1: \documentclass[twocolumn,showpacs,floatfix,aps]{revtex4}
2: \usepackage{graphicx}% Include figure files
3: \usepackage{dcolumn}% Align table columns on decimal point
4: \usepackage{bm}% bold math
5: \newcommand{\n}{\mbox{\boldmath $\nabla$}}
6: \renewcommand\baselinestretch{1.0}
7: \begin{document}
8:
9: %\draft
10:
11: \title{Two-particle localization and antiresonance \\ in disordered spin
12: and qubit chains}
13:
14: \author{L. F. Santos}
15: \email{santos@pa.msu.edu}
16: \author{M. I. Dykman}
17: \email{dykman@pa.msu.edu}
18: \affiliation{Department of Physics and Astronomy
19: and the Institute for Quantum Sciences,
20: Michigan State University, East Lansing, Michigan 48824}
21:
22: \date{\today}
23:
24: \begin{abstract}
25: We show that, in a system with defects, two-particle states may
26: experience destructive quantum interference, or antiresonance. It
27: prevents an excitation localized on a defect from decaying even
28: where the decay is allowed by energy conservation. The system studied
29: is a qubit chain or an equivalent spin chain with an anisotropic ($XXZ$)
30: exchange coupling in a magnetic field. The chain has a defect with an
31: excess on-site energy. It corresponds to a qubit with the
32: level spacing different from other qubits. We show that, because of
33: the interaction between excitations, a single defect may lead to
34: multiple localized states. The energy spectra and localization lengths
35: are found for two-excitation states. The localization of
36: excitations facilitates the operation of a quantum
37: computer. Analytical results for strongly anisotropic coupling are
38: confirmed by numerical studies.
39: \end{abstract}
40:
41: \pacs{03.67.Lx,75.10.Pq,75.10.Jm,73.21.-b}
42:
43: \maketitle
44: %\newpage
45:
46: \section{Introduction}
47:
48: One of the most important potential applications of quantum computers
49: (QC's) is studies of quantum many-body effects. It is particularly
50: interesting to find new many-body effects in condensed-matter systems
51: that could be easily simulated on a QC. In the present paper we
52: discuss one such effect: antiresonance, or destructive quantum
53: interference between two-particle excitations in a system with
54: defects. We also study interaction-induced two-particle localization on
55: a defect and discuss implications of the results for quantum
56: computing.
57:
58: The basic elements of a QC, qubits, are two-state systems. They are
59: naturally modeled by spin-1/2 particles. In many suggested
60: realizations of QC's, the qubit-qubit interaction is ``on'' all the time
61: \cite{recent_reviews}. In terms of spins, it corresponds to exchange
62: interaction. The dynamics of such QC's and spin systems in solids have
63: many important similar aspects that can be studied together.
64:
65: In most proposed QC's the energy difference between the qubit states is
66: large compared to the qubit-qubit interaction. This corresponds to a
67: system of spins in a strong external magnetic field. However, in
68: contrast to ideal spin systems, level spacings of different qubits
69: can be different. A major advantageous feature of QC's is that the qubit
70: energies can be often individually controlled
71: \cite{Makhlin99,JJ-all,mark}. This corresponds to controllable
72: disorder of a spin system, and it allows one to use QC's for studying a
73: fundamentally important problem of how the spin-spin interaction
74: affects spin dynamics in the presence of disorder.
75:
76: %The analogy with coupled-spin systems is particularly relevant for the
77: %QC's where the interqubit interaction is permanently on rather than
78: %being turned on and off.
79:
80: Several models of QC's where the interqubit interaction is permanently
81: ``on'' are currently studied. In these models the effective spin-spin
82: interaction is usually strongly anisotropic. It varies from the
83: essentially Ising coupling $\sigma_n^z\sigma_m^z$ in nuclear
84: magnetic resonance and some
85: other systems
86: \cite{Chuang9802,Yamamoto02,Cory00,Piermarocchi02,Demille02} ($n,m$
87: enumerate qubits, and $z$ is the direction of the magnetic field) to
88: the $XY$-type (i.e., $\sigma_n^x\sigma_m^x + \sigma_n^y\sigma_m^y$) or
89: the $XXZ$-type (i.e., $\Delta_{nm}\sigma_n^z\sigma_m^z+
90: \sigma_n^x\sigma_m^x + \sigma_n^y\sigma_m^y$) coupling in some
91: Josephson-junction based systems \cite{Makhlin99,JJ-all}.
92:
93: The Ising coupling describes the system in the case where the
94: transition frequencies of different qubits are strongly
95: different. Then $\sigma_n^z\sigma_m^z$ is the only part of the
96: interaction that slowly oscillates in time, in the Heisenberg
97: representation, and therefore is not averaged out. If the qubit
98: frequencies are close to each other, the terms $\sigma_n^x\sigma_m^x +
99: \sigma_n^y\sigma_m^y$ become smooth functions of time as well. They
100: lead to resonant excitation hopping between qubits. In a multiqubit
101: system with close frequencies, both Ising and $XY$ interactions are
102: present in the general case, but their strengths may be different
103: \cite{Silvestrov01,Kaminsky-Lloyd02}. In this sense the $XXZ$
104: coupling is most general, at least for qubits with high transition
105: frequencies.
106:
107: The interqubit interaction often rapidly falls off with the distance
108: and can be approximated by nearest neighbor coupling. Many important
109: results on anisotropic spin systems with such coupling have been
110: obtained using the Bethe ansatz. Initially the emphasis was placed on
111: systems without defects \cite{Bethe} or with defects on the edge of a
112: spin chain \cite{Baxter}. More recently these studies have been
113: extended to systems with defects that are described by integrable
114: Hamiltonians \cite{integrable}. However, the problem of a spin chain
115: with several coupled excitations and with defects of a general type
116: has not been solved.
117:
118: In this paper we investigate interacting
119: excitations in an anisotropic spin system
120: with defects. We show that the excitation localized on a defect does
121: not decay even where the decay is allowed by energy conservation. We
122: also find that, in addition to a single-particle excitation, a defect
123: leads to the onset of two types of localized two-particle excitations.
