The equivalence  of 
391: Eq.~(\ref{eq:tdse_continuum}) to other forms in the literature is established
392: in Appendix~\ref{app:pol}.
393: 
394: If at time $t$ the $M$ states $|v_{\kk n}\rangle$ at every $\kk$
395: are lattice-periodic and orthonormal,
396: the dynamics dictated by Eq.~(\ref{eq:tdse_continuum}) preserves those 
397: properties, i.e.,
398: $\dot{v}_{\kk n}(\rr+\RR)=\dot{v}_{\kk n}(\rr)$ and
399: $d\langle v_{\kk n}|v_{\kk m} \rangle/dt=0$. This can be seen
400: as follows. Starting from
401: %
402: \begin{equation}
403: i\hbar\, \dot{v}_{\kk n}(\rr+\RR)=\int 
404: H_k^0(\rr+\RR,\xx)v_{\kk n}(\xx)\,d\xx+
405: ie\ee\cdot\partial_\kk v_{\kk n}(\rr+\RR),
406: \end{equation}
407: %
408: making the change of variables $\xx'=\xx-\RR$, and invoking the assumed
409: lattice-periodicity of both $\hat{H}^0$ and $v_{\kk n}(\rr)$, the 
410: right-hand-side becomes $i\hbar\dot{v}_{\kk n}(\rr)$. 
411: As for orthonormality, Eq.~(\ref{eq:tdse_continuum}) 
412: yields\cite{foot:wavepacket}
413: %
414: \begin{equation}
415: \label{eq:orthonormality}
416: \frac{d}{dt}\langle v_{\kk n}|v_{\kk m}\rangle=\frac{e}{\hbar}\ee\cdot
417: \partial_\kk\langle v_{\kk n}|v_{\kk m}\rangle.
418: \end{equation}
419: Since by hypothesis 
420: %
421: \begin{equation}
422: \label{eq:norm_conv}
423: \langle v_{\kk n}|v_{\kk m}\rangle\equiv \int_v v_{\kk n}^*(\rr)
424: v_{\kk m}(\rr)\,d\rr=\delta_{n,m}, 
425: \end{equation}
426: %
427: where the integral is over a unit cell, the
428: right-hand-side of Eq.~(\ref{eq:orthonormality}) vanishes. 
429: This completes the proof of Eq.~(\ref{eq:dm_bloch}).
430: 
431: Two assumptions were made in the above derivation. The first is that the states
432: $\ket{v_{\kk n}}$ vary smoothly with $\kk$, so that $k$-derivatives
433: exist; we will come back to this point in Sec.~\ref{sec:dyn_pol_disc}.
434: The second is that the dynamics is scattering-free.
435: Note that Eq.~(\ref{eq:orthonormality}) is closely related to the
436: collisionless Boltzmann equation; incoherent scattering would destroy the
437: constancy of the total population $n_\kk$ by inducing transitions between
438: different $k$ points.
439: 
440: Having established that the occupied manifold is spanned by $M$ 
441: Bloch-like states at each $\kk$, we now transform them
442: into Wannier-like states
443: $\langle \rr|W_{{\bf R}n}(t)\rangle=W_n(\rr-{\bf R},t)$
444: in the usual way,
445: %
446: \begin{equation}
447: \label{eq:wannier_like}
448: \ket{W_{{\rm\bf R}n}(t)}=
449: \Omega_B^{-1}\sum_{m=1}^M\,\int d\kk\, 
450: e^{-i\kk\cdot{\rm\bf R}}\,U_{\kk,mn}(t)\,\ket{\phi_{\kk m}(t)}, 
451: \end{equation}
452: %
453: where a periodic gauge is assumed and we have inserted a unitary rotation
454: (\ref{eq:gauge_transf}) among the occupied states.
455: The assumption that by a judicious choice of the matrices $U_\kk(t)$
456: the Bloch-like states can be made to vary smoothly with $\kk$ is equivalent
457: to the assumption that the Wannier-like states can be chosen to be
458: well localized.\cite{mv97}
459:  
460: The density matrix (\ref{eq:dm_bloch}) can now be recast as
461: %
462: \begin{equation}
463: \label{eq:w-r}
464: n(\rr,\rr';t)=\sum_{n=1}^M\sum_{{\rm\bf R}}
465: W_{{\rm\bf R}n}(\rr,t)\,W_{{\rm\bf R}n}^*(\rr',t).
466: \end{equation}
467: %
468: We will term {\it Wannier-representable} (WR) a state whose density matrix 
469: is of this form.  An insulating ground state is WR,
470: while a metallic state is not. We have established that under the Hamiltonian
471: (\ref{eq:hamiltonian}) and in the absence of
472: scattering, an initially insulating system remains WR, or 
473: ``insulating-like'', even if at some later time the ground state of
474: $\hat{H}^0(t)$ becomes metallic.\cite{foot:frustrated_metal}
475: Unlike a true insulating ground state, or a stationary 
476: field-polarized state,\cite{souza02} a dynamic WR state
477: will in general break time-reversal symmetry and carry a macroscopic current. 
478: This is the subject of the next Section.
479: 
480: % ----------------------------------------------------------------------------
481: 
482: \section{Dynamic polarization and current} 
483: \label{sec:nonadiabatic_pol}
484: 
485: \subsection{Derivation}
486: \label{sec:dyn_pol_der}
487: 
488: Our aim in this Section is to show that
489: a WR state carries a current that can be expressed as the 
490: rate of change of a polarization per unit volume,
491: %
492: \begin{equation}
493: \label{eq:curr_pol}
494: \jj(t)=\frac{d\pp(t)}{dt},
495: \end{equation}
496: %
497: where $\pp(t)$ is given in a periodic gauge by
498: %
499: \begin{equation}
500: \label{eq:dyn_pol}
501: P_{\alpha}(t)=-\frac{ie}{{(2\pi)}^3}\,\sum_{n=1}^M\int d\kk\,
502: \langle v_{\kk n}(t)|\partial_{k_{\alpha}}v_{\kk n}(t)\rangle
503: \end{equation}
504: %
505: ($\alpha$ is a cartesian direction) or, equivalently, by
506: %
507: \begin{equation}
508: \label{eq:pol_wannier}
509: \pp(t)=-\frac{e}{v}\,\sum_{n=1}^M\,\int\,\rr\,
510: \vert W_n(\rr,t)\vert^2\,d\rr.
511: \end{equation}
512: %
513: Eqs.~(\ref{eq:dyn_pol})-(\ref{eq:pol_wannier}) are identical to the 
514: King-Smith--Vanderbilt expressions appropriate
515: for the adiabatic regime and $\ee=0$,\cite{ksv93} except that in 
516: (\ref{eq:dyn_pol}) the 
517: valence-band eigenstates $\ket{u_{\kk n}}$
518: have been replaced by the instantaneous solutions 
519: of the TDSE (\ref{eq:tdse_continuum}),
520: and the $W_n(\rr,t)$ in Eq.~(\ref{eq:pol_wannier})
521: are the Wannier states corresponding to the $v_{\kk n}(\rr,t)$.
522: Eq.~(\ref{eq:dyn_pol}) can be interpreted as
523: a {\it nonadiabatic} geometric phase.\cite{aa87}
524: 
525: As in the adiabatic case, 
526: $\pp(t)$ is invariant under the transformation (\ref{eq:gauge_transf}) only 
527: up to a ``quantum of polarization'' $(e/v){\bf R}$.
528: Naturally, this gauge indeterminacy does not affect the measurable $\jj(t)$.
529: The total change in bulk polarization in a time interval $[0,T]$ is also 
530: well-defined as the integrated current: $\Delta \pp=\int_0^T \jj(t)\,dt$.
531: It can be determined, apart from an integer multiple of the quantum, by 
532: evaluating $\pp(t)$ at the endpoints: $\Delta \pp=\pp(T)-\pp(0)$.
533: In practice the remaining indeterminacy can be removed in the manner described
534: in Ref.~\onlinecite{ksv93}, by evaluating $\pp(t)$ with sufficient frequency during
535: that interval.
536: 
537: To establish Eqs.~(\ref{eq:curr_pol})-(\ref{eq:dyn_pol}), we first
538: evaluate $d\pp/dt$ by taking the time derivative of
539: Eq.~(\ref{eq:dyn_pol}) and obtain, after an integration by parts,
540: %
541: \begin{equation}
542: \label{eq:dpdt}
543: \frac{dP_\alpha}{dt}=-\frac{ie}{{(2\pi)}^3}\,\sum_{n=1}^M\, \int d\kk\,
544: \big[
545:   \langle \dot{v}_{\kk n}|\partial_{k_{\alpha}}v_{\kk n}\rangle-\mbox{c.c.}
546: \big].
547: \end{equation} 
548: %
549: Inserting the TDSE, Eq.~(\ref{eq:tdse_continuum}), we note that the
550: contribution arising from the second term therein, which
551: explicitly involves $\ee$, may be written as
552: %
553: \begin{equation}
554: \label{eq:vanishing_part}
555: \widetilde{J}_{\alpha}=\frac{e^2}{{(2\pi)}^3\hbar}\,\sum_{n=1}^M\sum_{\beta}\,
556: {\cal E}_{\beta}\int d\kk\,\Omega^{(n)}_{\alpha\beta}(\kk)
557: \end{equation}
558: %
559: where
560: %
561: \begin{equation}
562: \label{eq:curvature}
563: \Omega^{(n)}_{\alpha\beta}(\kk)=i\big[ 
564: \langle\partial_{k_{\alpha}}v_{\kk n}|\partial_{k_{\beta}}v_{\kk n}\rangle-
565: \langle\partial_{k_{\beta}}v_{\kk n}|\partial_{k_{\alpha}}v_{\kk n}\rangle
566: \big].
567: \end{equation}
568: %
569: This takes the form of a (nonadiabatic) Berry
570: curvature.\cite{explan-notanom} Using Stokes' theorem,
571: its volume integral can be turned into
572: a surface integral around the edges of the BZ of the Berry connection
573: ${\bf A}_{\kk,nn}$, where
574: %
575: \begin{equation}
576: \label{eq:connection}
577: A_{\kk,mn}^\alpha=i\langle v_{\kk m}|\partial_{k_\alpha} v_{\kk n}\rangle.
578: \end{equation}
579: %
580: Such an integral vanishes in a periodic gauge, so that 
581: $\widetilde{J}_{\alpha}=0$.  The remaining contribution, arising
582: from the insertion of the first term of Eq.~(\ref{eq:tdse_continuum})
583: into Eq.~(\ref{eq:dpdt}), then gives
584: %
585: \begin{equation}
586: \label{eq:curr}
587: \frac{dP_{\alpha}}{dt}
588:   =\frac{e}{{(2\pi)}^3\hbar}\,\sum_{n=1}^M\,\int d\kk
589: \big[
590:   \langle v_{\kk n}|\hat{H}_{\kk}^0|\partial_{k_{\alpha}} v_{\kk n}\rangle
591:   +{\rm c.c.}
592: \big].
593: \end{equation}
594: %
595: 
596: On the other hand, the current is
597: %
598: \begin{equation}
599: \label{eq:current}
600: J_{\alpha}=-\frac{e}{v}\,{\rm Tr}_c(\hat{n}\hat{v}_{\alpha}).
601: \end{equation}
602: %
603: Here 
604: ${\rm Tr}_c$ denotes the trace per unit cell, 
605: %
606: \begin{equation}
607: \label{eq:trace}
608: {\rm Tr}_c(\hat{\cal O})=\frac{1}{N}\,\int 
609: {\cal O}(\rr,\rr)\,d\rr,
610: \end{equation}
611: %
612: where $N$ is the (formally infinite) number of real-space cells in the system.
613: The velocity operator is defined as
614: %
615: \begin{equation}
616: \label{eq:velocity}
617: \hat{v}_{\alpha}=\frac{1}{i\hbar}\,[\hat{r}_{\alpha},\hat{H}].
618: \end{equation}
619: %
620: Inserting the Hamiltonian (\ref{eq:hamiltonian}) and using
621: $[\hat{r}_\alpha,\hat{H}^{\boldsymbol{\cal E}}]=0$,
622: %
623: \begin{equation}
624: \label{eq:velocity_b}
625: \hat{v}_{\alpha}=\frac{1}{i\hbar}\,[\hat{r}_{\alpha},\hat{H}^0].
626: \end{equation}
627: % 
628: In the position representation we find, combining (\ref{eq:dm_bloch}), 
629: (\ref{eq:current}) and (\ref{eq:velocity_b}), 
630: invoking the lattice-periodicity of the integrand to replace
631: $(1/N)\int d\rr$ by $\int_v d\rr$,
632: and inserting the identity
633: $\hat{\bf 1}=\int d\rr'\ket{\rr'}\bra{\rr'}$,
634: %
635: \begin{eqnarray}
636: J_{\alpha}&=&
637: -\frac{e}{{(2\pi)}^3\hbar}\sum_{n=1}^M\,\int d\kk\,\int_v d\rr\,
638: \int d\rr'\,
639: v_{\kk n}^*(\rr')v_{\kk n}(\rr) \nonumber \\ 
640: &\times&
641: H^0(\rr',\rr)\partial_{k_{\alpha}}
642: e^{-i\kk\cdot (\rr'-\rr)}.
643: \end{eqnarray}
644: %
645: Integrating by parts in $k_\alpha$ 
646: (the boundary term vanishes in a periodic gauge),
647: and using
648: %
649: \begin{equation}
650: H_{\kk}^0(\rr',\rr)=
651: e^{-i\kk\cdot(\rr'-\rr)}H^0(\rr',\rr),
652: \end{equation}
653: %
654: $J_{\alpha}$ reduces to exactly the same expression
655: appearing on the right-hand side of Eq.~(\ref{eq:curr}).  This
656: completes the proof of Eqs.~(\ref{eq:curr_pol})-(\ref{eq:dyn_pol}) for WR
657: states evolving under the Hamiltonian (\ref{eq:hamiltonian}).
658: 
659: We note in passing that the integral on the right-hand side of
660: Eq.~(\ref{eq:curr}) can be recast as
661: %
662: \begin{equation}
663: \int d\kk
664: \big[
665:   \partial_{k_\alpha}
666:   \langle v_{\kk n}|\hat{H}_{\kk}^0| v_{\kk n}\rangle-
667:   \langle v_{\kk n}|\big(\partial_{k\alpha}\hat{H}_{\kk}^0\big)|v_{\kk n}
668:   \rangle
669: \big].
670: \end{equation}
671: %
672: The first term vanishes in a periodic gauge, leading to the more 
673: familiar-looking form
674: %
675: \begin{equation}
676: J_{\alpha}=-\frac{e}{{(2\pi)}^3}\,\sum_{n=1}^M\,\int d\kk\,
677: \langle v_{\kk n}|\hat{v}_\alpha(\kk)| v_{\kk n}\rangle,
678: \end{equation}
679: %
680: where 
681: $\hat{v}_\alpha(\kk)=(1/\hbar)
682: \big(\partial_{k_\alpha}\hat{H}_{\kk}^0\big)$.\cite{blount62}
683: 
684: The above derivations (and
685: indeed all the results in this paper) remain valid for nonlocal
686: pseudopotentials such as those used in {\it ab initio} calculations,
687: since the definition of the velocity as the commutator
688: (\ref{eq:velocity}) remains valid for such pseudopotentials.
689: 
690: 
691: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
692: 
693: \subsection{Discussion}
694: \label{sec:dyn_pol_disc}
695: 
696: It is remarkable that a knowledge of the wave functions at $t=0$ and
697: $t=T$ is sufficient to infer, to within a factor of $(e/v){\bf R}$, the net 
698: amount of current that flowed through the bulk in the intervening time. This 
699: is a direct consequence of representability by localized
700: Wannier functions, that is, of the 
701: insulating-like character of the many-electron system. 
702: For such systems the integral in Eq.~(\ref{eq:pol_wannier}) can be
703: evaluated, and it becomes possible to track the time evolution of the
704: electronic center of mass, i.e., of $\pp$. Indeed, the center of mass 
705: can be meaningfully defined within periodic boundary conditions
706: only for many-electron states that are localized in the manner of
707: insulating states.\cite{kohn64,souza00} 
708: Under these conditions, the history of the coherent current flow is contained 
709: (modulo the quantum of polarization) in
710: the initial and final wave functions, related by
711: the time evolution operator $\exp[-(i/\hbar)\int_0^T \hat{H}(t) dt]$.
712: 
713: This result was previously established for adiabatic 
714: charge flow,\cite{ksv93} under the assumption that the ground
715: state is separated from excited states by finite energy gaps
716: everywhere in the BZ. In nonadiabatic situations the occupied
717: manifold acquires a significant excited-state admixture, so that
718: it becomes impossible to identify an energy gap.
719: Instead, underlying the derivation in Sec.~\ref{sec:dyn_pol_der}
720: is a weaker assumption, namely, that the many-electron state has a
721: localized nature, as reflected by the ability to construct,
722: via Eq.~(\ref{eq:wannier_like}), Wannier functions having a finite localization
723: length.\cite{souza00,resta99} (Numerical calculations of the
724: localization length will be presented in Sec.~\ref{sec:results}.)
725: For instance, when taking $k$-derivatives, we assumed a  
726: ``differentiable gauge'' for the $\ket{v_{\kk n}}$.  This is only
727: possible if the character of the electronic manifold changes
728: slowly with $\kk$, which is precisely what is measured by the
729: localization length.\cite{mv97} These observations are
730: in line with Kohn's viewpoint that the defining feature of the
731: insulating state is wave function localization, not the existence of
732: an energy gap.\cite{kohn64}
733: 
734: 
735: % ----------------------------------------------------------------------------
736: 
737: \section{Dynamical equations}
738: \label{sec:dynamical}
739: 
740: Having found the Berry-phase formula (\ref{eq:dyn_pol}) for the dynamic
741: polarization in the presence of a field $\ee(t)$, 
742: let us now use it to obtain computationally tractable dynamical equations under
743: the Hamiltonian (\ref{eq:hamiltonian}). The starting point is the 
744: observation that the dipole term $\hat{H}^{\ee}$ contributes
745: $-v\pp(t)\cdot\ee(t)$ to the energy per unit cell. 
746: An energy functional valid for periodic boundary conditions is then
747: obtained by expressing $\pp(t)$ via the TBP 
748: formulas. This program was previously carried out for insulators in static
749: fields,\cite{nunes94,nunes01,souza02,umari02} where 
750: stationary states were computed by minimizing that energy
751: functional after applying a regularization procedure (truncation of 
752: the Wannier functions in real space or discretization of $k$-space).
753: Our strategy for the time-dependent problem is to impose stationarity on the 
754: corresponding action functional. Following
755: Refs.~\onlinecite{nunes01,souza02}, we adopt here a $k$-space formulation,
756: which is particularly well-suited for numerical work.
757: Special emphasis will be put on
758: the discrete-$k$ case since this is the relevant one for numerical
759: implementations.
760: 
761: 
762: 
763: \subsection{Continuum-$k$ case}
764: \label{sec:td_continuum}
765: 
766: In the continuum-$k$ limit the TDSE may be formally obtained 
767: from a Lagrangian density ${\cal L}(\kk)$
768: such that the Lagrangian per unit cell is
769: $L=\Omega_B^{-1}\int d\kk\,{\cal L}(\kk)$.
770: For WR states under the Hamiltonian (\ref{eq:hamiltonian}) we have
771: %
772: \begin{equation}
773: \label{eq:lagrangian_density}
774: {\cal L}(\kk)=i\hbar \sum_{n=1}^M\,\langle v_{\kk n} | \dot{v}_{\kk n} \rangle 
775: -E(\kk),
776: \end{equation}
777: %
778: where
779: %
780: \begin{equation}
781: \label{eq:energy_k_density}
782: E(\kk)=\sum_{n=1}^M\,
783: \langle v_{\kk n} | \hat{H}_{\kk}^0+
784: ie{\ee}\cdot\partial_\kk | v_{\kk n} \rangle.
785: \end{equation}
786: %
787: Using Eq.~(\ref{eq:dyn_pol}) and defining the zero-field energy functional
788: %
789: \begin{equation}
790: \label{eq:energy0}
791: E^0=\Omega_B^{-1}\sum_{n=1}^M\,\int d\kk\,
792: \bra{v_{\kk n}}\hat{H}_\kk^0\ket{v_{\kk n}},
793: \end{equation}
794: %
795: one finds the total energy functional
796: %
797: \begin{equation}
798: \label{eq:energy_continuum}
799: E=\Omega_B^{-1}\int d\kk\,E(\kk)=E^{0}-
800: v\pp\cdot\ee.
801: \end{equation}
802: %
803: The Euler-Lagrange equation\cite{goldstein80} 
804: %
805: \begin{equation}
806: \label{eq:euler-lagrange}
807: \frac{d}{dt}
808: \frac{\delta {\cal L}}{ \langle\delta \dot{v}_{\kk n} |}+
809: \frac{d}{d\kk} 
810: \frac{\delta {\cal L}}{ \langle \delta \partial_{\kk} v_{\kk n} |}-
811: \frac{\delta {\cal L}}{ \langle \delta v_{\kk n} |}=0
812: \end{equation}
813: %
814: then leads to the dynamical equation (\ref{eq:tdse_continuum}).
815: 
816: As already mentioned, the choice of dynamical equation for the 
817: $\ket{v_{\kk n}}$
818: is not unique. An
819: alternative to Eq.~(\ref{eq:tdse_continuum}) is
820: %
821: \begin{equation}
822: \label{eq:tdse_continuum_cov}
823: i\hbar\ket{\dot{v}_{\kk n}}=\big(\hat{H}^0_\kk+
824: ie{\ee}\cdot\widetilde{\partial}_\kk\big)\ket{{v}_{\kk n}}.
825: \end{equation}
826: %
827: The bare derivative $\partial_\kk$ has been 
828: replaced by 
829: %
830: \begin{equation}
831: \label{eq:cov_der}
832: \widetilde{\partial}_\kk= 
833: \partial_\kk+i\sum_{m,n=1}^M\,{\bf A}_{\kk,mn}\ket{v_{\kk m}}\bra{v_{\kk n}},
834: \end{equation}
835: %
836: where ${\bf A}_{\kk,mn}$ is given by Eq.~(\ref{eq:connection}).
837: The operator $\widetilde{\partial}_\kk$ is a multiband version of the covariant
838: derivative\cite{fradkin} and is discussed further in Appendix~\ref{app:cov}.
839: Although the field-coupling term in Eq.~(\ref{eq:tdse_continuum_cov})
840: is no longer a scalar potential term in the strict sense, we will continue
841: to view it as such in a generalized sense. 
842: 
843: Eq.~(\ref{eq:tdse_continuum_cov}) preserves the orthonormality of the
844: $\ket{v_{\kk n}}$ and generates the correct dynamics for the density matrix.
845: (These properties rely on
846: $(A_{\kk}^\alpha)^{\dagger}=A_{\kk}^\alpha$, which follows from 
847: $\partial_{k_\alpha}\bra{v_{\kk n}}v_{\kk m}\rangle=0$.\cite{foot:wavepacket_b})
848: The latter is most easily seen from the dynamics of the projector
849: %
850: \begin{equation}
851: \label{eq:projector}
852: \hat{P}_\kk=\sum_{n=1}^M \ket{v_{\kk n}}\bra{v_{\kk n}},
853: \end{equation}
854: %
855: which completely specifies the occupied subspace at $\kk$ while being 
856: insensitive to unitary rotations inside that subspace.
857: After some algebra, it can be shown
858: that while the individual $\ket{v_{\kk n}}$ behave differently under
859: Eqs.~(\ref{eq:tdse_continuum}) and (\ref{eq:tdse_continuum_cov}),
860: $\hat{P}_\kk$ stays the same.
861: 
862: An advantage of introducing Eq.~(\ref{eq:tdse_continuum_cov}) in
863: place of Eq.~(\ref{eq:tdse_continuum}) is that, upon the discretization
864: of $k$-space, the former leads to an evolution equation at point
865: $\kk$ that is gauge-covariant (in the sense of transforming in the
866: obvious way under unitary rotations among occupied states at $\kk$ and
867: being invariant under such rotations at neighboring points
868: $\kk'$), as will become clear in the next section.
869: 
870: % ----------------------------------------------------------------------------
871: 
872: \subsection{Discrete-$k$ case}
873: \label{sec:td_discrete}
874:  
875: This is the relevant case for numerical work.
876: The Lagrangian for a uniform mesh of $N$ points in the BZ is
877: %
878: \begin{equation}
879: \label{eq:lagrangian_discrete}
880: L=\frac{i\hbar}{N}\sum_{n=1}^M \sum_{\kk}\,
881: \langle v_{\kk n}|\dot{v}_{\kk n} \rangle-E,
882: \end{equation}
883: %
884: where $E$ is the energy in an electric field,
885: %
886: \begin{equation}
887: \label{eq:energy}
888: E=E^0-v\ee\cdot\pp,
889: \end{equation}
890: %
891: with
892: %
893: \begin{equation}
894: \label{eq:energy0_discrete}
895: E^0=\frac{1}{N}\sum_{n=1}^M \sum_{\kk}\,
896: \langle v_{\kk n} | \hat{H}_{\kk}^0 | v_{\kk n} \rangle
897: \end{equation}
898: %
899: and a discretized expression for $\pp$ to be given
900: shortly.  Applying the Lagrangian equations of motion
901: \cite{goldstein80}
902: %
903: \begin{equation}
904: \label{eq:euler-lagrange_b}
905: \frac{d}{dt} 
906: \frac{\delta L}{ \langle \delta \dot{v}_{\kk n} |}-
907: \frac{\delta L}{\langle \delta v_{\kk n} |}=0
908: \end{equation}
909: %
910: yields
911: %
912: \begin{equation}
913: \label{eq:td_dva}
914: i\hbar\frac{d}{dt}|v_{\kk n}\rangle=\hat{H}_{\kk}^0 | v_{\kk n} \rangle
915: -N v \ee \cdot
916: \frac{\delta \pp}{\bra{\delta v_{\kk n}}}.
917: \end{equation}
918: %
919: Writing
920: %
921: \begin{equation}
922: \pp=\frac{1}{2\pi} \sum_{i=1}^3\, \aa_i\,(\pp\cdot\bb_i)
923: \end{equation}
924: %
925: where $\aa_i$ and $\bb_i$ are the direct and reciprocal lattice
926: vectors respectively, and defining
927: %
928: \begin{equation}
929: \label{eq:pdotb}
930: v\pp\cdot{\bf b}_i=-e\overline{\Gamma}_i,
931: \end{equation} 
932: %
933: the last term in Eq.~(\ref{eq:td_dva}) becomes
934: %
935: \begin{equation}
936: \label{eq:td_dvb}
937: +\frac{N e}{2\pi} \, \sum_{i=1}^3\, (\ee\cdot\aa_i)\,
938: \frac{\delta \overline{\Gamma}_i}{\langle \delta v_{\kk n} |} .
939: \end{equation}
940: %
941: 
942: According to the Berry-phase theory of polarization,\cite{ksv93}
943: $\overline{\Gamma}_i$ is the string-averaged 
944: discretized geometric phase along the $\bb_i$ direction,
945: %
946: \begin{equation}
947: \label{eq:discrete_phase}
948: \overline{\Gamma}_i=
949: -\frac{1}{N^\perp_i}\,\sum_{l=1}^{N^\perp_i}\,{\rm Im}\,
950: {\rm ln}\prod_{j=0}^{N_i^\parallel-1}\,{\rm det}\,S(\kk_j^{(i)},\kk_{j+1}^{(i)}).
951: \end{equation}
952: %
953: Here
954: $S_{mn}(\kk,\kk')=\langle v_{\kk m}|v_{\kk'n}\rangle$ is the $M\times M$ 
955: overlap matrix, $N^\perp_1$ is the number of strings
956: along ${\bf b}_1$, each containing $N_1^\parallel$ points 
957: $\kk_j^{(1)}=\kk_{\perp}^{(1)}+j\Delta \kk_1$,
958: $\kk_{\perp}^{(1)}$ is a point on the $(\bb_2,\bb_3)$ plane labeled by $l$, 
959: and $\Delta \kk_1=\bb_1/N_1^\parallel$.
960: Eqs.~(\ref{eq:pdotb}) and (\ref{eq:discrete_phase}) provide the discretized
961: %formulation 
962: version
963: of the nonadiabatic Berry-phase polarization,
964: Eq.~(\ref{eq:dyn_pol}). 
965: (A discrete-$k$ formula for the macroscopic current $\jj(t)$ is
966: given in Appendix~\ref{app:discr}.)
967: As in the continuum case, a periodic gauge is assumed.
968: 
969: 
970: A compact expression for 
971: $\delta \overline{\Gamma}_i/\bra{\delta v_{\kk n}}$
972: is derived in Appendix \ref{app:grad-b}
973: using the following notation. Let 
974: $\kk i\sigma=\kk+\sigma\Delta\kk_i$,
975: where $\sigma=\pm1$. The overlap matrix becomes $S_{\kk i\sigma,mn}=
976: \langle v_{\kk m}|v_{\kk i\sigma,n}\rangle$. Next we define
977: %
978: \begin{equation}
979: \label{eq:def-vtkis}
980: |\widetilde{v}_{\kk i\sigma,n}\rangle=\sum_{m=1}^M
981: {\big(S_{\kk i\sigma}^{-1}\big)}_{mn}
982: |v_{\kk i\sigma,m}\rangle
983: \end{equation} 
984: %
985: which is a ``dual'' of $|v_{\kk n}\rangle$ in the space of
986: the $|v\rangle$'s at the neighboring point $\kk i\sigma$, since
987: %
988: \begin{equation}
989: \label{eq:duality}
990: \langle v_{\kk n}|\widetilde{v}_{\kk i\sigma,m}\rangle=\delta_{n,m}.
991: \end{equation}
992: %
993: The $|\widetilde{v}_{\kk i\sigma,n}\rangle$ are gauge-covariant in the
994: sense that (i) they are invariant under
995: unitary rotations among the $\ket{v_{\kk i\sigma,n}}$ at any neighboring
996: point $\kk i\sigma$, and (ii) they transform under
997: unitary rotations among the $\ket{v_{\kk n}}$ in the same
998: manner as the $\ket{v_{\kk n}}$ themselves, i.e.,
999: %
1000: \begin{equation}
1001: \label{eq:gauge_cov}
1002: \ket{\widetilde{v}_{\kk i\sigma,n}}\rightarrow
1003: \sum_{m=1}^M\,U_{\kk,mn}\ket{\widetilde{v}_{\kk i\sigma,m}}.