124:
125: The analysis is done for a system with the $XXZ$ coupling. The coupling
126: anisotropy is assumed to be strong, as in the case of a QC based on
127: electrons on helium, for example \cite{mark}. The ground state of the
128: system corresponds to all spins pointing in the same direction
129: (downwards, for concreteness). A single-particle excitation
130: corresponds to one qubit being excited, or one spin being flipped. If
131: the qubit energies are tuned in resonance with each other, a QC
132: behaves as an ideal spin system with no disorder. A single-particle
133: excitation is then magnon-type, it freely propagates through the
134: system.
135:
136: In the opposite case where the qubit energies are tuned far away from
137: each other (as for diagonal disorder in tight-binding models), all
138: single-particle excitations are localized. If the excitation density
139: is high, the interaction between them may affect their localization,
140: leading to quantum chaos, cf. Refs.
141: \onlinecite{Silvestrov01,Shepelyansky00,Izrailev01,Lloyd02}.
142: Understanding the interplay between interaction and disorder is a
143: prerequisite for building a QC. We will consider the case where the
144: excitation density is low, yet the interaction is important. In
145: particular, excitations may form bound pairs (but the pair density is
146: small).
147:
148: One of the important questions is whether the interaction leads to
149: delocalization of excitations. More specifically, consider an
150: excitation, which is localized on a defect in the absence of other
151: excitations. We now create an extended magnon-type excitation (a
152: propagating wave), that can be scattered off the localized one.
153: % \footnote{In the case of a
154: %finite-length spin chain, we use the term ``extended'' for the states
155: %with a sinusoidal wave function.}.
156: The problem is whether this will
157: cause the excitation to move away from the defect. We show below that,
158: due to unexpected destructive quantum interference, the scattering
159: does not lead to delocalization.
160:
161:
162: \subsection{Model and preview}
163: We consider a one-dimensional array of qubits which models a spin-1/2
164: chain. For nearest-neighbor coupling, the Hamiltonian is
165: %
166: \begin{eqnarray}
167: \label{hamiltonian}
168: &&H = \frac{1}{2}\sum_n \varepsilon^{(n)} \sigma_{n}^{z} +
169: \frac{1}{4}\sum_n \sum_{i=x,y,z}J_{ii}\sigma_n^i\sigma_{n+1}^i\\
170: &&J_{xx}=J_{yy}=J,\;J_{zz}=J\Delta\nonumber .
171: \end{eqnarray}
172: %
173: Here, $\sigma_n^i$ are the Pauli matrices and $\hbar =1$. The
174: parameter $J$ characterizes the strength of the exchange coupling, and
175: $\Delta$ determines the coupling anisotropy. We assume that
176: $|\Delta|\gg 1$; for a QC based on electrons on helium, $|\Delta|$
177: lies between $20$ and $8$, for typical parameter values
178: \cite{mark}.
179:
180: We will consider effects due to a single defect. Respectively, all
181: on-site spin-flip energies $\varepsilon^{(n)}$ are assumed to be the
182: same except for the site $n=n_0$ where the defect is located, that is,
183: %
184: \begin{equation}
185: \label{chain_energies}
186: \varepsilon^{(n)}= \varepsilon + g\delta_{n,n_0} .
187: \end{equation}
188:
189: In order to formulate the problem of interaction-induced decay of
190: localized excitations, we preview in Fig.~\ref{fig:energies} a part of
191: the results on the energy spectrum of the system. In the absence of
192: the defect, the energies of single-spin excitations (magnons) lie
193: within the band ${\varepsilon}_1 \pm J$, where ${\varepsilon}_1
194: =\varepsilon-J\Delta $ (the energy is counted off from the
195: ground-state energy). The defect has a spin-flip energy that differs by $g$
196: (a qubit with a transition frequency different from that of other
197: qubits). It leads to a localized single-spin excitation with no
198: threshold in $g$, for an infinite chain. The energy of the localized
199: state is shown by a dashed line on the left panel of
200: Fig.~\ref{fig:energies}.
201:
202: \begin{figure}[ht]
203: \includegraphics[width=3.2in]{fig1_XXZ.eps}
204: \caption{Left panel: the one-excitation energy spectrum in an infinite
205: spin chain with a defect. The energies of extended states (magnons)
206: form a band of width $J$ centered at ${\varepsilon}_1$. The dashed
207: line indicates the energy of the excitation localized on the
208: defect. Right panel: the two-excitation energy spectrum. The band
209: $2({\varepsilon}_1\pm J)$ is formed by uncoupled magnons. The band
210: centered at $2{\varepsilon}_1+g$ is formed by localized-delocalized
211: pairs (LDP's) in which one excitation is localized on the defect and
212: the other is in an extended state. The narrow band centered at
213: $2{\varepsilon}_1 + J\Delta$ is formed by propagating bound pairs (BP's)
214: of excitations. The dashed lines show the energies of the
215: states where both excitations are bound to the defect.}
216: \label{fig:energies}
217: \end{figure}
218:
219: We now discuss excitations that correspond to two flipped spins. A
220: defect-free $XXZ$ system has a two-magnon band of independently
221: propagating noninteracting magnons. However, the anisotropy of the
222: exchange coupling leads also to the onset of bound pairs (BP's) of
223: excitations. The BP band is much narrower than the two-magnon band and
224: is separated from it by a comparatively large energy difference
225: $J\Delta$, see the right panel of Fig.~\ref{fig:energies}. In the
226: presence of a defect, there are two-excitation states with one
227: excitation localized on the defect and the other being in an extended
228: state. We call them localized-delocalized pairs (LDP's). An interplay
229: between disorder and interaction may lead to new types of states where
230: both excitations are localized near the defect. Their energies are
231: shown in the right panel of Fig.~\ref{fig:energies} by dashed lines.
232:
233: A localized one-spin excitation cannot decay by emitting a magnon, by
234: energy conservation. But it might experience an
235: induced decay when a magnon is inelastically scattered off the excited
236: defect into an extended many-spin state. Magnon-induced decay is
237: allowed by energy conservation when the total energy of the localized
238: one-spin excitation and the magnon coincides with the energy of
239: another two-particle state. In the $XXZ$ model the total number of
240: excitations (flipped spins) is conserved, and therefore decay is only
241: possible into extended states of two bound magnons. In other words, it
242: may only happen when the LDP band overlaps with the BP band in
243: Fig.~\ref{fig:energies}.