1004: \end{equation}
1005: %
1006: Then it is shown in Appendix \ref{app:grad-b} that
1007: %
1008: \begin{equation}
1009: \label{eq:grad_Gamma}
1010: \frac{\delta \overline{\Gamma}_i}{\bra{\delta v_{\kk n}}}=
1011: \frac{i}{2N^\perp_i}\sum_{\sigma=\pm1}\,\sigma |\widetilde{v}_{\kk i\sigma,n}
1012: \rangle.
1013: \end{equation}
1014: %
1015: Combining Eqs.~(\ref{eq:td_dva}), (\ref{eq:td_dvb}), and
1016: (\ref{eq:grad_Gamma}), and defining
1017: %
1018: \begin{equation}
1019: \label{eq:ket_w_k}
1020: |w_{\kk n}\rangle=\frac{ie}{4\pi}\,\sum_{i=1}^3\,N_i^\parallel
1021: (\ee\cdot\aa_i)
1022: \sum_{\sigma=\pm1}\sigma |\widetilde{v}_{\kk i\sigma,n}\rangle,
1023: \end{equation}
1024: %
1025: the dynamical equation becomes
1026: %
1027: \begin{equation}
1028: \label{eq:td_discrete_a}
1029: i\hbar\frac{d}{dt}|v_{\kk n}\rangle=\hat{H}_{\kk}^0|v_{\kk n}\rangle
1030: +|w_{\kk n}\rangle.
1031: \end{equation}
1032: 
1033: Eq.~(\ref{eq:td_discrete_a}) is a discretized version of  
1034: Eq.~(\ref{eq:tdse_continuum_cov}), i.e.,
1035: %
1036: \begin{equation}
1037: \label{eq:cov_dis}
1038: \ket{w_{\kk n}}\simeq ie\ee\cdot\widetilde{\partial}_\kk\ket{v_{\kk n}}.
1039: \end{equation}
1040: %
1041: This 
1042: is connected with the fact that
1043: the duals provide a natural framework for writing a finite-difference 
1044: representation of $\widetilde{\partial}_\kk\ket{v_{\kk n}}$.\cite{sai02}
1045: In our notation,
1046: %
1047: \begin{equation}
1048: \label{eq:cov_discrete}
1049: \Delta \kk_i \cdot \widetilde{\partial}_\kk| v_{\kk n}\rangle\simeq
1050: \frac{1}{2}\sum_{\sigma=\pm1}\,\sigma|\widetilde{v}_{\kk i\sigma,n}
1051: \rangle.
1052: \end{equation}
1053: 
1054: Both dynamical equations
1055: (\ref{eq:tdse_continuum}) and (\ref{eq:tdse_continuum_cov}) lead to
1056: $d\langle v_{\kk n}|v_{\kk m}\rangle/dt=0$ for WR manifolds, so that 
1057: the time evolution of the individual states $|v_{\kk n}\rangle$
1058: is unitary. This property is 
1059: preserved in the discretized form (\ref{eq:td_discrete_a}), 
1060: since
1061: $\langle v_{\kk n}|w_{\kk m}\rangle=0$.
1062: In order to take advantage of certain
1063: unitary integration algorithms, it is useful to recast the term
1064: $|w_{\kk n}\rangle$ on the right-hand-side 
1065: as an hermitian operator acting on $|v_{\kk n}\rangle$.
1066: For that purpose let us define
1067: %
1068: \begin{equation}
1069: \label{eq:pkis}
1070: \hat{P}_{\kk i\sigma}=\sum_{n=1}^M |\widetilde{v}_{\kk i\sigma,n}\rangle
1071: \langle v_{\kk n}|,
1072: \end{equation}
1073: %
1074: which converts an occupied state at $\kk$ into its dual at $\kk i\sigma$ and
1075: is invariant under gauge transformations (i.e., under independent unitary
1076: rotations among occupied states at both $\kk$ and $\kk i\sigma$).
1077: It follows that the operator
1078: %
1079: \begin{equation}
1080: \label{eq:wkhat}
1081: \hat{\rm w}_{\kk}(\ee)=
1082: \frac{ie}{4\pi}\sum_{i=1}^3\,N_i^\parallel(\ee\cdot\aa_i)
1083: \sum_{\sigma}\,\sigma\hat{P}_{\kk i\sigma}
1084: \end{equation}
1085: %
1086: turns $|v_{\kk n}\rangle$ into $|w_{\kk n}\rangle$, which is the property
1087: we seek. Lastly, for the purpose of acting on $|v_{\kk n}\rangle$ the
1088: nonhermitian $\hat{\rm w}_{\kk}$
1089: can be replaced by 
1090: $\hat{\rm w}_{\kk}+\hat{\rm w}_{\kk}^{\dagger}$ since 
1091: $\hat{P}_\kk \hat{P}_{\kk i \sigma}=\hat{P}_\kk$, so that
1092: $\hat{Q}_\kk\hat{\rm w}_{\kk}=\hat{\rm w}_{\kk}$ 
1093: (where $\hat{Q}_\kk=1-\hat{P}_\kk$)
1094: and therefore
1095: $\hat{\rm w}_{\kk}^{\dagger}|v_{\kk n}\rangle=0$.
1096: We have thus achieved our goal: Eq.~(\ref{eq:td_discrete_a}) now takes
1097: the canonical form of a TDSE,
1098: %
1099: \begin{equation}
1100: \label{eq:td_discrete_b}
1101: i\hbar\frac{d}{dt}|v_{\kk n}\rangle=\tk|v_{\kk n}\rangle,
1102: \end{equation}
1103: %
1104: with an hermitian operator on the right-hand side:
1105: %
1106: \begin{equation}
1107: \label{eq:t_k}
1108: \tk(\ee)=\hat{H}_{\kk}^0+\hat{\rm w}_{\kk}(\ee)+
1109: \hat{\rm w}_{\kk}^{\dagger}(\ee).
1110: \end{equation}
1111: %
1112: We remarked previously that in the continuum-$k$ limit 
1113: Eq.~(\ref{eq:td_discrete_a}) reduces to Eq.~(\ref{eq:tdse_continuum_cov}).
1114: The corresponding analysis for Eqs.~(\ref{eq:td_discrete_b})-(\ref{eq:t_k}) is
1115: left to Appendix~\ref{app:cov}.
1116: 
1117: The operator $\hat{\rm w}_{\kk}$
1118: appearing in Eq.~(\ref{eq:t_k}) is defined via 
1119: Eqs.~(\ref{eq:def-vtkis}), (\ref{eq:pkis}), and (\ref{eq:wkhat}).  It
1120: should be emphasized that it depends explicitly on the occupied
1121: states at $\kk$ and $\kk i\sigma$.
1122: In particular, even when $\hat{H}^0_\kk$ and $\ee$ are time-independent,
1123: if the occupied manifolds at $\kk$ and $\kk i\sigma$ are changing over time, so
1124: is $\tk$. However, $\tk$ remains invariant under unitary rotations
1125: at $k$-points $\kk$ and $\kk i\sigma$.
1126: Hence the resulting dynamics of
1127: the occupied manifold (Eq.~(\ref{eq:td_projector}) below) has the essential
1128: property of being insensitive to the gauge arbitrariness that is always present 
1129: in numerical simulations.
1130: 
1131: Note that
1132: when the Lagrangian procedure was applied in Sec.~\ref{sec:td_continuum} to
1133: the continuum-$k$ problem,
1134: we arrived at Eq.~(\ref{eq:tdse_continuum}), which contains 
1135: $\partial_\kk\ket{v_{\kk n}}$. When the same was done after discretization, the
1136: resulting dynamical equation contained instead 
1137: $\widetilde{\partial}_\kk\ket{v_{\kk n}}$. The 
1138: reason is that the gradient of the discretized Berry's phase, 
1139: Eq.~(\ref{eq:grad_Gamma}), is by construction orthogonal to the 
1140: occupied subspace at $\kk$,
1141: whereas the corresponding continuum-$k$ term used in 
1142: Sec.~\ref{sec:td_continuum}
1143: was not. Had we orthogonalized that term, we
1144: would have obtained $\hat{Q}_\kk\partial_\kk\ket{v_{\kk n}}$ instead of
1145: $\partial_\kk\ket{v_{\kk n}}$, which is
1146: equivalent to $\widetilde{\partial}_\kk\ket{v_{\kk n}}$
1147: (see Appendix~\ref{app:cov}).
1148: 
1149: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1150: 
1151: \subsection{Numerical time integration}
1152: \label{sec:integration}
1153: 
1154: In the applications of Sec.~\ref{sec:results} we use the
1155: algorithm\cite{koonin90}
1156: 
1157: \begin{equation}
1158: \label{eq:time_evolution}
1159: \ket{v_{\kk n}(t+\Delta t)}= \frac{1-i\hbar(\Delta t/2)\hat{T}_{\kk}(t)}
1160: {1+i\hbar(\Delta t/2)\hat{T}_{\kk}(t)}\,
1161: \ket{v_{\kk n}(t)}
1162: \end{equation}
1163: %
1164: to perform the time integration.
1165: Note that in order to use this algorithm it was necessary to invoke
1166: the form (\ref{eq:td_discrete_b}) of the TDSE.
1167: The hermiticity of $\hat{T}_{\kk}$ guarantees that the time evolution is
1168: strictly unitary for any value of $\Delta t$.
1169: Since the system under study in Sec.~\ref{sec:results} is a tight-binding model
1170: with only three basis orbitals, the matrix inversion
1171: is very inexpensive. 
1172: The same algorithm has been successfully used to perform
1173: self-consistent time-dependent density-functional calculations of the
1174: optical properties of atomic clusters using localized  
1175: orbitals as a basis set.\cite{tsolakidis02}
1176: For calculations with large basis sets
1177: (e.g., plane-waves) more efficient algorithms are 
1178: available.\cite{sugino99,watanabe02}
1179: 
1180: Owing to the hermiticity of $\tk$, the projector
1181: (\ref{eq:projector}) obeys
1182: %
1183: \begin{equation}
1184: \label{eq:td_projector}
1185: \frac{d\hat{P}_\kk}{dt}=\frac{1}{i\hbar}\,[\tk,\hat{P}_\kk].
1186: \end{equation}
1187: %
1188: Hence, a variation of the above approach would be to replace
1189: $\tk$ by 
1190: %
1191: \begin{equation}
1192: \hat{\cal T}_\kk=\hat{Q}_\kk \tk + \tk \hat{Q}_\kk,
1193: \end{equation}
1194: %
1195: which is also hermitian. Because
1196: $[\hat{\cal T}_\kk,\hat{P}_\kk]=[\tk,\hat{P}_\kk]$,
1197: this choice does not change the dynamics of the occupied subspace
1198: $\hat{P}_\kk$, but it does change the dynamics of the individual states
1199: $|v_{\kk n}\rangle$.  In fact, $\hat{\cal T}_\kk$ generates
1200: a {\it parallel transport} evolution characterized by
1201: $\langle v_{\kk n}|\dot{v}_{\kk n}\rangle=0$, thus discarding the
1202: ``irrelevant'' part of the dynamics associated with phase factors and
1203: unitary rotations inside the occupied subspace.
1204: We have found empirically, however, that the use of $\hat{\cal T}_\kk$ in 
1205: Eq.~(\ref{eq:time_evolution}) appears to result in a less stable
1206: numerical time evolution, and we have therefore chosen to retain the
1207: original $\tk$ dynamics in our practical implementation.
1208: 
1209: % ----------------------------------------------------------------------------
1210: 
1211: \subsection{Discussion}
1212: \label{sec:dynamics_discussion}
1213: 
1214: It may seem surprising that a linear potential can be accommodated in a
1215: theoretical description of a periodic bulk system.
1216: A commonly-held viewpoint is that  a linear potential can be
1217: implemented within periodic boundary conditions only for the case
1218: of a finite system (molecule or cluster) in a supercell, in which
1219: case it becomes possible to introduce a sawtooth potential as long as its
1220: discontinuity is located in a region of negligible electron density.
1221: To the contrary, Eqs.~(\ref{eq:tdse_continuum}), 
1222: (\ref{eq:tdse_continuum_cov}), and (\ref{eq:td_discrete_b}) 
1223: demonstrate that it is perfectly permissible to insist on the usual periodic
1224: boundary conditions on the wave functions while allowing for
1225: nonperiodicity of the potential. This can be done because
1226: the potential takes the special form of a sum of spatially periodic and
1227: linear contributions, relevant to a crystal in a homogeneous
1228: electric field.  As shown in Sec.~\ref{sec:periodicity}, the 
1229: action of the nonperiodic Hamiltonian, Eq.~(\ref{eq:hamiltonian}), then
1230: preserves the lattice periodicity of the density matrix,
1231: which can therefore be represented by periodic wave functions.
1232: 
1233: Incidentally, we note that 
1234: a sawtooth operator of sorts is ``hiding'' behind the TBP
1235: formulas. The Berry-phase polarization has been recast 
1236: as the expectation value of a properly-defined center-of-mass 
1237: position operator of the many-electron periodic system.\cite{souza00} That 
1238: operator, 
1239: introduced by Kohn,\cite{kohn64} is a sawtooth, not in real space but
1240: in the configuration space of the many-body wave function. 
1241: It can only be constructed for wave functions having a certain
1242: disconnectedness (localization) property in configuration space
1243: characteristic of the insulating state.
1244: This observation is closely related to the discussion in
1245: Sec.~\ref{sec:dyn_pol_disc}.
1246: 
1247: Finally, we mention that an alternative approach for introducing a linear 
1248: potential into a periodic solid
1249: is via the crystal momentum representation (CMR) formalism.\cite{blount62}  
1250: This approach is summarized in
1251: Appendix~A, where the connection with our formalism is established.
1252: The CMR dynamical equations appear to be
1253: less convenient for computational work.
1254: However, the advantages of the present formulation
1255: came at the expense of generality,
1256: since our equations are restricted to the scattering-free dynamics of
1257: initially insulating systems.
1258: 
1259: % ----------------------------------------------------------------------------
1260: 
1261: \section{Stable Stationary solutions}
1262: \label{sec:stationary}
1263: 
1264: \subsection{Formulation}
1265: \label{sec:stat_formulation}
1266: 
1267: Let us try to find, for a constant $\ee\not=0$, 
1268: solutions to Eq.~(\ref{eq:td_discrete_b}) 
1269: for which the occupied manifold remains unchanged over time. A natural 
1270: guess is the manifold spanned by $M$ eigenstates of $\tk$ at each 
1271: $\kk$,
1272: %
1273: \begin{equation}
1274: \label{eq:eigenequation_tk}
1275: \tk|v_{\kk n}\rangle=E_{\kk n}(\ee)|v_{\kk n}\rangle.
1276: \end{equation}
1277: %
1278: Since $\tk$ depends on the occupied states at the neighboring $k$-points, 
1279: Eq.~(\ref{eq:eigenequation_tk})  must be solved 
1280: self-consistently among all $\kk$. If a solution exists,
1281: the corresponding
1282: $\hat{T}_{\kk}$ and $\hat{P}_{\kk}$ commute and, according to
1283: Eq.~(\ref{eq:td_projector}), $d\hat{P}_{\kk}/dt=0$, i.e., the solution is stationary.
1284: 
1285: We are now ready to make contact with Refs.~\onlinecite{souza02,umari02}, where
1286: the energy functional $E$ of Nunes and Gonze,\cite{nunes01} 
1287: Eq.~(\ref{eq:energy}), was minimized at fixed
1288: $\ee$. A stationary point of $E$ has zero gradient: 
1289: %
1290: $|G_{\kk n}\rangle=\delta E/\langle \delta v_{\kk n}|=0$,
1291: where the functional derivative is taken in such a way that the gradient
1292: is orthogonal to the occupied space.
1293: In Appendix~\ref{app:grad-b} it is shown that
1294: %
1295: \begin{equation}
1296: \label{eq:gradient_energy}
1297: |G_{\kk n}\rangle=
1298: (1/N)\,\hat{Q}_\kk\tk\,|v_{\kk n}\rangle,
1299: \end{equation}
1300: %
1301: so that solutions of Eq.~(\ref{eq:eigenequation_tk}) obey
1302: $|G_{\kk n}\rangle=0$.  Thus, stationary solutions
1303: of the dynamical equation are
1304: stationary points of $E$. 
1305: 
1306: A Hessian stability analysis\cite{unpublished} 
1307: shows that a necessary condition for a stationary
1308: point of $E$ to be a minimum is that the $M$ lowest-lying eigenstates of 
1309: $\tk$ are chosen.
1310: Since doing so at $\ee=0$  yields the 
1311: ground state, at finite $\ee$ that procedure yields a state that is
1312: is adiabatically connected to it by slowly ramping up
1313: the field, keeping the system in a minimum of $E$.
1314: Such ``polarized manifolds'' have been 
1315: discussed previously in a perturbative 
1316: framework,\cite{nenciu91,wannier55,adams57}
1317: treating $\kk$ as a continuous variable. In that limit the electric 
1318: field perturbation becomes singular.
1319: That is, even an arbitrary small field induces a 
1320: current via Zener tunneling to higher bands, and the polarized manifolds
1321: are not stationary, but rather, are long-lived resonances.
1322: In other words, an infinite crystal in the presence of a static electric field
1323: does not have a ground state. This is reflected in $E$ loosing its 
1324: minima as soon as $\ee$ departs from zero. 
1325: 
1326: Instead, for a discrete mesh of $k$-points, arguments can be 
1327: given\cite{nunes01,souza02} suggesting that $E$ loses its minima only when 
1328: $\ee$ exceeds a critical value ${\cal E}_c(N)$ that decreases as 
1329: the number $N$ of $k$ points increases; this is supported by numerical 
1330: calculations.\cite{souza02,umari02} It follows from the preceding discussion 
1331: that the minima of $E$ below ${\cal E}_c(N)$ are stable
1332: stationary solutions of the dynamical equation.
1333: Conversely, above ${\cal E}_c(N)$ there are no 
1334: such solutions. 
1335: 
1336: These two regimes --~below and above the
1337: critical field~-- will be explored numerically in 
1338: Sec.~\ref{sec:efield_turn_on} via time-dependent calculations. If one stays
1339: below ${\cal E}_c(N)$, the stationary solutions can be computed
1340: using time-independent methods, such as the diagonalization algorithm
1341: described next or the minimization methods of 
1342: Refs.~\onlinecite{souza02,umari02}.
1343: 
1344: % ----------------------------------------------------------------------------
1345: 
1346: \subsection{Diagonalization algorithm}
1347: \label{sec:algebraic}
1348: 
1349: We have in Eq.~(\ref{eq:eigenequation_tk}) the basis for an algebraic method 
1350: of computing stationary states at finite $\ee$ on a uniform
1351: $k$-point mesh, for $|\ee|<{\cal E}_c(N)$: loop over the $k$ points; for 
1352: each one select the $M$ eigenstates of $\tk$ with the
1353: lowest eigenvalues; iterate until the procedure
1354: converges at all $\kk$ and the occupied subspace stabilizes (this will only
1355: happen below ${\cal E}_c$). 
1356: Even in a tight-binding model without charge self-consistency, the set of
1357: equations (\ref{eq:eigenequation_tk}) has to be solved self-consistently
1358: throughout the BZ,
1359: since the operators $\tk$ couple neighboring $k$ points via their
1360: dependence on the $\ket{v_{\kk n}}$. One may choose to update 
1361: $\tk$ either inside or outside the loop over $\kk$; the latter
1362: option renders the algorithm parallelizable over $k$ points.
1363: 
1364: We have tested this scheme on the tight-binding model of 
1365: Sec.~\ref{sec:results}, and confirm that it produces the same state 
1366: as a direct steepest-descent or conjugate-gradients minimization of 
1367: the functional $E$.\cite{souza02}
1368: This algorithm may be especially suited for implementation in certain 
1369: total-energy codes
1370: that are based on iteratively diagonalizing the
1371: Kohn-Sham Hamiltonian expanded in a small basis set of local 
1372: orbitals.\cite{soler02}
1373: % ----------------------------------------------------------------------------
1374: 
1375: \subsection{Discussion}
1376: \label{sec:stat_discussion}
1377: 
1378: Eq.~(\ref{eq:eigenequation_tk}) is a discretization of the time-independent
1379: version of Eq.~(\ref{eq:tdse_continuum_cov}):
1380: %
1381: \begin{equation}
1382: \label{eq:stationary_continuum_cov}
1383: \big(\hat{H}^0_\kk+
1384: ie{\ee}\cdot\widetilde{\partial}_\kk\big)\ket{{v}_{\kk n}}=
1385: E_{\kk n}(\ee)\ket{{v}_{\kk n}}.
1386: \end{equation}
1387: %
1388: An analysis of the eigenvalues of this equation will serve as a guide for 
1389: discussing those of Eq.~(\ref{eq:eigenequation_tk}).
1390: (For the present purposes we will assume that the continuum form 
1391: (\ref{eq:stationary_continuum_cov})
1392: has solutions for $\ee\not=0$.) As a result of the properties of the
1393: covariant derivative (Appendix~\ref{app:cov}) the $E_{\kk n}(\ee)$ are 
1394: invariant under diagonal gauge transformations
1395: $U_{\kk,mn}=e^{i\theta_{\kk m}}\delta_{m,n}$. 
1396: Upon multiplying on the left by
1397: $\bra{v_{\kk n}}$ the second term on the left-hand-side of
1398: Eq.~(\ref{eq:stationary_continuum_cov}) vanishes. Integrating 
1399: over $\kk$ and summing over $n$, we then find
1400: %
1401: \begin{equation}
1402: \label{eq:inequality}
1403: \Omega_B^{-1}\sum_{n=1}^M\,\int d\kk\, E_{\kk n}(\ee)=
1404: E^0(\ee)\geq E^0(\ee=0),
1405: \end{equation}
1406: %
1407: where $E^0(\ee)$ is the zero-field energy functional
1408: (\ref{eq:energy0}) evaluated at the field-polarized stationary state, and the 
1409: inequality follows from the variational principle. 
1410: The same properties hold 
1411: for the eigenvalues of the discretized form~(\ref{eq:eigenequation_tk}),
1412: which can be obtained by diagonalizing $\hat{H}_\kk^0$ inside the occupied
1413: manifold.
1414: We have here the interesting situation that a minimum of the
1415: {\it total} energy $E$ can be obtained by solving the eigenvalue equations
1416: (\ref{eq:eigenequation_tk}) whose eigenvalues, summed over $n$ and $\kk$
1417: give instead the {\it zero-field} contribution $E^0$. This can be traced 
1418: back to Eq.~(\ref{eq:orthogonal}), which expresses the
1419: ``parallel-transport-like'' nature of the covariant derivative.
1420: 
1421: The above is to be compared with the time-independent
1422: version of Eq.~(\ref{eq:tdse_continuum}),
1423: %
1424: \begin{equation}
1425: \label{eq:stationary_continuum}
1426: \big(\hat{H}^0_\kk+
1427: ie{\ee}\cdot\partial_\kk\big)\ket{{v}'_{\kk n}}=
1428: E'_{\kk n}(\ee)\ket{{v}'_{\kk n}}.
1429: \end{equation}
1430: %
1431: Under a diagonal transformation
1432: $\ket{{v}'_{\kk n}}\rightarrow e^{i\theta_{\kk m}}\,\ket{{v}'_{\kk n}}$
1433: its eigenvalues change as
1434: $E'_{\kk n}(\ee) \rightarrow E'_{\kk n}(\ee)-e\ee\cdot\partial_\kk
1435: \theta_{\kk n}$. The analog of Eq.~(\ref{eq:inequality})
1436: is
1437: %
1438: \begin{equation}
1439: \Omega_B^{-1}\sum_{n=1}^M\,\int d\kk\, E'_{\kk n}(\ee)=
1440: E^0(\ee)-v\ee\cdot\pp(\ee).
1441: \end{equation}
1442: %
1443: The quantity on the right-hand side is now the total energy $E$,
1444: which is invariant only modulo $e\ee\cdot{\bf R}$.
1445: 
1446: Although the individual eigenstates of Eq.~(\ref{eq:stationary_continuum_cov})
1447: are in general different from those of Eq.~(\ref{eq:stationary_continuum}),
1448: the self-consistent solutions for all $\kk$ and $n$ span the same
1449: space in both cases, i.e., they differ only by a gauge transformation.
1450: It is then a matter of convenience to choose which of the two
1451: equations to solve in practice. 
1452: Our particular approach is to discretize
1453: Eq.~(\ref{eq:stationary_continuum_cov}) 
1454: in a gauge-covariant manner, and then solve the resulting 
1455: Eq.~(\ref{eq:eigenequation_tk}).
1456: 
1457: % ----------------------------------------------------------------------------
1458: 
1459: \section{Numerical results}
1460: \label{sec:results}
1461: 
1462: % ----------------------------------------------------------------------------
1463: 
1464: \subsection{Tight-binding model}
1465: \label{sec:tb}
1466: 
1467: We have applied our scheme to the one-dimensional tight-binding model 
1468: of Ref.~\onlinecite{nunes94}, a three-band Hamiltonian with three 
1469: atoms per unit cell of length $a=1$ and one orbital per atom,
1470: %
1471: \begin{equation}
1472: \label{eq:tb_ham}
1473: \hat{H}^0(\alpha)=\sum_j\,
1474: \big\{
1475:   \epsilon_j(\alpha)\hat{c}_j^\dagger \hat{c}_j+
1476:   t\big[\hat{c}_j^\dagger \hat{c}_{j+1}+\rm{h.c.}\big]
1477: \big\},
1478: \end{equation}
1479: %
1480: with the site energy given by $\epsilon_{3m+l}(\alpha)=
1481: \Delta\cos(\alpha-\beta_l)$. Here $m$ is the cell index, 
1482: $l=\{ -1,0,1\}$ is the site index, and $\beta_l=2\pi l/3$.
1483: Before the
1484: Berry-phase polarization can be computed\cite{bennetto96} (or an electric field
1485: applied to the system\cite{nunes94,dunlap86}), the position operator must
1486: be specified.
1487: Although this may be done without introducing additional 
1488: parameters,\cite{foreman02} we adopt the simple prescription
1489: of Ref.~\onlinecite{nunes94}:
1490: $\hat{x}=\sum_j\, x_j\hat{c}_j^\dagger \hat{c}_j$, with $x_j=j/3$.
1491: In the results reported below we have set $e=\hbar=1$, $t=1$, and $\Delta=-1$,
1492: and only the lowest band is filled (with single occupancy). 
1493: Fig.~\ref{fig1} shows the band structure at zero field for $\alpha=0$.
1494: %
1495: \begin{figure}
1496: \centerline{\epsfig{file=fig1.ps,width=3.3in,angle=0}}
1497: \caption{Energy dispersion of the tight-binding model for the choice of
1498: parameters $t=1$, $\Delta=-1$, and $\alpha=0$.}
1499: \label{fig1}
1500: \end{figure}
1501: 
1502: % ----------------------------------------------------------------------------
1503: 
1504: \subsection{Sliding charge-density wave}
1505: \label{sec:cdw}
1506: 
1507: The Hamiltonian of Eq.~(\ref{eq:tb_ham}) is a simple model of a
1508: commensurate charge-density wave which slides by one period as the parameter
1509: $\alpha$ evolves adiabatically through $2\pi$.
1510: It is easiest to see this by noting that in the space of parameters
1511: $\Delta_x=\Delta\cos\alpha$ and $\Delta_y=\Delta\sin\alpha$, cycling
1512: $\alpha$ by $2\pi$ corresponds to tracing a circle
1513: about the origin in the $\Delta_x$--$\Delta_y$ plane.  The system
1514: is insulating (i.e., a gap remains open) at all points in this plane
1515: except for a singular point at the origin where the system is metallic.
1516: Thus, this cyclic adiabatic change in $\hat{H}^0$ takes the 
1517: system along an insulating path that encircles this singular point,
1518: so that a quantized particle transport $\Delta P=\int_0^T
1519: J(t)\,dt$ of a unit charge is obtained.\cite{thouless83}
1520: 
1521: Away from the adiabatic regime, deviations from exact quantization are 
1522: expected.
1523: This can be understood from the fact that under nonadiabatic conditions 
1524: the state at time $t$ depends on the history at times $t'<t$.
1525: In particular, the final state may be be 
1526: different from the initial one even though $\hat{H}^0(T)=\hat{H}^0(0)$, in
1527: which case $P(T)-P(0)\not=1$.
1528: By contrast, an adiabatically-evolving system has no memory, being
1529: completely determined by the instantaneous $\hat{H}^0(t)$ and $d\hat{H}^0/dt$.
1530: 
1531: To illustrate this point, we increased $\alpha$ from $0$ to $2\pi$ 
1532: during a time interval  $t\in[0,T]$ according to
1533: $\alpha(t)=2\pi \sin^2(\pi t/2T)$, and held it
1534: constant afterwards. The system was prepared at $t=0$ in its ground state, and 
1535: the wave functions evolved in time 
1536: according to Eq.~(\ref{eq:time_evolution}), using $\Delta t=0.005$ (the same
1537: time step was used in all other simulations in this work). 
1538: At each time step we computed the dynamic 
1539: polarization using Eq.~(\ref{eq:discrete_phase}).
1540: %
1541: \begin{figure}
1542: \centerline{\epsfig{file=fig2.ps,width=3.3in,angle=0}}
1543: \vspace{0.2cm}
1544: \caption{Time evolution of the polarization as a result of changing the sliding
1545: parameter $\alpha$ from $0$ to $2\pi$ over the time interval $[0,80]$, using
1546: 200 $k$ points. The 
1547: solid line shows the actual dynamic polarization, while the dashed line shows 
1548: the ground-state polarization of the instantaneous Hamiltonian.
1549: Inset: Detail of the remnant oscillations of the polarization after the
1550: Hamiltonian stops changing, at $t=80$. For comparison, the results using
1551: 100 $k$ points are also shown.}
1552: \label{fig2}
1553: \end{figure}
1554: 
1555: The resulting $P(t)$ for $T=80$ and 200 $k$ points is shown in Fig.~\ref{fig2},
1556: where we also display the exact adiabatic ($T\rightarrow\infty$) limit
1557: $P_{\rm static}[\alpha(t)]$ obtained by diagonalizing 
1558: $\hat{H}^0_k(\alpha(t))$ on the same mesh of $k$ points.\cite{foot:thouless}
1559: (To check that our calculations are converged with respect to the number of 
1560: $k$ points, the inset of
1561: Fig.~\ref{fig2} compares the results for 100 and 200 $k$ points.)