244:
245: Decay into BP states may occur directly or via the two-excitation
246: state located next to the defect. The amplitudes of the corresponding
247: transitions turn out to be nearly equal and opposite in sign. As a
248: result of this quantum interference, even though the band of bound
249: magnons is narrow and has high density of states, the LDP to BP
250: scattering does not happen, i.e., the excitation on the defect is not
251: delocalized. The BP to LDP scattering does not happen either, i.e., a
252: localized excitation is not created as a result of BP decay.
253:
254: In Sec.~II below and in the Appendix we briefly analyze localization of
255: one excitation in a finite chain with a defect, for different boundary
256: conditions. In Sec.~III we discuss the two-excitation states localized
257: near a defect. In Sec.~IV we consider the resonant situation where the
258: energy band of extended bound pair states is within the band of
259: energies of the flipped defect spin plus a magnon, i.e., where the BP
260: band overlaps with the LDP band in Fig.~\ref{fig:energies}. We find
261: that the localized excitation remains on the defect site in this
262: case. Analytical results for a chain with strong anisotropy $|\Delta|$
263: are compared with numerical calculations. Section V contains concluding
264: remarks.
265:
266:
267:
268: \section{One excitation: localized and extended states}
269:
270: In order to set the scene for the analysis of the two-excitation case,
271: in this section and in the Appendix we briefly discuss the well-known case
272: of one excitation (flipped spin) in an $XXZ$ spin chain with a defect
273: \cite{Economou} and the role of boundary conditions. The Hamiltonian
274: of the chain with the defect on site $n_0$ has the form
275: (\ref{hamiltonian}), (\ref{chain_energies}). We assume that the
276: excitation energy $\varepsilon$ largely exceeds both the coupling
277: constant $|J|$ and the energy excess on the defect site $|g|$.
278: In this case the ground-state of the system corresponds to all spins
279: being parallel, with $\langle\sigma_n^z\rangle=-1$
280: irrespective of the signs of $J,g$, and $\Delta$.
281:
282:
283: Without a defect, one-spin excitations are magnons. They freely
284: propagate throughout the chain. The term in the Hamiltonian
285: (\ref{hamiltonian}) responsible for one-excitation hopping is
286: $H^{(t)}=\sum\nolimits_n H^{(t)}_n$, with
287: %
288: \begin{eqnarray}
289: \label{transfer_one}
290: &H^{(t)}_n = {1\over 4}J \sum_{i=x,y} \sigma_n^i\sigma_{n+1}^i \equiv {1\over
291: 8}J (t_n^{(l)} + t_n^{(r)}),\\
292: %
293: &t_n^{(r)}= [t_{n}^{(l)}]^{\dagger}=
294: \sigma_{n+1}^{+} \sigma_{n}^{-}.\nonumber
295: \end{eqnarray}
296: %
297: The operators $t_{n}^{(r)}$ and $t_{n}^{(l)}$ cause excitation
298: shifts $n\to n+1$ and $n+1\to n$, respectively.
299:
300: A defect leads to magnon scattering and to the onset of localized
301: states. Both propagation of excitations and their localization are
302: interesting for quantum computing. Coherent excitation transitions allow
303: one to have a QC geometry where remote ``working'' qubits are
304: connected by chains of ``auxiliary'' qubits, which form ``transmission
305: lines'' \cite{mark}. Localization, on the other hand, allows one to
306: perform single-qubit operations on targeted qubit.
307:
308: A QC makes it possible to model spin chains with different boundary
309: conditions. The simplest models are an open spin chain with free
310: boundaries, which is mimicked by a finite-length array of qubits (for
311: electrons on helium, it can be implemented using an array of equally
312: spaced electrodes, cf. Ref.~\onlinecite{Goodkind01}) or a periodic
313: chain, which can be mimicked by a ring of qubits.
314:
315: An open $N$-spin chain is described by the Hamiltonian
316: (\ref{hamiltonian}), where the first sum runs over $n=1,\ldots,N$ and
317: the second sum runs over $n=1,\ldots,N-1$ (the edge spins have
318: neighbors only inside the chain). In what follows, we count energy off
319: the ground-state energy ${E}_{0} = -(N\varepsilon+g)/2 + (N-1) J
320: \Delta /4 $, i.e., we replace in Eq.~(\ref{hamiltonian}) $H\to H-E_0$.
321:
322: The eigenfunctions of $H$ in the case of one
323: excitation can be written as \cite{Economou}
324:
325: \begin{equation}
326: \psi_1 = \sum_{n=1}^{N} a(n) \phi (n)
327: \label{one_exc} ,
328: \end{equation}
329: %
330: where $\phi (n) $ corresponds to the spin on site $n$ being up
331: and all other spins being down. The Schr\"odinger equation for
332: $a(n) $ has the form
333:
334: \begin{eqnarray}
335: && \left( {\varepsilon}_{1} + g \delta_{n,n_0}
336: +\frac{J\Delta }{2} \delta_{n,1}
337: +\frac{J\Delta }{2} \delta_{n,N} \right) a(n)
338: \nonumber \\
339: &&+
340: \frac{J }{2} [a(n-1) + a(n+1)]=E_1a(n),
341: \label{eq_a}
342: \end{eqnarray}
343: %
344: where ${\varepsilon}_{1} = \varepsilon -J\Delta$ is the energy of a
345: flipped spin in an ideal infinite chain in the absence of excitation
346: hopping and $E_1$ is the one-excitation energy eigenvalue. For an
347: open $N$-spin chain we set $a(0)= a(N+1)=0$ in
348: Eq.~(\ref{eq_a}).
349:
350: The Hamiltonian of a closed $N$-spin chain has the form
351: (\ref{hamiltonian}), where both sums over $n$ go from 1 to $N$ and the
352: site $N+1$ coincides with the site 1. Here, the defect
353: location $n_0$ can be chosen arbitrarily.
354: The wave function can be sought in the
355: form (\ref{one_exc}). The Schr\"odinger equation then has the form
356: (\ref{eq_a}), except that there are no terms proportional to $\delta_{n,1},\,
357: \delta_{n,N}$ from the end points of the chain. It has to be solved
358: with the boundary condition $a(n+N)=a(n)$.