1562: The dynamic polarization $P(t)$ obtained by solving the TDSE follows closely, 
1563: but not exactly, the adiabatic curve. In particular, at the time $t=80$ when 
1564: the Hamiltonian stops changing, the polarization differs 
1565: slightly from unity (see inset in Fig.~\ref{fig2}), indicating that the system is
1566: not in the ground state. The oscillations that
1567: follow arise from quantum interference
1568: (beats) between valence and conduction states, as a 
1569: result of having excited electrons across the gap during $[0,T]$. Their period,
1570: of 5.5 time units, corresponds to the fundamental gap in 
1571: Fig.~\ref{fig1}, $E_{\rm gap}=1.137$. This is consistent with
1572: the $k$-space distribution of the (small) electron-hole pair
1573: amplitude present in the system after time $T$: for a sliding period of
1574: $T=80$, the distribution 
1575: is mostly 
1576: concentrated around $k=0$, and it is essentially the lowest conduction band 
1577: that gets populated. 
1578: As the adiabatic limit is approached by increasing $T$, the amplitude of the
1579: remnant oscillations of the polarization decreases. This is illustrated in
1580: the upper panel of Fig.~\ref{fig3}, where we compare $T=80$ with $T=120$.
1581: 
1582: Besides the macroscopic polarization $P(t)$, another quantity of interest
1583: is the electronic localization length $\xi(t)$ that 
1584: characterizes the root-mean-square quantum fluctuations of the macroscopic
1585: polarization.\cite{souza00} It is given
1586: by $\xi^2=\Omega_{\rm I}/M$, where $\Omega_{\rm I}$ is a
1587: gauge-invariant quantity which in one dimension is equal to the
1588: spread of the maximally-localized Wannier functions
1589: (\ref{eq:wannier_like}).\cite{mv97} We have computed $\Omega_{\rm I}$ using 
1590: Eq.~(34) from Ref.~\onlinecite{mv97}, and in the lower panel of Fig.~\ref{fig3}
1591: we plot $\xi(t)$ against $\alpha(t)/2\pi$.
1592: In the adiabatic limit the resulting curve consists of three identical 
1593: oscillations, reflecting the existence of three equivalent atoms in the unit 
1594: cell. As nonadiabaticity 
1595: increases, $\xi$ tends to increase as well. Nevertheless, the electrons remain 
1596: localized, i.e., ``insulating-like'', in the sense discussed in
1597: Sec.~\ref{sec:dyn_pol_disc}.\cite{foot:insulating-like}
1598: %
1599: \begin{figure}
1600: \centerline{\epsfig{file=fig3.ps,width=3.3in,angle=0}}
1601: \vspace{0.2cm}
1602: \caption{Upper panel: Same as Fig.~\ref{fig2}, but now using two different
1603: time intervals, $[0,80]$ and $[0,120]$, for changing the sliding parameter
1604: $\alpha$.
1605:  Lower panel: electron localization 
1606: length $\xi(t)$ versus the instantaneous value of $\alpha$, during the
1607: time intervals $[0,T]$ during which $\alpha$ is changing.}
1608: \label{fig3}
1609: \end{figure}
1610: 
1611: The above results are representative of the regime where deviations
1612: from adiabaticity are small. If we increase the degree of
1613: nonadiabaticity by choosing a smaller $T$ (e.g., $T=40$), we begin
1614: to notice a linear increase of the polarization at later
1615: times.  This new behavior can be traced to the excitation of
1616: electron and hole wavepackets centered at some $k_0$ and
1617: propagating at different group velocities.  Let $\Delta E_{k_0}$ be
1618: the interband separation, and $\Delta v_g$ be the difference of
1619: group velocities of the two lowest bands, at $k_0$.  Then, in addition to
1620: the quantum beats of period $2\pi\hbar/\Delta E_{k_0}$ caused by
1621: the interband dynamics, we observe a linear-in-$t$ term in $P(t)$
1622: with slope proportional to $\Delta v_g$, reflecting the change in
1623: dipole moment as the electron-hole pair separates.  (More
1624: precisely, the preceeding statements apply only in the limit of a
1625: dense $k$-point mesh; for any finite mesh spacing $\Delta k$, the
1626: linear behavior is replaced by an oscillatory one with an amplitude
1627: scaling as $1/\Delta k$ and period $2\pi/(\Delta v_g\,\Delta k)$. 
1628: Thus, an especially fine $k$-point mesh should be used 
1629: if these effects are to be investigated.) 
1630: 
1631: 
1632: 
1633: % ----------------------------------------------------------------------------
1634: 
1635: \subsection{Gradual turn-on of an electric field}
1636: \label{sec:efield_turn_on}
1637: 
1638: In the previous example the electric field was held at zero,
1639: and the dynamics was
1640: produced by varying the parameter $\alpha$ in $\hat{H}^0$. Let us now study the
1641: polarization response of the system when an electric field ${\cal E}(t)$  is 
1642: switched on linearly over a time interval $[0,T]$ and is held fixed afterwards.
1643: We have set $\alpha=0$, so that the ground state is centrosymmetric, with zero 
1644: spontaneous polarization.
1645: 
1646: We begin by considering a situation where the final 
1647: value of the field, ${\cal E}_{\rm max}$, is smaller than the 
1648: $k$-mesh-dependent critical field ${\cal E}_c(N)$
1649: above which the energy functional (\ref{eq:energy}) has no
1650: minima. This allows us to compare the dynamic polarization $P(t)$ with the
1651: static polarization $P_{\rm static}[{\cal E}(t)]$ of the stationary state in 
1652: the presence of the same field, which we find by minimizing the 
1653: energy.\cite{souza02} In Fig.~\ref{fig4} we display the results for 
1654: ${\cal E}_{\rm max}=0.025$ and two
1655: different switching times, $T=40$ and $T=80$. The simulation was done using 200
1656: $k$ points, to which corresponds a critical field 
1657: ${\cal E}_c(N=200)\sim 0.037$.
1658: (The inset shows the agreement between the results obtained
1659: using 100 and 200 $k$ points.)
1660: Clearly, $P(t)$ tracks quite closely the adiabatic curve 
1661: $P_{\rm static}[{\cal E}(t)]$, the more so as $T$ increases. This 
1662: illustrates the point, 
1663: emphasized in Ref.~\onlinecite{nunes94}, that the
1664: state obtained by minimizing a field-dependent energy functional should be 
1665: thought of as
1666: the one which is generated from the zero-field state by adiabatically turning
1667: on ${\cal E}$. 
1668: %
1669: \begin{figure}
1670: \centerline{\epsfig{file=fig4.ps,width=3.3in,angle=0}}
1671: %\vspace{-0.4cm}
1672: \caption{Time evolution of the polarization as a result of increasing the 
1673: electric field from $0$ to ${\cal E}_{\rm max}=0.025$ over two time intervals,
1674: $[0,40]$ and $[0,80]$, using 200 $k$ points.
1675: The solid line shows the actual dynamic polarization, while the dashed line 
1676: shows the static polarization for the instantaneous value of the field.
1677: Inset: Comparison of the dynamic polarization for 100 and 200 $k$ points.}
1678: \label{fig4}
1679: \end{figure}
1680: 
1681: Let us now explore the regime above ${\cal E}_c(N)$, where energy-minimization 
1682: schemes fail. For ${\cal E}_{\rm max}>{\cal E}_c(N)$ the exact adiabatic limit
1683: of the process of ramping up the field is unattainable. Nevertheless, if 
1684: ${\cal E}_{\rm max}$ is small compared to the field scale at which
1685: {\it intrinsic} breakdown occurs (i.e., at which the Zener tunneling
1686: rate becomes on the of order interband frequencies, which is a bulk property
1687: \cite{odwyer73}), a {\it quasistationary} state should be reachable
1688: by turning on the field at a rate that is slow compared to the
1689: usual electronic processes, but fast compared to the characteristic
1690: tunneling rate at the maximum field encountered.  After the ramp-up
1691: is completed, but at times still short compared to the tunneling
1692: rate, this state should provide the appropriate extrapolation to
1693: fields above ${\cal E}_c(N)$ of the truly stationary state that
1694: exists below ${\cal E}_c(N)$.
1695: 
1696: To illustrate this situation, we repeated the calculation with 200 
1697: $k$ points depicted in Fig.~\ref{fig4}, but increasing the maximum field from 
1698: 0.025 to 0.05, somewhat larger than ${\cal E}_c(200)\sim 0.037$. The resulting 
1699: curve for $P(t)$ is very similar
1700: to that in Fig.~\ref{fig4}, without any sign of runaway behavior.
1701: As a more striking example, we show in Fig.~\ref{fig5} the outcome of
1702: calculations with the same final field of ${\cal E}_{\rm max}=0.05$, but with 
1703: even denser sets of 400 and 800 $k$ points. For the latter ${\cal E}_c\sim0.01$,
1704: considerably smaller than ${\cal E}_{\rm max}$, and still there is no sign of
1705: instability. (Note also that the $P(t)$ curve in Fig.~\ref{fig5} --~whose 
1706: vertical scale differs from that in Fig.~\ref{fig4}  by
1707: the same factor of two that exists between the respective values of 
1708: ${\cal E}_{\rm max}$ --~looks almost identical to that in Fig.~\ref{fig4}.)
1709: These results confirm that, as long as we are solving
1710: a {\it time-dependent} Schr\"odinger equation for a given history of
1711: switching on the field, there is no such thing as a $\Delta k$-dependent
1712: critical field; the thermodynamic limit of an infinitely dense $k$-point mesh
1713: is perfectly well defined. The only breakdown behavior that may  
1714: be observed in short time scales
1715: is the physical one that occurs when the applied field is
1716: large enough that the Zener tunneling rate becomes 
1717: significant.\cite{kane59,odwyer73} 
1718: The concept of a $\Delta k$-dependent critical field applies only to the
1719: attemp to obtain solutions in the presence of a static electric field from
1720: an energy variational principle. By going back to the original dynamical 
1721: problem of slowly ramping up the field, we circumvent the difficulties that 
1722: ultimately resulted from trying to treat as a (stable) stationary state what is
1723: really a long-lived resonance. 
1724: %
1725: \begin{figure}
1726: \centerline{\epsfig{file=fig5.ps,width=3.3in,angle=0}}
1727: %\vspace{-0.4cm}
1728: \caption{Same as Fig.~\ref{fig4}, but now using
1729: ${\cal E}_{\rm max}=0.05$ and 800 $k$ points. Since ${\cal E}_{\rm max}$ is
1730: larger than the critical field for this number of $k$ points
1731: (${\cal E}_c\sim 0.01$), no adiabatic curve $P_{\rm static}[{\cal E}(t)]$ is
1732: shown. Inset: Comparison of the dynamic polarization for 400 and 800 $k$ 
1733: points.}
1734: \label{fig5}
1735: \end{figure}
1736: 
1737: 
1738: % ----------------------------------------------------------------------------
1739: 
1740: \subsection{Dielectric function in a static field}
1741: \label{sec:franz_keldysh}
1742: 
1743: There is great interest in modulating the optical properties of crystals and
1744: superlattices by applying static electric fields. An example
1745: of such an electro-optical effect is the modification of the dielectric
1746: function. This is known as the Franz-Keldysh effect, or electroabsorption.
1747: Although it has been extensively studied
1748: in bulk semiconductors,\cite{yu96} quantum wells,\cite{rink89} and 
1749: superlattices,\cite{ando97} we are not aware of any first-principles 
1750: investigations. The present method may provide a route to such calculations.
1751: %
1752: \begin{figure}
1753: \centerline{\epsfig{file=fig6.ps,width=3.3in,angle=0}}
1754: %\vspace{-0.4cm}
1755: \caption{Susceptibility $\chi^{[{\cal E}_0]}(\omega)$ in the presence of a static
1756: field ${\cal E}_0$, for $\alpha=0$ and 100 
1757: $k$ points, using a level-broadening $\delta=0.04$. 
1758: Dotted lines: Kubo formula
1759: result for ${\cal E}_0=0$; solid lines: 
1760: results using our method, for both ${\cal E}_0=0$ and ${\cal E}_0=0.05$.
1761: The latter displays the Franz-Keldysh effect. The inset compares 
1762: the Franz-Keldysh oscillations for two different bias fields, ${\cal E}_0=0.05$
1763: and ${\cal E}_0=0.03$.}
1764: \label{fig6}
1765: \end{figure}
1766: 
1767: We compute the dielectric function in the presence of a static field 
1768: ${\cal E}_0$ as
1769: follows. The system is prepared at $t=0$ in the stationary state polarized
1770: by a field ${\cal E}_0+\Delta{\cal E}$, with $|\Delta{\cal E}|<<|{\cal E}_0|$. 
1771: By using a field of magnitude below the
1772: critical field, we are able to find that state by 
1773: minimizing the energy. 
1774: For $t>0$ we let the system evolve in time in the presence of the target field 
1775: ${\cal E}_0$.
1776: Let $P_{\rm static}[{\cal E}_0]$ be static polarization of the system under the
1777: field ${\cal E}_0$. The polarization response to the step-function
1778: discontinuity in ${\cal E}(t)={\cal E}_0+\Delta {\cal E}\,\theta(-t)$ is
1779: $\Delta P(t)=P(t)-P_{\rm static}[{\cal E}_0]$. 
1780: To obtain the frequency-dependent response we need the Fourier
1781: transform of $\Delta P(t)$ for $t>0$ only:
1782: %
1783: \begin{equation}
1784: \Delta P(\omega)=\int_0^{+\infty}\,\Delta P(t)\,
1785: e^{(i\omega-\delta)t}\,dt,
1786: \end{equation}
1787: %
1788: where a damping factor $\delta$ has been introduced as an approximate way to
1789: account for level broadening.\cite{tsolakidis02}
1790: To linear order in $\Delta {\cal E}$ the susceptibility is
1791: %
1792: \begin{equation}
1793: {\rm Re}\,\chi^{[{\cal E}_0]}(\omega)=
1794: \left. \frac{dP_{\rm static}[{\cal E}]}{d{\cal E}}\right|_{{\cal E}={\cal E}_0} -
1795: \frac{\omega}{\Delta \cal E}\,{\rm Im}\,\Delta P(\omega),
1796: \end{equation}
1797: %
1798: \begin{equation}
1799: {\rm Im}\,\chi^{[{\cal E}_0]}(\omega)=\frac{\omega}{\Delta \cal E}\,
1800: {\rm Re}\,\Delta P(\omega).
1801: \end{equation}
1802: 
1803: 
1804: With this real-time approach the need to perform a summation over 
1805: conduction-band states is circumvented.
1806: Previous real-time, scalar potential
1807: approaches\cite{tsolakidis02,yabana99} were restricted to finite systems,
1808: since it was unclear how to evaluate the dynamic macroscopic polarization of an
1809: extended system. A real-time, vector-potential
1810: scheme valid for bulk systems was proposed in Ref.~\onlinecite{bertsch00}.
1811: 
1812: We validate our method by comparing in Fig.~\ref{fig6}
1813: the ground-state susceptibility 
1814: with the analytic Kubo-formula (sum-over-states)
1815: result, using in both cases the same 
1816: broadening $\delta$ and $k$-point mesh. Also shown in Fig.~\ref{fig6} is the
1817: susceptibility in the presence of a ${\cal E}_0=0.05$ bias field,
1818: displaying the Franz-Keldysh effect: an
1819: absorption tail below the gap caused by photon-assisted tunneling, and 
1820: oscillations above the gap.\cite{yu96} 
1821: The Franz-Keldysh oscillations become more widely spaced with increasing
1822: ${\cal E}_0$. This is illustrated in the inset of 
1823: Fig.~\ref{fig6}, where we compare them for ${\cal E}_0=0.05$ and
1824: ${\cal E}_0=0.03$.