359:
360: For an open chain, the solution of the Schr\"odinger equation
361: (\ref{eq_a}) can be sought in the form of plane
362: waves propagating between the chain boundaries and the defect,
363: %
364: \begin{equation}
365: \label{left_right}
366: a (n) =
367: C_{l,r}e^{i\theta n} + C'_{l,r} e^{- i\theta n}, \quad |n-n_0|\geq 1 .
368: \end{equation}
369: %
370: The subscripts $l$ and $r$ refer to the coefficients for the waves to
371: the left ($n<n_0$) and to the right ($n>n_0$) from the defect. The
372: interrelations between these coefficients and the coefficient $a(n_0)$
373: follow from the boundary conditions and from matching the solutions at
374: $n_0$. They are given by Eqs.~(\ref{bound_cond}) and
375: (\ref{boundary_defect}).
376:
377: For a closed chain, on the other hand, the solution can be sought
378: in the form
379: %
380: \begin{eqnarray}
381: \label{right}
382: a(n) &=& Ce^{i\theta n} + C'e^{-i\theta n}, \quad n_0<n\leq
383: N, \\
384: %
385: a(n) &=& Ce^{i\theta (n+N)} + C'e^{-i\theta (n+N)}, \quad 1\leq n
386: < n_0.\nonumber
387: \end{eqnarray}
388:
389:
390: The energy $E_1$ as a function of $\theta$ can be obtained by
391: substituting Eqs.~(\ref{left_right}) and (\ref{right}) into
392: Eq.~(\ref{eq_a}). Both for the open and closed chains it has the form
393: %
394: \begin{equation}
395: E_1 = {\varepsilon}_{1} +
396: J \cos \theta .
397: \label{energy_one}
398: \end{equation}
399:
400: The eigenfunctions (\ref{left_right}), (\ref{right}) with real
401: $\theta$ correspond to sinusoidal waves (extended states). From
402: (\ref{energy_one}), their energies lie within the band
403: ${\varepsilon}_{1} \pm J$. In contrast, localized states have complex
404: $\theta$, and their energies lie outside this band. The corresponding
405: solutions for both types of chains are discussed in the Appendix. In a
406: sufficiently long chain, there is always one localized one-excitation
407: state on a defect. Its energy is given by Eq.~(\ref{E_d}). In the case
408: of an open chain with the coupling anisotropy parameter $|\Delta|
409: > 1$, there are also localized states on the chain boundaries. Their
410: energy is given by Eq.~(\ref{E_s}).
411:
412:
413:
414: \section{Two excitations: unbound, bound, and localized states}
415:
416: A spin chain with a defect displays rich behavior in the presence
417: of two excitations. It is determined by the interplay between
418: disorder and inter-excitation coupling. Solutions of the
419: two-excitation problem have been obtained in the case of a
420: disorder potential of several special types, where the system is
421: integrable \cite{integrable}. Here we will study the presumably
422: nonintegrable but physically interesting problem where the
423: on-site energy of the defect differs from that of the host sites.
424:
425: The system is described by Hamiltonian (\ref{hamiltonian}).
426: In order to concentrate on the effects of
427: disorder rather than boundaries, we will consider a closed chain of
428: length $N$. We will also assume that the anisotropy is strong,
429: $|\Delta| \gg 1$.
430:
431: The wave function of a chain with two excitations is given by a linear
432: superposition
433: %
434: \begin{equation}
435: \label{two_exc}
436: \psi_2 = \sum_{n<m} a(n,m) \phi (n,m),
437: \end{equation}
438: %
439: where $\phi(n,m)$ is the state where spins on the sites $n$ and $m$
440: are pointing upward, whereas all other spins are pointing downward. In
441: a periodic chain, the sites with numbers that differ by $N$ are
442: identical, therefore we have $a(n,m) = a(m, n+N)$.
443:
444: From Eq.~(\ref{hamiltonian}), the Schr\"odinger equation for the
445: coefficients $a(n,m)$ is
446: %
447: \begin{widetext}
448: \begin{eqnarray}
449: \label{a(n,m)}
450: &&\left( 2{\varepsilon}_{1} + g \delta_{n,n_0}+ g \delta_{m,n_0}
451: +J\Delta \delta_{m,n+1}\right)a(n,m)
452: \nonumber \\
453: %
454: && +\frac{1}{2}J \bigl[a(n-1,m) + [a(n+1,m) +
455: a(n,m-1)](1-\delta_{m,n+1}) + a(n,m+1)\bigr] =E_2a(n,m).
456: \end{eqnarray}
457: \end{widetext}
458: %
459: Here, $E_2$ is the energy of a two-excitation state [${\varepsilon}_1 =
460: \varepsilon - J\Delta$ is the on-site one-excitation energy,
461: cf. Eq.~(\ref{eq_a})]. As before, we assume that the defect is located
462: on site $n_0$.
463:
464: \subsection{An ideal chain}
465:
466: In the absence of a defect the system is integrable. The solution of
467: Eq.~(\ref{a(n,m)}) can be found using the Bethe ansatz \cite{Bethe},
468: %
469: \begin{equation}
470: \label{two_bethe_ideal}
471: a(n,m) = C e^{i (\theta_1 n+ \theta_2 m)} +
472: C' e^{i (\theta_2 n + \theta_1 m)}.
473: \end{equation}
474: %
475: The energy of the state with given $\theta_1,\theta_2$ is obtained by
476: substituting Eq.~(\ref{two_bethe_ideal}) into Eq.~(\ref{a(n,m)})
477: written for $m>n+1$. This gives
478: %
479: \begin{equation}
480: \label{E2}
481: E_2 = 2{\varepsilon}_{1} + J (\cos \theta_1 + \cos \theta_2).
482: \end{equation}
483:
484: By requiring that the ansatz (\ref{two_bethe_ideal}) apply also for
485: $m=n+1$, we obtain an interrelation between the coefficients $C$ and $C'$,
486: %
487: \begin{equation}
488: \label{CC'}
489: \frac{C}{C'} =
490: -\frac{1 - 2 \Delta e^{i \theta_1 } + e^{i (\theta_1 + \theta_2)}}
491: {1 - 2 \Delta e^{i \theta_2 } + e^{i (\theta_1 + \theta_2)}}.