1825: 
1826: % ----------------------------------------------------------------------------
1827: 
1828: \section{Summary}
1829: 
1830: The work of King-Smith and Vanderbilt demonstrated that the bulk electronic
1831: polarization, defined in terms of the current flowing during the
1832: {\it adiabatic} evolution of an insulating system in a {\it vanishing
1833: macroscopic electric field}, could be related to a Berry's phase
1834: defined over the manifold of occupied Bloch states.\cite{ksv93}
1835: We have generalized this result by considering the time evolution of
1836: an initially insulating electron system under the very general
1837: Hamiltonian (\ref{eq:hamiltonian}), where
1838: the lattice-periodic part $\hat{H}^0(t)$ and the homogeneous
1839: electric field $\ee(t)$ may have an arbitrarily strong and
1840: rapid variation in time. In the absence of scattering, we have proved that the
1841: integrated current $\Delta {\bf P}=\int {\bf J}(t)\,dt$ is still
1842: given by the King-Smith--Vanderbilt formula, but written in terms
1843: of the instantaneous Bloch-like solutions of the time-dependent
1844: Schr\"odinger equation.  The coherent dynamic polarization ${\bf P}(t)$ was
1845: interpreted as a nonadiabatic geometric phase.\cite{aa87} These
1846: generalizations of the theory allowed us to justify recent developments in 
1847: which 
1848: the energy functional $E$ of Nunes and Gonze\cite{nunes01} has been used as the
1849: basis for direct DFT calculations of insulators in a static
1850: homogeneous electric field.\cite{souza02,umari02} 
1851: The limitation of those methods to fields of magnitude smaller than a 
1852: $\Delta \kk$-dependent
1853: critical field that vanishes in the thermodynamic limit 
1854: has been removed: we have shown numerically that quasistationary states in 
1855: finite fields exist for arbitrarily dense $k$-point meshes, and can be obtained
1856: by solving the time-dependent Shr\"odinger equation for a slowly-increasing
1857: field.
1858: The present method also provides a
1859: convenient framework for the computation of coherent time-dependent
1860: excitations in insulators.   
1861: As an example, the dielectric function
1862: was calculated for a tight-binding model by considering the response
1863: to a step-function discontinuity in $\ee(t)$, illustrating
1864: effects such as photon-assisted tunneling and Franz-Keldysh oscillations.
1865: A full {\it ab initio} implementation within the framework of time-dependent
1866: density-functional theory should be possible.
1867: 
1868: \begin{acknowledgements}
1869: This work was supported by NSF Grant DMR-0233925 and by DARPA/ONR Grant
1870: N00014-01-1-1061.  We wish to thank M.~H.~Cohen for many useful discussions.
1871: \end{acknowledgements}
1872: 
1873: % ----------------------------------------------------------------------------
1874: 
1875: \appendix
1876: 
1877: \section{Crystal-momentum representation}
1878: \label{app:pol}
1879: 
1880: The 
1881: introduction of linear scalar potentials in crystals is usually discussed in
1882: the language of the crystal-momentum representation (CMR).\cite{blount62}
1883: Instead, we have used the Berry-phase theory of polarization, and the purpose 
1884: of this Appendix is to show how to switch from one to the other. The
1885: CMR uses 
1886: as a basis the eigenstates $\ket{\psi_{\kk m}}$ of $\hat{H}^0$ with eigenvalues
1887: $E_{\kk m}$. In accordance with Eq.~(\ref{eq:norm_conv}) we assume that 
1888: $|\psi_{\kk m}(\rr)|^2$
1889: integrates to one over the unit cell volume $v$. That 
1890: implies\cite{jones73,foot:normalization}
1891: %
1892: \begin{equation}
1893: \label{eq:normalization}
1894: \bra{\psi_{\kk m}}\psi_{\kk' l}\rangle\equiv\int \psi_{\kk m}^*(\rr)
1895: \psi_{\kk' l}(\rr)\,d\rr=\Omega_B\delta(\kk-\kk')\delta_{ml}.
1896: \end{equation}
1897: % 
1898: The CMR expansion of the identity operator is
1899: %
1900: \begin{equation}
1901: \hat{\bf 1}=\Omega_B^{-1}\sum_{m=1}^\infty\int d\kk\,
1902: \ket{\psi_{\kk m}}\bra{\psi_{\kk m}},
1903: \end{equation}
1904: %
1905: so that a general 
1906: one-electron state $\ket{\phi}$ is expanded as
1907: %
1908: \begin{equation}
1909: \label{eq:cmr}
1910: \ket{\phi}=\hat{\bf 1}\ket{\phi}=
1911: \sum_{m=1}^\infty\,\int\,d\kk' f_{\kk' m}\ket{\psi_{\kk' m}},
1912: \end{equation}
1913: %
1914: where $f_{\kk' m}=\Omega_B^{-1} \bra{\psi_{\kk' m}}\phi\rangle$.
1915: For the occupied Bloch-like states $\ket{\phi_{\kk n}}$ in a WR manifold, the 
1916: CMR wave function $f_{\kk' m}(\kk,n)$ takes the form
1917: %
1918: \begin{equation}
1919: f_{\kk' m}(\kk,n)=c_{\kk',nm}\delta(\kk'-\kk) 
1920: \end{equation}
1921: %
1922: with $\sum_{m=1}^\infty |c_{\kk,nm}|^2=1$,
1923: which leads to Eq.~(\ref{eq:linear_comb}).
1924: 
1925: \subsection{Current and the CMR velocity operator}
1926: 
1927: The velocity operator (\ref{eq:velocity_b})
1928: is diagonal in $\kk$ and is conveniently split into a sum of two 
1929: operators, one diagonal and the other off-diagonal in the band
1930: index:\cite{blount62}
1931: %
1932: \begin{equation}
1933: \label{eq:velocity_cmr}
1934: \hat{\bf v}=\hat{\bf v}^{\rm d}+\hat{\bf v}^{\rm od}.
1935: \end{equation}
1936: %
1937: The matrix elements of $\hat{\bf v}^{\rm d}$ are
1938: %
1939: \begin{equation}
1940: \langle \psi_{\kk m}|\hat{\bf v}^{\rm d}|\psi_{\kk' l}\rangle=
1941: \Omega_B\delta(\kk-\kk')\delta_{ml}{\bf v}^{\rm d}_{\kk m},
1942: \end{equation} 
1943: %
1944: where
1945: %
1946: \begin{equation}
1947: \label{eq:vel_intra}
1948: {\bf v}^{\rm d}_{\kk m}
1949: = \frac{1}{\hbar}\,\partial_\kk E_{\kk m}.
1950: \end{equation}
1951: %
1952: The matrix elements of $\hat{\bf v}^{\rm od}$ are
1953: %
1954: \begin{equation}
1955: \langle \psi_{\kk m}|\hat{\bf v}^{\rm od}|\psi_{\kk' l}\rangle=
1956: \Omega_B\delta(\kk-\kk'){\bf v}^{\rm od}_{\kk,ml},
1957: \end{equation}
1958: %
1959: where
1960: %
1961: \begin{equation} 
1962: \label{eq:vel_inter}
1963: {\bf v}^{\rm od}_{\kk,ml}
1964: =\frac{i}{\hbar}\,{\bf X}_{\kk,ml}\,[E_{\kk m}-E_{\kk l}]
1965: \end{equation}
1966: %
1967: and we have defined the hermitian matrix
1968: %
1969: \begin{equation}
1970: \label{eq:x_matrix}
1971: X_{\kk,ml}^\alpha=i\,\langle 
1972:                       u_{\kk m}|\partial_{k_\alpha} u_{\kk l}
1973:                     \rangle,
1974: \end{equation}
1975: %
1976: which is analogous to Eq.~(\ref{eq:connection}) for the $\ket{v_{\kk n}}$.
1977: 
1978: The current, Eq.~(\ref{eq:current}), is split into intraband and interband 
1979: parts,
1980: %
1981: \begin{equation}
1982: \label{eq:total_current}
1983: \jj(t)=\jj_{\rm intra}(t)+\jj_{\rm inter}(t).
1984: \end{equation}
1985: %
1986: Writing the density matrix as
1987: %
1988: \begin{equation}
1989: \label{eq:cmr_dm}
1990: \langle \psi_{\kk m}|\hat{n}|\psi_{\kk' l}\rangle=
1991: \Omega_B\delta(\kk-\kk')n_{\kk,ml},
1992: \end{equation}
1993: %
1994: where
1995: %
1996: \begin{equation}
1997: \label{eq:dm_cmr}
1998: n_{\kk,ml}= 
1999: \sum_{n=1}^M\,c_{\kk,nm}\,{[c_{\kk,nl}]}^{*},
2000: \end{equation}
2001: %
2002: we find
2003: %
2004: \begin{equation}
2005: \label{eq:intra_current}
2006: \jj_{\rm intra}= -\frac{e}{v}\,{\rm Tr}_c\big(\hat{n}\hat{\bf v}^{\rm d}\big)
2007: =\frac{-e}{(2\pi)^3} \, \sum_{m=1}^{\infty}\,\int\,d\kk\,
2008: n_{\kk,mm} \, {\bf v}^{\rm d}_{\kk m}
2009: \end{equation}
2010: %
2011: and
2012: %
2013: \begin{equation}
2014: \label{eq:inter_current}
2015: \jj_{\rm inter}=-\frac{e}{v}\,{\rm Tr}_c\big(\hat{n}\hat{\bf v}^{\rm od}\big)
2016: =\frac{-e}{(2\pi)^3} \sum_{m,l=1}^{\infty} \, \int \, d\kk\,
2017: n_{\kk,ml} {\bf v}^{\rm od}_{\kk,lm}.
2018: \end{equation} 
2019: %
2020: In the above we used the CMR form of Eq.~(\ref{eq:trace}),
2021: %
2022: \begin{equation}
2023: \label{eq:trace_cmr}
2024: {\rm Tr}_c(\hat{\cal O})=\Omega_B^{-1}\sum_{m=1}^\infty\,
2025: \int d\kk\,\frac{1}{N}\bra{\psi_{\kk m}}\hat{\cal O}\ket{\psi_{\kk m}},
2026: \end{equation}
2027: %
2028: where $N$ should be taken to signify 
2029: $\Omega_B\delta(0)$.\cite{foot:normalization}
2030: 
2031: Plugging (\ref{eq:linear_comb}) into (\ref{eq:curr}) yields, after some
2032: manipulations,
2033: Eqs.~(\ref{eq:total_current}), (\ref{eq:intra_current}), and
2034: (\ref{eq:inter_current}), confirming that
2035: the Berry-phase polarization correctly accounts for both 
2036: intraband and interband contributions.
2037: It is instructive to consider some particular cases. The adiabatic current
2038: $\jj=(d\pp/d\lambda)\dot{\lambda}$
2039: discussed in Refs.~\onlinecite{ksv93,thouless83} is purely interband.
2040: If the perturbation is a sinusoidal electric field, the linear 
2041: response is again a purely interband current, while the nonlinear response 
2042: has also an intraband component.\cite{sipe00,lambrecht00}
2043: 
2044: \subsection{Polarization and the CMR position operator}
2045: 
2046: Along the same lines, one can show that the Berry-phase expression for
2047: $\pp$ is consistent with the CMR position operator, which takes the 
2048: form\cite{blount62}
2049: %
2050: \begin{equation}
2051: \label{eq:cmr_x}
2052: \bra{\psi_{\kk m}}\hat{\rr}\ket{\psi_{\kk'l}}=-i\Omega_B\partial_{\kk'}
2053: \delta(\kk'-\kk)\delta_{mn}+\Omega_B\delta(\kk'-\kk){\bf X}_{\kk,ml}.
2054: \end{equation}
2055: %
2056: Combined with Eqs.~(\ref{eq:cmr_dm}) and (\ref{eq:trace_cmr}) this yields
2057: %
2058: \begin{equation}
2059: \pp=-\frac{e}{v}{\rm Tr}_c(\hat{n}\hat{\rr})
2060: =\frac{-e}{(2\pi)^3}\sum_{m,l=1}^\infty\int d\kk\,
2061: n_{\kk,ml}{\bf X}_{\kk,lm},
2062: \end{equation}
2063: %
2064: which is the same results one gets from inserting the CMR expansion
2065: (\ref{eq:linear_comb}) into the nonadiabatic Berry-phase formula 
2066: (\ref{eq:dyn_pol}). The linear character of $\hat{\rr}$ is reflected in the 
2067: above equation being defined only up to a quantum of polarization.
2068: 
2069: \subsection{CMR dynamical equations}
2070: 
2071: In the case where $\hat{H}^0$ (and hence the CMR basis) is constant in time,
2072: plugging (\ref{eq:linear_comb}) into the TDSE (\ref{eq:tdse_continuum}) yields
2073: the CMR form of the Schr\"odinger equation,\cite{adams57,callaway91}
2074: %
2075: \begin{equation}
2076: \label{eq:tdse_cmr}
2077: i\hbar\,\dot{c}_{\kk m}=
2078: \big(
2079:   E_{\kk m}+ie\ee\cdot{\cal D}_\kk
2080: \big) c_{\kk m},
2081: \end{equation}
2082: %
2083: where we have simplified $c_{\kk,nm}$ to $c_{\kk m}$ and defined
2084: %
2085: \begin{equation}
2086: {\cal D}_\kk c_{\kk m}=\partial_\kk c_{\kk m}-
2087: i\sum_{l=1}^\infty\,{\bf X}_{\kk l}c_{\kk l},
2088: \end{equation}
2089: %
2090: which is reminiscent of the covariant derivative, Eq.~(\ref{eq:cov_der})
2091: (but note the difference in the sign of the last term).
2092: It is customary to write Eq.~(\ref{eq:tdse_cmr}) as
2093: %
2094: \begin{equation}
2095: \label{eq:tdse_cmr_b}
2096: i\hbar\,\dot{c}_{\kk m}=
2097: \big(
2098:   E_{\kk m}^{(1)}+ie\ee\cdot\partial_\kk
2099: \big) c_{\kk m}+e\ee\,\cdot\,\sum_{l\not=m}^\infty\,c_{\kk l}
2100: {\bf X}_{\kk,ml},
2101: \end{equation}
2102: %
2103: where
2104: %
2105: \begin{equation}
2106: \label{eq:shifted_band}
2107: E_{\kk m}^{(1)}=E_{\kk m}+e\ee\cdot {\bf X}_{\kk,mm}
2108: \end{equation}
2109: %
2110: is a shifted energy eigenvalue. $E_{\kk m}^{(1)}$
2111: is identical to Eq.~(\ref{eq:energy_k_density}) except that
2112: $\ket{v_{\kk m}}$ has been replaced by the zero-field eigenstate 
2113: $\ket{u_{\kk m}}$.
2114: Upon averaging over $\kk$ the last term on the right-hand-side becomes
2115: the first-order shift in total energy,
2116: $-v\pp_0\cdot\ee$, where $\pp_0$ is the spontaneous Berry-phase polarization.
2117: 
2118: In general the above TDSE has no stationary solutions.
2119: Approximate solutions --~the Wannier-Stark states~-- result from 
2120: restricting the wavepacket dynamics to a single band
2121: (the semiclassical approximation). That is achieved by
2122: dropping the sum on the right-hand-side of Eq.~(\ref{eq:tdse_cmr_b}),
2123: which is responsible for interband tunneling.\cite{callaway91,kane59}
2124: 
2125: Finally, combining (\ref{eq:dm_cmr}) and (\ref{eq:tdse_cmr}) produces the
2126: dynamical equation for the CMR density matrix:
2127: %
2128: \begin{eqnarray}
2129: \label{eq:sbe}
2130: i\hbar\,\dot{n}_{\kk,nm}&=&
2131: (E_{\kk n}-E_{\kk m})n_{\kk,nm}+ie\ee\cdot\partial_\kk n_{\kk,nm} \nonumber \\
2132: &-&e\ee\,\cdot\,\sum_{l=1}^\infty\,
2133: \big(
2134:   n_{\kk,nl}{\bf X}_{\kk,lm}-{\bf X}_{\kk,nl}n_{\kk,lm}
2135: \big).\nonumber \\
2136: \end{eqnarray}
2137: %
2138: A closely related form has been used to study the nonlinear optical 
2139: susceptibilities of semiconductors.\cite{aversa95,lambrecht00}
2140: 
2141: % ---------------------------------------------------------------------------
2142: 
2143: \section{Covariant derivative and related operators}
2144: \label{app:cov}
2145: 
2146: In Sec.~\ref{sec:td_continuum} we introduced a modified TDSE that contains
2147: the multiband covariant derivative $\widetilde{\partial}_\kk$, 
2148: Eq.~(\ref{eq:cov_der}), that was instrumental for making contact with
2149: the discrete-$k$ dynamical equations of Sec.~\ref{sec:td_discrete}.
2150: Here we summarize the properties of the covariant derivative and other closely
2151: related operators.
2152: 
2153: The covariant derivative $\widetilde{\partial}_\kk\ket{v_{\kk n}}$ 
2154: of an occupied state transforms
2155: in the same way as that state under a gauge transformation, 
2156: Eq.~(\ref{eq:gauge_transf}):
2157: %
2158: \begin{equation}
2159: \label{eq:gauge_cov_der}
2160: \widetilde{\partial}_\kk\ket{v_{\kk n}}\rightarrow
2161: \sum_{m=1}^M\,U_{\kk,mn}\,\widetilde{\partial}_\kk\ket{v_{\kk m}}.
2162: \end{equation}
2163: %
2164: Moreover, it is orthogonal to the occupied subspace at $\kk$,
2165: %
2166: \begin{equation}
2167: \label{eq:orthogonal}
2168: \bra{v_{\kk m}}\widetilde{\partial}_\kk v_{\kk n}\rangle=0.
2169: \end{equation}
2170: %
2171: Recalling that parallel transport is characterized by
2172: $\bra{v_{\kk n}} \partial_\kk v_{\kk n}\rangle=0$, for
2173: $m=n$ this relation shows that $\widetilde{\partial}_\kk$
2174: acting in an arbitrary gauge gives the same result as $\partial_\kk$
2175: acting in the parallel-transport gauge that shares the same states
2176: at $\kk$. In the discretized form (\ref{eq:cov_discrete}) the property 
2177: (\ref{eq:gauge_cov_der}) is a consequence of Eq.~(\ref{eq:gauge_cov}), and the
2178: property (\ref{eq:orthogonal}) is a consequence of Eq.~(\ref{eq:duality}).
2179: Like $i\partial_\kk$, $i\widetilde{\partial}_\kk$ is hermitian.
2180: By this we mean that
2181: its matrix representation in an orthonormal basis 
2182: (e.g., the $\ket{v_{\kk n}}$, $n=1,\ldots,M$ complemented by a set of 
2183: unoccupied states $\ket{c_{\kk j}}$) is hermitian. This is closely related
2184: to the hermiticity of the matrix $A_{\kk}^\alpha$ defined in
2185: Eq.~(\ref{eq:connection}). Finally, note that
2186: %
2187: \begin{equation}
2188: \label{eq:related_ops}
2189: i\widetilde{\partial}_\kk\ket{v_{\kk n}}=
2190: i\hat{Q}_\kk\partial_\kk\ket{v_{\kk n}}
2191: =i\hat{Q}_\kk\partial_\kk\hat{P}_\kk\ket{v_{\kk n}},
2192: \end{equation}
2193: %
2194: i.e., the action of $i\widetilde{\partial}_\kk$ on an occupied state is 
2195: identical to that of $i\hat{Q}_\kk\partial_\kk$ and
2196: $i\hat{Q}_\kk\partial_\kk\hat{P}_\kk$. They differ in how they
2197: act on the unoccupied states. Unlike
2198: $i\widetilde{\partial}_\kk$, the other two
2199: are not hermitian: for instance, 
2200: $%(i\hat{Q}_\kk\partial_\kk)^{\dagger}\ket{v_{\kk n}}=
2201: (i\hat{Q}_\kk\partial_\kk\hat{P}_\kk)^{\dagger}\ket{v_{\kk n}}=0$.
2202: It follows from these considerations that 
2203: Eq.~(\ref{eq:tdse_continuum_cov}) can be recast as
2204: %
2205: \begin{equation}
2206: \label{eq:tdse_continuum_cov_mod}
2207: i\hbar\ket{\dot{v}_{\kk n}}=\big[\hat{H}^0_\kk+
2208: e\ee\cdot\big(i\hat{Q}_\kk\partial_\kk\hat{P}_\kk+
2209: \text{h.c.}
2210: \big)
2211: \big]\ket{{v}_{\kk n}}.
2212: \end{equation}
2213: %
2214: This is the form of the TDSE to which 
2215: Eqs.~(\ref{eq:td_discrete_b})-(\ref{eq:t_k}) reduce in the continuum-$k$
2216: limit, since 
2217: %
2218: \begin{equation}
2219: \hat{\rm w}_\kk\simeq ie\ee\cdot\hat{Q}_\kk\partial_\kk\hat{P}_\kk
2220: \end{equation}
2221: %
2222: (compare with Eq.~(\ref{eq:cov_dis})).
2223: 
2224: % ---------------------------------------------------------------------------
2225: 
2226: 
2227: \section{Gradient of the energy functional}
2228: 
2229: The purpose of this Appendix is to obtain expressions for the
2230: derivatives of the two terms in the energy functional
2231: of Eq.~(\ref{eq:energy}) with respect to the occupied Bloch-like
2232: states in the discrete-$k$ case.  The results have been used
2233: in Secs.~\ref{sec:td_discrete} and \ref{sec:stat_formulation} for
2234: the discussion of the time-dependent evolution equations and the
2235: stationary solutions, respectively.
2236: 
2237: \subsection{Band-structure contribution}
2238: \label{app:grad-a}
2239: 
2240: To find the gradient $\delta E/\bra{\delta v_{\kk n}}$ of the energy
2241: functional (\ref{eq:energy}), let us isolate the terms that depend on 
2242: $\bra{v_{\kk n}}$. Using (\ref{eq:projector}) the zero-field part 
2243: (\ref{eq:energy0_discrete}) can be expressed as
2244: $E^0=(1/N)\,\sum_\kk\,{\rm tr}[\hat{P}_\kk\hat{H}^0_\kk]$, so that
2245: %
2246: \begin{equation}
2247: \frac{\delta E^0}{\bra{\delta v_{\kk n}}}=\frac{1}{N}
2248: \frac{\delta {\rm tr}[\hat{P}_\kk\hat{H}^0_\kk]}{\bra{\delta v_{\kk n}}}.
2249: \end{equation}
2250: %
2251: In order to allow for arbitrary variations of $\bra{v_{\kk n}}$,
2252: even those for which the $\bra{v_{\kk n}}$ do not remain orthonormal,
2253: we write
2254: %
2255: \begin{equation}
2256: \label{eq:proj_nonorthonormal}
2257: \hat{P_\kk}=\sum_{m,n=1}^M {({\cal S}_\kk^{-1})}_{mn} 
2258: \ket{v_{\kk m}} \bra{v_{\kk n}},
2259: \end{equation}
2260: %
2261: where ${\cal S}_{\kk,mn}=\langle v_{\kk m}|v_{\kk n}\rangle$. Dropping the 
2262: subscript $\kk$,
2263: %
2264: \begin{eqnarray}
2265: &{\delta{\rm tr}[\hat{P}\hat{H}^0]}&=
2266: {{\rm tr}[(\delta\hat{P})\hat{H}^0]} \nonumber \\
2267: &&=\sum_{m,n}\,({\cal S}^{-1})_{mn}
2268: \Big[\langle v_n|\hat{H}^0|\delta v_m\rangle+
2269:     \langle \delta v_n|\hat{H}^0|v_m\rangle
2270: \Big]
2271: \nonumber \\
2272: &&\qquad+\sum_{m,n}\,\langle v_n|\hat{H}^0|v_m\rangle\,
2273: \delta({\cal S}^{-1})_{mn}.
2274: \end{eqnarray}
2275: %
2276: Using $\delta({\cal S}^{-1})=-{\cal S}^{-2}\,\delta{\cal S}$ and
2277: $
2278: \delta {\cal S}_{mn}=
2279: \langle v_m|\delta v_n\rangle+\langle \delta v_m|v_n\rangle,
2280: $
2281: and evaluating at ${\cal S}={\bf 1}$, we arrive at
2282: %
2283: \begin{equation}
2284: \label{eq:grad_zerofld}
2285: \frac{\delta {\rm tr}[\hat{P}_\kk\hat{H}^0_\kk]}{\bra{\delta v_{\kk n}}}=
2286: \hat{Q}_\kk\,\hat{H}_\kk^0\,\ket{v_{\kk n}}.
2287: \end{equation}
2288: %
2289: Thus the consequence of expressing $\hat{P}_\kk$ as 
2290: (\ref{eq:proj_nonorthonormal}) 
2291: instead of (\ref{eq:projector}) is to 
2292: render the gradient orthogonal to the occupied manifold at $\kk$.
2293: (When we derived the dynamical equation (\ref{eq:td_discrete_a})
2294: using (\ref{eq:euler-lagrange_b}), the
2295: gradient of $E^0$ was not orthogonalized, which is why the
2296: dynamics did not follow parallel transport (see Sec.~\ref{sec:integration})).
2297: 
2298: \subsection{Polarization contribution}
2299: \label{app:grad-b}
2300: 
2301: To find the gradient of the field-coupling term $-v\ee\cdot\pp$
2302: we need $\delta\overline{\Gamma}_i/\bra{\delta v_{\kk n}}$. Let us start by
2303: recasting Eq.~(\ref{eq:discrete_phase}) as
2304: %
2305: \begin{equation}
2306: \overline{\Gamma}_i=
2307: \frac{1}{N^\perp_i}\,\sum_{l=1}^{N^\perp_i}\,\sum_{j=0}^{N_i^\parallel-1}\,
2308: \phi(\kk_j^{(i)},\kk_j^{(i)}+\Delta \kk_i),
2309: \end{equation}
2310: %
2311: where we have defined the phase
2312: %
2313: \begin{equation}
2314: \label{eq:phase}
2315: \phi(\kk,\kk')=-{\rm Im}\,{\rm ln}\,{\rm det}\,
2316: S(\kk,\kk').
2317: \end{equation}
2318: %
2319: Using $\phi(\kk',\kk)=-\phi(\kk,\kk')$, this becomes
2320: %
2321: \begin{equation}
2322: \label{eq:gamma_avg}
2323: \overline{\Gamma}_i=\frac{1}{N^\perp_i}\,\sum_{\sigma=\pm 1}\,\sigma\,
2324: \phi(\kk,\kk i\sigma)+\ldots,
2325: \end{equation}
2326: %
2327: where only the terms depending on $\bra{v_{\kk n}}$ were written explicitly.
2328: Hence
2329: %
2330: \begin{equation}
2331: \frac{\delta \overline{\Gamma}_i}{\bra{\delta v_{\kk n}}}=
2332: \frac{1}{N^\perp_i}\,\sum_{\sigma=\pm 1}\,\sigma\,
2333: \frac{\delta}{\bra{\delta v_{\kk n}}}\phi(\kk,\kk i\sigma).
2334: \end{equation}
2335: %
2336: The phase $\phi(\kk,\kk')$ can be expressed as
2337: %
2338: \begin{eqnarray}
2339: \label{eq:phase_b}
2340: \phi(\kk,\kk')&=&-{\rm Im}\,{\rm tr}\,{\rm ln}\,
2341: S(\kk,\kk')\nonumber\\
2342: &=&\frac{i}{2}\,{\rm tr}\,{\rm ln}\,S(\kk,\kk')\,-\,
2343: \frac{i}{2}\,{\rm tr}\,{\rm ln}\,S(\kk',\kk).
2344: \end{eqnarray}
2345: %
2346: For an arbitrary non-singular matrix $A$ we have
2347: %
2348: \begin{eqnarray}
2349: \delta {\rm tr}\,{\rm ln}A&=&{\rm tr}\,{\rm ln}(A+\delta A)-{\rm tr}\,
2350: {\rm ln}(A)\nonumber \\
2351: &=&{\rm tr}\,{\rm ln}[(A+\delta A)A^{-1}]=
2352: {\rm tr}\,{\rm ln}[{\bf 1}+(\delta A)A^{-1}]\nonumber\\
2353: &=&{\rm tr}[A^{-1}\,\delta A]+{\cal O}(\delta A^{2}),
2354: \end{eqnarray}
2355: %
2356: so that
2357: %
2358: \begin{eqnarray}
2359: \frac{\delta {\rm tr}\,{\rm ln}S(\kk,\kk')}{\bra{\delta v_{\kk n}}}&=&
2360: {\rm tr}
2361: \left[
2362:   S^{-1}(\kk,\kk')\,\frac{\delta S(\kk,\kk')}{\bra{\delta v_{\kk n}}}
2363: \right]\nonumber \\
2364: &=&\sum_{m=1}^M\,S^{-1}_{mn}(\kk,\kk')\,\ket{v_{\kk' m}}\nonumber\\
2365: &=&\ket{\widetilde{v}_{\kk'n}}.
2366: \end{eqnarray}
2367: %
2368: The corresponding derivative of the last term of (\ref{eq:phase_b})
2369: vanishes since $S(\kk',\kk)=\langle v_{\kk'n}\vert v_{\kk n}\rangle$
2370: does not contain $\bra{v_{\kk n}}$ as a bra.  We thus arrive at
2371: %
2372: \begin{equation}
2373: \label{eq:grad_phase}
2374: \frac{\delta \phi(\kk,\kk')}{\bra{\delta v_{\kk n}}}=\frac{i}{2}\,
2375: \ket{\widetilde{v}_{\kk'n}},
2376: \end{equation}
2377: %
2378: which combined with (\ref{eq:gamma_avg}) gives 
2379: %
2380: \begin{equation}
2381: \label{eq:grad_berry_phase}
2382: \frac{\delta \overline{\Gamma}_i}{\bra{\delta v_{\kk n}}}=
2383: \frac{i}{2N^\perp_i}\sum_{\sigma=\pm1}\,\sigma |\widetilde{v}_{\kk i\sigma,n}
2384: \rangle.
2385: \end{equation}
2386: %
2387: This is automatically orthogonal to the occupied manifold at $\kk$.