492: \end{equation}
493: %
494: With account taken of normalization, Eqs.~(\ref{two_bethe_ideal}) and
495: (\ref{CC'}) fully determine the wave function.
496: The states with real $\theta_{1,2}$ form a two-magnon band with width
497: $4|J|$, as seen from the dispersion relation (\ref{E2}). The magnons
498: are not bound to each other and propagate independently.
499:
500: For $|\Delta| > 1$, Eq.~(\ref{CC'}) also has a solution $C=0$, which
501: gives a complex phase $\theta_{2}= \theta_1^* = {1\over 2}\theta
502: -i\kappa$ with $\kappa > 0$. This solution corresponds to the wave
503: function $a(n,m)\propto \exp[i\theta(n+m)/2 - \kappa (m-n)]$;
504: we have $m>n$. From
505: Eqs.~(\ref{two_bethe_ideal})--(\ref{CC'}) we obtain
506: %
507: \begin{eqnarray}
508: &&e^{-\kappa}=\Delta^{-1}\cos(\theta/2),\quad E_{BP} = E_{BP}^{(0)}+
509: {J\over 2\Delta}\cos \theta,\nonumber\\
510: %
511: &&
512: E_{BP}^{(0)}= 2{\varepsilon}_1 + J\Delta + {J\over 2\Delta}.
513: \label{bound}
514: \end{eqnarray}
515:
516: Eq.~(\ref{bound}) describes a bound pair of excitations. Such a
517: pair can freely propagate along the chain. The wave function is
518: maximal when the excitations are on neighboring sites. The size of
519: the BP, i.e. the typical distance between the excitations, is
520: determined by the reciprocal decrement $\kappa^{-1}$, and ultimately
521: by the anisotropy parameter $\Delta$. For large $|\Delta|$, the
522: excitations in a BP are nearly completely bound to nearest sites. Then
523: the coefficients $a(n,m)\propto \delta_{n+1,m}$, to the lowest order
524: in $|\Delta|^{-1}$.
525:
526: The distance between the centers of the BP band and the two-magnon
527: band $E_{BP}^{(0)}-2\varepsilon_1$ is given approximately by the BP
528: binding energy $J\Delta$. The width of the BP band $|J/\Delta|$ is
529: parametrically smaller than the width of the two-magnon band $4|J|$,
530: see Fig.~\ref{fig:energies}.
531:
532: For nearest-neighbor coupling and for large $|\Delta|$, transport of
533: bound pairs can be visualized as occurring via an intermediate
534: step. First, one of the excitations in the pair makes a virtual
535: transition to the neighboring empty site, and as a result the parallel
536: spins in the pair are separated by one site. The corresponding state
537: differs in energy by $J\Delta$ from the bound-pair state. At the next
538: step the second spin can move next to the first, and then the whole
539: pair moves by one site. From perturbation theory, the bandwidth should
540: be $J^2/J\Delta\equiv J/\Delta$, which agrees with Eq.~(\ref{bound}).
541:
542: The above arguments can be made quantitative by introducing an
543: effective Hamiltonian $\tilde H^{(t)} = \sum_{n=1}^{N} \tilde
544: H_{n}^{(t)} $ of BP's. It is obtained from the Hamiltonian
545: $H^{(t)}$ (\ref{transfer_one}) in the second order of perturbation
546: theory in the one-excitation hopping constant $J$,
547: %
548: \begin{eqnarray}
549: \label{transfer_bound}
550: \tilde H_n^{(t)}&=&\frac{J }{64 \Delta }
551: [t^{(r)}_{n-1} t^{(l)}_{n-1} + t^{(l)}_{n+1} t^{(r)}_{n+1}]
552: \\
553: &+& \frac{J }{64 \Delta }
554: [t^{(l)}_{n} t^{(l)}_{n-1} + t^{(r)}_{n} t^{(r)}_{n+1}].
555: \nonumber
556: \end{eqnarray}
557: %
558: The operators $t^{(r,l)}_n$ of excitation hopping to the right or
559: left are given by Eq.~(\ref{transfer_one}). The first pair of terms in
560: $\tilde H_n^{(t)}$ [Eq.~(\ref{transfer_bound})] describes virtual
561: transitions in which a BP dissociates and then recombines on the same
562: site. This leads to a shift of the on-site energy level of the BP by
563: $J/2\Delta$. The second pair of terms describes the motion of a BP as
564: a whole to the left or to the right.
565:
566: The action of the Hamiltonian $\tilde H^{(t)}$ on the wave function
567: $a(n,n+1)$ is given by
568: %
569: \begin{eqnarray}
570: \label{transferBP}
571: &&\tilde H^{(t)}a(n,n+1)=\\
572: &&{J\over 2\Delta}a(n,n+1) + {J\over
573: 4\Delta}[a(n+1,n+2) + a(n-1,n)]. \nonumber
574: \end{eqnarray}
575: %
576: The Schr\"odinger equation for a BP is given by the sum of
577: the diagonal part (the first term) of Eq.~(\ref{a(n,m)}) with $m=n+1$
578: and the right-hand side of Eq.~(\ref{transferBP}). In this
579: approximation a BP eigenfunction is $a(n,m)=\delta_{n+1,m}\exp(i\theta
580: n)$, and the dispersion law is of the form (\ref{bound}).
581:
582: \subsection{Localized states in a chain with a defect}
583:
584: We now consider excitations in the presence of a defect. In this section we
585: assume that the defect excess energy $g$ is such that
586: %
587: \begin{equation}
588: \label{nonresonant_cond}
589: |g|\gg |J|, \quad |J\Delta-g| \gg |J| .
590: \end{equation}
591: The first inequality guarantees that the localization length of an
592: excitation on the defect is small [its inverse Im~$\theta_d\approx
593: \ln 2|g/J| \gg 1$ cf. Eq.~(\ref{one_localized})].
594:
595: The second condition in Eq.~(\ref{nonresonant_cond}) can be understood by
596: noticing that,
597: %, besides the bands of two-excitation states with unbound
598: %and bound excitations of an ideal chain,
599: in a chain with a defect, there is a two-excitation state where one
600: excitation is localized on the defect whereas the other is in an
601: extended magnon-type state. These excitations are not bound
602: together. The energy of such an unbound localized-delocalized pair
603: (LDP) should differ from the energy of unbound pair of magnons by
604: $\approx g$, and from the energy of a bound pair of magnons
605: (\ref{bound}) by $\approx J\Delta - g$. In this section we consider
606: the case where both these energy differences largely exceed the magnon
607: bandwidth $J$ (the case where $|J\Delta - g|\alt |J|$ will be
608: discussed in the following section).
609:
610: \subsubsection{Unbound localized-delocalized pairs (LDP's)}
611:
612: In the neglect of excitation hopping, the energy of an excitation
613: pair where one excitation is far from the defect ($|n-n_0|\gg 1$) and
614: the other is localized on the defect is $2{\varepsilon}_1 + g$. [This
615: can be seen from Eq.~(\ref{a(n,m)}) for the coefficients $a(n,m)$ in
616: which off-diagonal terms are disregarded.] At the same time, if one
617: excitation is on the defect and the other is on the neighboring site
618: $n_0\pm 1$, this energy becomes $2{\varepsilon}_1 + g + J\Delta$. The
619: energy difference $J\Delta$ largely exceeds the characteristic
620: bandwidth $J$. Therefore, if the excitation was initially far from the
621: defect, it will be reflected before it reaches the site $n_0\pm 1$.
622:
623: The above arguments suggest to seek the solution for the wave function
624: of an LDP in a periodic chain in the form
625: %
626: \begin{equation}
627: \label{LDP_wave}
628: a(n_0,m)=Ce^{i\theta m} + C'e^{-i\theta m} ,
629: \end{equation}
630: %
631: with the boundary condition $a(n_0, n_0 +1)= a(n_0,n_0+N-1) = 0$. This
632: boundary condition and the form of the solution are similar to what
633: was used in the problem of one excitation in an open chain.
634:
635: From Eq.~(\ref{a(n,m)}), the energy of an LDP is
636: %
637: \begin{equation}
638: \label{energy_LDP}
639: E_{LDP}\approx 2{\varepsilon}_1 + g + J\cos\theta .
640: \end{equation}
641: The wave number $\theta$ takes on $N-3$ values $\pi k/(N-2)$, with
642: $k=1,\ldots, N-3$.
643: As expected, the bandwidth of the LDP is $2|J|$, as in the case of
644: one-excitation band in an ideal chain.
645:
646: We have compared Eq.~(\ref{energy_LDP}) with numerical results
647: obtained by direct solution of the eigenvalue problem
648: (\ref{a(n,m)}). For $N=100, \Delta = 20, g/J=10$ we obtained excellent
649: agreement once we took into account that the energy levels
650: (\ref{energy_LDP}) are additionally shifted by $J^2/2g$. This shift
651: can be readily obtained from Eq.~(\ref{a(n,m)}) as the second-order
652: correction (in $J$) to the energy $E_d$ of the excitation localized on
653: the defect. It follows also from Eq.~(\ref{E_d}) for $|g|\gg |J|$.
654:
655: We note that the result is trivially generalized to the case of a
656: finite but small density of magnons. The wave function $a(n_0,
657: m_1,m_2,\ldots,m_M)$ of $M$ uncoupled magnons and an excitation on the
658: site $n_0$ is given by a sum of the appropriately weighted
659: permutations of $\exp(\pm i\theta_1m_1 \pm i\theta_2m_2 \ldots \pm
660: i\theta_Mm_M)$ over the site numbers $m_i$ (these numbers can be
661: arranged so that $m_1<\ldots <m_k<n_0<m_{k+1}<\ldots < m_M$), with
662: real $\theta_i$. The weighting factors for large $|\Delta|$ are found
663: from the boundary conditions and the condition that $a(n_0,
664: m_1,m_2,\ldots)=0$ whenever any two numbers $m_i, m_{i+1}$ differ by
665: 1. When the ratio of the number of excitations $M$ to the chain length
666: $N$ is small, the energy is just a sum of the energies of uncoupled
667: magnons and the localized excitation, i.e., it is
668: $(M+1){\varepsilon}_1 + g + J\sum\nolimits_{i=1}^M\cos\theta_i$. For
669: small $M/N$, scattering of a magnon by the excitation on a defect
670: occurs as if there were no other magnons, i.e., the probability of a
671: three-particle collision is negligibly small.
672:
673:
674: \subsubsection{Bound pairs localized on the defect: the doublet}
675:
676: For large $|\Delta|,|g/J|$, a bound pair of neighboring excitations
677: should be strongly localized when one of the excitations is on the
678: defect site. Indeed, if we disregard intersite excitation hopping,
679: the energy of a BP sitting on the defect is
680: $E_D^{(0)}=2{\varepsilon}_1+ g+ J\Delta $. It differs significantly
681: from the energy of freely propagating BP's (\ref{bound}), causing
682: localization.
683:
684: The major effect of the excitation hopping is that the pair can make
685: resonant transitions between the sites $(n_0, n_0+1)$ and $(n_0-1,
686: n_0)$. Such transitions lead to splitting of the energy level of the
687: pair into a doublet. To second order in $J$ the energies of the
688: resulting symmetric and antisymmetric states are
689: %
690: \begin{equation}
691: \label{doublet}
692: E_{D}^{(\pm)} = E_D^{(0)} + \frac{J(2J\Delta +g)}{4\Delta (J\Delta +g)}
693: \pm \frac{J^2}{4 (J\Delta + g)} .
694: \end{equation}
695: %
696: The energy splitting between the states is small if $J\Delta$ and $g$
697: have same sign. If on the other hand, $|J\Delta +g| \alt |J|$, the
698: theory has to be modified. Here, the bound pairs with one excitation
699: localized on the defect are resonantly mixed with extended unbound
700: two-magnon states. We do not consider this case in the present paper.
701:
702: We note that, in terms of quantum computing, the onset of a doublet
703: suggests a simple way of creating entangled states. Indeed, by
704: applying a $\pi$ pulse at frequency $\varepsilon_1+g$ one can
705: selectively excite the qubit $n_0$. If then a $\pi$ pulse is applied
706: at the frequency $E_D^{+}-(\varepsilon_1+g)$, it will selectively
707: excite a Bell state $\left[|01\rangle + |10\rangle\right]/\sqrt{2}$,
708: where $|ij\rangle$ describes the state where the qubits on sites
709: $n_0-1$ and $n_0+1$ are in the states $|i\rangle$ and $|j\rangle$,
710: respectively. The excitations on sites $n_0\pm 1$ can then be
711: separated without breaking their entanglement using two-qubit
712: gate operations.
713:
714:
715: \subsubsection{Localized states split off the bound-pair band}
716:
717: The energy difference $E_D^{(\pm)}-E_{BP}\approx g$ between BP states
718: on the defect site (\ref{doublet}) and extended BP states largely
719: exceeds the bandwidth $|J/\Delta|$ of the extended states. Therefore
720: it is a good approximation to assume that the wave functions of
721: extended BP states are equal to zero on the defect. In other words,
722: such BP's are reflected {\it before} they reach the defect. In this
723: sense, the defect acts as a boundary for them. One may expect that
724: there is a surface-type state associated with this boundary.
725:
726: The emergence of the surface-type state is facilitated by the
727: defect-induced change of the on-site energy of a BP located next to
728: the defect on the sites $(n_0+1, n_0+2)$ [or $(n_0-2,n_0-1)$]. This
729: change arises because virtual dissociation of a BP with one excitation
730: hopping onto a defect site gives a different energy denominator
731: compared to the case where a virtual transition is made onto a regular
732: site. It is described by an extra term $\delta E_{BP}$ in the
733: expression (\ref{transfer_bound}) for the diagonal part of the
734: Hamiltonian $\tilde H_n^{(t)}$ with $n=n_0+1$ [or $n=n_0-2$],
735: %
736: \begin{equation}
737: \label{delta_BP}
738: \delta E_{BP}= \frac{Jg}{4\Delta(J\Delta - g)}.
739: \end{equation}
740:
741: Using the transformed Hamiltonian (\ref{transferBP}), one can analyze
742: BP states in a way similar to the analysis of one-excitation states in
743: an open chain, see the Appendix. The Schr\"odinger equation for BP states
744: away from the defect is given by the first term in Eq.~(\ref{a(n,m)})
745: and by Eqs.~(\ref{transferBP}) and (\ref{delta_BP}). The BP wave functions
746: can be sought in the form
747: %
748: \begin{equation}
749: \label{BPfull}
750: a(n,n+1) = C e^{i\theta n} + C' e^{-i \theta n}.
751: \end{equation}
752: %
753: Then the BP energy as a function of $\theta$ is given by $E_{BP}=
754: E_{BP}^{(0)} + (J/2\Delta)\cos\theta$, cf. Eq.~(\ref{bound}).
755:
756: The values of $\theta$ can be found from the boundary condition
757: that the BP wave function is equal to zero on the defect, i.e.,
758: $a(n_0,n_0+1) = a(n_0-1,n_0) \equiv a(n_0+N-1,n_0+N)= 0$. From the
759: Schr\"odinger equations for $a(n,n+1)$ with $n=n_0+1$ and $n=N+n_0-2$,
760: we obtain an equation for $\theta$ of the form
761: %
762: \begin{eqnarray}
763: \label{theta_BP}
764: f(\theta ) = f(-\theta ),\; f(\theta ) = \left[\delta E_{BP} -
765: {J\over 4 \Delta} e^{i \theta }\right]^2e^{i\theta (N-3)}.
766: \end{eqnarray}
767:
768: Equation (\ref{theta_BP}) has $2(N-1)$ solutions for $\exp(i\theta)$. The
769: solutions $\exp(i\theta)=\pm 1$ are spurious, in the general case. The
770: roots $\theta$ and $-\theta$ describe one and the same wave
771: function. Therefore there are $N-2$ physically distinct roots
772: $\theta$, as expected for an $N$-spin chain with two excluded BP
773: states located at $(n_0, n_0\pm 1)$.
774:
775: Depending on the ratio $q=J/(4\Delta\,\delta E_{BP})\equiv (J\Delta -
776: g)/g$, the roots $\theta$ are either all real or there is one or two
777: pairs of complex roots with opposite signs. Real roots correspond to
778: extended states, whereas complex roots correspond to the states that
779: decay away from the defect. The onset of complex roots can be analyzed
780: in the same way as described in the Appendix for one excitation. There is
781: much similarity, formally and physically, between the onset of
782: localized BP states next to the defect and the onset of surface states
783: at the edge of an open chain.
784:
785: We
786: rewrite Eq.~(\ref{theta_BP}) in the form
787: %
788: \begin{equation}
789: \label{tangent_BP}
790: \tan[\theta(N-1)]={2\sin\theta \,(\cos \theta - q)\over (\cos\theta -
791: q)^2 - \sin^2\theta}.
792: \end{equation}
793: %
794: For $|q| > 1$ all roots of Eq.~(\ref{tangent_BP}) are real. At $|q|=1$
795: there occurs a bifurcation where two real roots with opposite signs
796: merge (at $\theta = 0$, for $q=1$, or at $\theta = \pi$, for
797: $q=-1$). They become complex for $|q| < 1$. For $|q| = 1-2(N-1)^{-1}$
798: two other real roots coalesce at $\theta = 0$ or $\pi$ and another
799: pair of complex roots emerges.
800:
801: As the length of the chain increases, the difference between the pairs
802: of complex roots decreases. In the limit $N\to \infty$ the roots merge
803: pairwise. The imaginary part of one of the roots is%
804: \begin{equation}
805: \label{BP_s}
806: {\rm Im}~\theta_{BP}^{(s)}= \ln |\delta E_{BP}/(J/4\Delta )|.
807: \end{equation}
808: %
809: The second root has opposite sign.
810:
811:
812: One can show from the Schr\"odinger equation for the BP's that the
813: solution (\ref{BP_s}) describes a ``surface-type'' BP state localized
814: next to the defect on sites $(n_0+1, n_0+2)$. The solution with
815: $-\theta_{BP}^{(s)}$ describes the surface state on
816: $(n_0-2,n_0-1)$. The amplitudes of these states exponentially decay
817: away from the defect. For $J\Delta >g> J\Delta/2 > 0$ or $J\Delta <g<
818: J\Delta/2 < 0$ we have Re~$\theta_{BP}^{(s)}=0$. For $g> J\Delta > 0$
819: or $g< J\Delta < 0$, we have Re~$\theta_{BP}^{(s)} =\pi$, and decay of
820: the localized state is accompanied by oscillations. The complex roots
821: in a finite chain can be pictured as describing those same states on
822: the opposite sites of the defect. But now the states are
823: ``tunnel'' split because of the overlap of their tails inside the
824: chain, which leads to the onset of two slightly different localization
825: lengths.
826:
827: The energy of the localized state in a long chain is
828: %
829: \begin{equation}
830: \label{BP_s_energy}
831: E_{BP}^{(s)}= E_{BP}^{(0)}+ \delta E_{BP} +
832: \frac{(J/4\Delta )^2}{\delta E_{BP}}.
833: \end{equation}
834: %
835: It lies outside the band $ E_{BP}^{(0)}\pm (J/2\Delta)$ (\ref{bound})
836: of the extended BP states. The distance to the band edge strongly
837: depends on the interrelation between the defect excess energy $g$ and
838: the BP binding energy $J\Delta$. It is of the order of the BP
839: bandwidth $|J/\Delta|$, except for the range where the difference
840: between $J\Delta$ and $g$ becomes small. In this range the energy of
841: the surface-type state sharply increases in the absolute value. This
842: is illustrated in Fig.~\ref{fig:BP_s}. The localization length $|{\rm
843: Im}~\theta_{BP}^{(s)}|^{-1}$ is large when the state energy is close
844: to the band edge and shrinks down with decreasing $|J\Delta -g|$,
845: i.e., with increasing $|\delta E_{BP}|$.
846:
847:
848: \begin{figure}[ht]
849: \includegraphics[width=2.8in]{fig2_XXZ.eps}
850: \caption{The distance between the energy level of the BP localized
851: next to the defect and the center of the band of extended BP states,
852: $\varepsilon_{BP} = (E_{BP}^{(s)} - E_{BP}^{(0)})/(J/4\Delta) $, vs
853: the scaled defect excess energy $g/J\Delta$. The localized state
854: exists in an infinite chain for $g/J\Delta> 1/2$. The vertical and
855: horizontal dot-dashed lines show the asymptotes for $g\to J\Delta$ and
856: $g/J\Delta\to\infty$, respectively. The results refer to the range
857: where the BP and LDP bands are far from each other, compared to the
858: LDP bandwidth $|J|$. These bands are sketched in the inset (the BP band
859: is above the LDP band for $J\Delta/g > 1$). The dashed line in the
860: inset shows where the energy level of the localized BP state is
861: located with respect to the bands.}
862: \label{fig:BP_s}
863: \end{figure}
864:
865:
866: \section{Antiresonant decoupling of two-excitation states}
867:
868: The analysis of the preceding section does not apply if the pair
869: binding energy $J\Delta$ is close to the defect excess energy $g$.
870: When $|g-J\Delta|$ is of the order of the LDP bandwidth $|J|$, the BP
871: and LDP states are in resonance, their bands overlap or nearly overlap
872: with each other. One might expect that there would occur mixing of
873: states of these two bands. In other words, a delocalized magnon in the
874: LDP band might be scattered off the excitation on the defect, and as a
875: result they both would move away as a bound pair. However, as we show,
876: such mixing does not happen. In order to simplify notations we will
877: assume in what follows that $g,J,\Delta >0$.
878:
879:
880: \subsection{A bound pair localized next to the defect}
881:
882: In the resonant region we should reconsider the analysis of the
883: next-to-the-defect bound pair localized on sites $(n_0+1, n_0+2)$ or
884: $(n_0-2,n_0-1)$. As $|g-J\Delta|$ decreases, the energy of this pair
885: (\ref{BP_s_energy}) moves away from the BP band, see
886: Fig.~\ref{fig:BP_s}. At the same time, the distance between the pair
887: energy and the LDP band $E_{BP}^{(s)}- 2\varepsilon_1-g$ becomes {\it
888: smaller} with decreasing $|g-J\Delta|$ as long as $|J\Delta - g|>
889: |J|/2$. For $|J\Delta - g|\sim J$, the next-to-the-defect pairs are
890: hybridized with LDP's. The hybridization occurs in first order in the
891: nearest-neighbor coupling constant $J$.
892:
893: To describe the hybridization, we will seek
894: the solution of the Schr\"odinger equation (\ref{a(n,m)}) in the form
895: of a linear superposition of an LDP state (\ref{LDP_wave}) and a
896: pair on the next to the defect sites,
897: %
898: \begin{eqnarray}
899: \label{hybridization}
900: a(n,m)&=& (Ce^{i\theta m} +C'e^{-i\theta m}) \delta_{n,n_0} \nonumber\\ &+&
901: a(n_0+1,n_0+2) \delta_{n,n_0+1}\delta_{m,n_0+2} \nonumber\\
902: &+&
903: a(n_0-2,n_0-1) \delta_{n,n_0-2}\delta_{m,n_0-1}
904: \end{eqnarray}
905: (we remind that $m > n$). The energy of a state with given $\theta$
906: can be found from (\ref{a(n,m)}) as before by considering $m$ far away
907: from $n_0$. It is given by Eq.~(\ref{energy_LDP}).
908:
909: The interrelation between the coefficients $C, C', a(n_0+ 1, n_0+2)$,
910: and $a(n_0-2, n_0-1) $, as well as the values of $\theta$ should be
911: obtained from the boundary conditions. These conditions follow from
912: the fact that the energy of a pair on sites $(n_0, n_0\pm 1)$ and the
913: energies of the pairs described by Eq.(\ref{hybridization}) differ by
914: $\sim g\approx J\Delta$, cf. (\ref{doublet}). Therefore the pairs on
915: sites $(n_0, n_0\pm 1)$ are decoupled from the states
916: (\ref{hybridization}), and in the analysis of the LDP's we can set
917: $a(n_0, n_0\pm 1)=0$. Decoupled also are unbound two-excitation states
918: with no excitation on the defect, i.e., two-magnon states. Therefore
919: $a(n,m)=0$ if simultaneously $m-n>1$ and $(n-n_0)(m-n_0)\neq 0$.
920:
921: From Eq.~(\ref{a(n,m)}) written for $n=n_0,m=n_0+ 2$ and
922: $n=n