2388: (See Refs.~\onlinecite{nunes01,sai02} for alternative derivations.)
2389: Collecting terms and using Eq.~(\ref{eq:ket_w_k}),
2390: we obtain Eq.~(\ref{eq:gradient_energy}) for the gradient of the full
2391: energy functional $E$.
2392: 
2393: % ---------------------------------------------------------------------------
2394: 
2395: \section{Discretized formula for the current}
2396: \label{app:discr}
2397: 
2398: Just as the macroscopic polarization $\pp$
2399: is evaluated in practice via a finite-difference formula on a mesh of $k$ 
2400: points, the same can be done for the macroscopic current $\jj=d\pp/dt$.
2401: The invariance of Eq.~(\ref{eq:curr}) under the
2402: replacement $\partial_\kk\rightarrow\widetilde{\partial}_\kk$ allows us to then
2403: use the discretization rule (\ref{eq:cov_discrete}), 
2404: leading to
2405: %
2406: \begin{equation}
2407: \label{eq:curr_discrete}
2408: \jj=\frac{e}{4\pi\hbar v}\,\sum_{n,\kk}\,\sum_{i=1}^3\,\sum_{\sigma=\pm 1}\,
2409: \frac{\sigma}{N_i^\perp}\,
2410: \langle v_{\kk n}|\hat{H}_\kk^0|\widetilde{v}_{\kk i\sigma,n}\rangle\aa_i
2411: +{\rm c.c.},
2412: \end{equation}
2413: %
2414: which is significantly easier and cheaper to compute than the spatial average 
2415: of the microscopic current.
2416: We have checked numerically on our one-dimensional tight-binding model
2417: that Eq.~(\ref{eq:curr_discrete}) yields, for small $\Delta t$, the same result
2418: as $[P(t+\Delta t)-P(t)]/\Delta t$ computed with the
2419: discretized Berry-phase formula. 
2420: 
2421: The same strategy as outlined above can be used to derive a discretized
2422: formula for the Berry curvature (\ref{eq:curvature}) summed over
2423: bands, which is also invariant under 
2424: $\partial_\kk\rightarrow\widetilde{\partial}_\kk$. 
2425: This may be useful in other contexts, such as semiclassical wavepacket dynamics
2426: in crystals.\cite{chang96}
2427: 
2428: 
2429: 
2430: % ---------------------------------------------------------------------------
2431: 
2432: \begin{thebibliography}{0}
2433: 
2434: 
2435: \bibitem{ksv93} R. D. King-Smith and D. Vanderbilt, Phys. Rev. B {\bf 47},
2436: 1651 (1993).
2437: 
2438: \bibitem{berry84} M. V. Berry, Proc. R. Soc. London Ser. A {\bf 392}, 45
2439: (1984).
2440: 
2441: \bibitem{mv97} N. Marzari and D. Vanderbilt, Phys. Rev. B  {\bf 56},
2442: 12847 (1997).
2443: 
2444: \bibitem{nunes94} R. W. Nunes and D. Vanderbilt, Phys. Rev. Lett. {\bf 73},
2445: 712 (1994).
2446: 
2447: \bibitem{dalcorso94} A. Dal Corso and F. Mauri, Phys. Rev. B {\bf 50},
2448: 5756 (1994)
2449: 
2450: \bibitem{dalcorso96} A. Dal Corso, F. Mauri, and A. Rubio, Phys. Rev. B
2451: {\bf 53}, 15638 (1996).
2452: 
2453: \bibitem{nunes01} R. W. Nunes and X. Gonze, Phys. Rev. B {\bf 63}, 155107
2454: (2001).
2455: 
2456: \bibitem{souza02} I. Souza, J. \'I\~niguez, and D. Vanderbilt, Phys. Rev.
2457: Lett. {\bf 89}, 117602 (2002).
2458: 
2459: \bibitem{umari02} P. Umari and A. Pasquarello, Phys. Rev.
2460: Lett. {\bf 89}, 157602 (2002). 
2461: 
2462: \bibitem{sai02} N. Sai, K. M. Rabe, and D. Vanderbilt, Phys. Rev. B {\bf 66},
2463: 104108 (2002).
2464: 
2465: \bibitem{mazur02} S. K. Sundaram and E. Mazur, Nature Materials {\bf 1},
2466: 217 (2002).
2467: 
2468: \bibitem{lazzeri03} See also M. Lazzeri and F. Mauri, Phys. Rev. Lett.
2469: {\bf 90}, 036401 (2003).
2470: 
2471: \bibitem{krieger86} J. B. Krieger and G. J. Iafrate, Phys. Rev. B {\bf 33},
2472: 5494 (1986); G. J. Iafrate, J. P. Reynolds, J. He, and J. B. Krieger,
2473: Int. Journ. of High Speed Electronics and Systems {\bf 9}, 223 (1998).
2474: 
2475: \bibitem{blount62} E. I. Blount, in {\em Solid State Physics, Advances in
2476: Research and Applications}, edited by F. Seitz and D. Turnbull (Academic,
2477: New York, 1962), Vol. 13, p.~305, and references cited therein.
2478: 
2479: \bibitem{foot:mistake} An incorrect interpretation was given
2480: in Ref.~\onlinecite{nunes01}
2481: to the term $ie{\ee}\cdot\partial_\kk\ket{{v}_{\kk n}}$. It does not come
2482: from the ``periodic part'' of the position operator, but rather from
2483: the full position operator, as can be seen from the present derivation. 
2484: 
2485: \bibitem{foot:wavepacket} After removing the ``band'' indices
2486: Eq.~(\ref{eq:orthonormality}) becomes the well-known equation
2487: for the $k$-space dynamics of an electronic wavepacket 
2488: $|\psi\rangle=\int |\phi_{\kk}\rangle\,d\kk$:
2489: $\dot{n}_{\kk}=(e/\hbar)\ee\cdot\partial_{\kk}n_{\kk}$, where
2490: $n_{\kk}=\langle v_{\kk}|v_{\kk} \rangle$. Eq.~(\ref{eq:tdse_continuum})
2491: is also valid for such wavepackets, even if the derivation given in the text
2492: was for WR manifolds. 
2493: 
2494: \bibitem{foot:frustrated_metal}  Such a coherent ``frustrated metal'' state
2495: can exist in the early moments following a strong femtosecond laser-pulse
2496: excitation of a semiconductor, but it will eventually be destroyed by thermal 
2497: scattering.\cite{mazur02}
2498: 
2499: \bibitem{aa87} Y. Aharonov and J. Anandan, Phys. Rev. Lett. {\bf 58}, 1593
2500: (1987).
2501: 
2502: \bibitem{explan-notanom} 
2503: The resemblance of Eqs.~(\ref{eq:vanishing_part})-(\ref{eq:curvature}) to the
2504: so-called anomalous current\cite{chang96} in semiclassical wavepacket
2505: dynamics is misleading, since the Berry
2506: curvature entering the anomalous current pertains to the band states
2507: $\ket{u_{\kk n}}$, not to the $\ket{v_{\kk n}}$. 
2508: In fact, it can be shown that
2509: the anomalous current is contained in the other term,
2510: Eq.~(\ref{eq:curr}), once it is properly modified for wavepackets.
2511: 
2512: \bibitem{kohn64} W. Kohn, Phys. Rev. {\bf 133}, A171 (1964);
2513: W. Kohn, in {\em Many-Body Physics}, edited by C. DeWitt and R. Balian
2514: (Gordon and Breach, New York, 1968), p. 351.
2515: 
2516: \bibitem{souza00} I. Souza, T. Wilkens, and R. M. Martin, Phys. Rev. B
2517: {\bf 62}, 1666 (2000).
2518: 
2519: \bibitem{resta99} R. Resta and S. Sorella, Phys. Rev. Lett. {\bf 82},
2520: 370 (1999).
2521: 
2522: \bibitem{goldstein80} H. Goldstein, {\em Classical Mechanics}, 2nd ed.
2523: (Addison-Wesley, Reading, Massachusetts, 1980).
2524: 
2525: \bibitem{fradkin} E. Fradkin, {\em Field Theories of Condensed Matter
2526: Systems} (Addison-Wesley, Reading, Massachusetts, 1991).
2527: 
2528: \bibitem{foot:wavepacket_b} Note that for a purely electronic wavepacket 
2529: $\psi(\rr)=\int e^{i\kk\cdot\rr}v_\kk(\rr)\,d\kk$
2530: the property $\partial_\kk\bra{v_\kk}v_\kk\rangle=0$ does not hold, and hence
2531: $(A_{\kk}^\alpha)^* \not=A_{\kk}^\alpha$. As a consequence, 
2532: unlike Eq.~(\ref{eq:tdse_continuum}), which correctly
2533: describes the dynamics of wavepackets,\cite{foot:wavepacket}
2534: Eq.~(\ref{eq:tdse_continuum_cov}) is only valid for WR manifolds.
2535: Of course, when wavepackets of {\it electron-hole pairs} are created
2536: by vertical electronic transitions in $k$ space, the electronic manifold
2537: remains Wannier-representable, and can be described within the present
2538: formalism (see, e.g., Sec.~\ref{sec:cdw}).
2539: 
2540: \bibitem{koonin90} S. E. Koonin and D. C. Meredith, {\em
2541: Computational Physics} (Addison-Wesley, Reading, Massachusetts, 1990).
2542: 
2543: \bibitem{sugino99} O. Sugino and Y. Miyamoto, Phys. Rev. B {\bf 59}, 2579
2544: (1999).
2545: 
2546: \bibitem{watanabe02}  N. Watanabe and M. Tsukada, Phys. Rev. E {\bf 65},
2547: 036705 (2002).
2548: 
2549: \bibitem{nenciu91} G. Nenciu, Rev. Mod. Phys. {\bf 63}, 91 (1991).
2550: 
2551: \bibitem{unpublished} J. \'I\~niguez, I. Souza, and D. Vanderbilt 
2552: (unpublished).
2553: 
2554: \bibitem{wannier55} G. Wannier, Phys. Rev. {\bf 100}, 1227 (1955);
2555: {\bf 101}, 1835 (1956).
2556: 
2557: \bibitem{adams57} E. N. Adams, Phys. Rev. {\bf 107}, 698 (1957).
2558: 
2559: \bibitem{soler02} J. M. Soler {\it et al.}, J. Phys.: Condens. Matter
2560: {\bf 14}, 2745 (2002).
2561: 
2562: \bibitem{bennetto96} J. Bennetto and D. Vanderbilt, Phys. Rev. B {\bf 53}, 
2563: 15417 (1996).
2564: 
2565: \bibitem{dunlap86} D. H. Dunlap and V. M. Kenkre, Phys. Rev. B {\bf 34}, 3625
2566: (1986).
2567: 
2568: \bibitem{foreman02} B. A. Foreman, Phys. Rev. B {\bf 66}, 165212 (2002), and
2569: references cited therein.
2570: 
2571: \bibitem{thouless83} D. J. Thouless, Phys. Rev. B {\bf 27}, 6083 (1983).
2572: 
2573: \bibitem{foot:thouless} It may seem surprising that even though we use a 
2574: discrete mesh of
2575: $k$-points --~ which effectively turns the system into a finite torus~-- 
2576: our calculations display exact quantization of adiabatic particle transport. 
2577: According to Thouless,\cite{thouless83} that  should only
2578: occur in the thermodynamic (continuum-$k$) limit. This has to do with our
2579: using the discretized Berry-phase formula (\ref{eq:discrete_phase}),
2580: which is not fully consistent with the velocity operator (\ref{eq:velocity})
2581: on the finite torus (we recall that the derivation in 
2582: Sec.~\ref{sec:dyn_pol_der} was done in the
2583: continuum-$k$ limit). From this viewpoint it seems preferable to regard the 
2584: discretization procedure as a numerical approximation
2585: to the continuum-$k$ limit, 
2586: rather than as an exact calculation on the finite torus.
2587: 
2588: \bibitem{foot:insulating-like} In the limit where the number $N$ of $k$ points 
2589: is large,
2590: $\xi$ is finite for the ground state of insulators and diverges for that of
2591: metals.\cite{souza00} 
2592: For finite $N$ it cannot really diverge. In particular,
2593: when $\xi=(\Omega_{\rm I}/M)^{1/2}$ is evaluated using Eq.~(34)
2594: of Marzari and Vanderbilt\cite{mv97} it cannot exceed, in one dimension, 
2595: $aN/2\pi$, well above the values that we observe
2596: in all our simulations.
2597: 
2598: \bibitem{odwyer73} J. J. O'Dwyer, {\em The Theory of Electrical Conduction
2599: and Breakdown in Solid Dielectrics} (Clarendon, Oxford, 1973), p.~13.
2600: 
2601: \bibitem{kane59} E. O. Kane, J. Phys. Chem. Solids {\bf 12}, 181 (1959).
2602: 
2603: \bibitem{yu96} P. Y. Yu and M. Cardona, {\em Fundamentals of Semiconductors}
2604: (Springer, Berlin, 1996).
2605: 
2606: \bibitem{rink89} S. Schmitt-Rink, D. S. Chemla, and D. A. B. Miller,
2607: Adv. Phys. {\bf 38}, 89 (1989).
2608: 
2609: \bibitem{ando97} M. Ando {\it et al.}, Superlatt. Microstruct. {\bf 22},
2610: 459 (1998).
2611: 
2612: \bibitem{tsolakidis02} A. Tsolakidis, D. S\'anchez-Portal, and R. M. Martin,
2613: Phys. Rev. B {\bf 66}, 235416 (2002).
2614: 
2615: \bibitem{yabana99} K. Yabana and G. F. Bertsch, Int. J. Quant. Chem. {\bf 75},
2616: 55 (1999).
2617: 
2618: \bibitem{bertsch00} G.F. Bertsch, J.-I. Iwata, A. Rubio, and K. Yabana,
2619: Phys. Rev. B {\bf 62}, 7998 (2000).
2620: 
2621: \bibitem{jones73} W. Jones and N. H. March, {\em Theoretical Solid State
2622: Physics}, Vol. 1 (Wiley, New York, 1973).
2623: 
2624: \bibitem{foot:normalization} Since $\delta(\kk-\kk')$ has units of real-space
2625: volume, the right-hand side of Eq.~(\ref{eq:normalization}) is dimensionless.
2626: In the discrete-$k$ case the integral on the left-hand side is over the
2627: supercell volume $V=Nv$, and $\Omega_B\delta(\kk-\kk')$ is replaced 
2628: by $N\delta_{\kk \kk'}$.
2629: 
2630: \bibitem{sipe00} J. E. Sipe and A. I. Shkrebtii, Phys. Rev. B {\bf 61}, 5337
2631: (2000).
2632: 
2633: \bibitem{lambrecht00} W. R. L. Lambrecht and S. N. Rashkeev, Phys. Stat.
2634: Sol. (b) {\bf 217}, 599 (2000).
2635: 
2636: \bibitem{callaway91} J. Callaway, {\em Quantum Theory of the Solid State},
2637: 2nd ed. (Academic Press, New York, 1991).
2638: 
2639: \bibitem{aversa95} C. Aversa and J. E. Sipe, Phys. Rev. B {\bf 52}, 14636
2640: (1995).
2641: 
2642: \bibitem{chang96} M. C. Chang and Q. Niu, Phys. Rev. B {\bf 53}, 7010 (1996).
2643: 
2644: 
2645: \end{thebibliography}
2646: 
2647: \end{document}
2648: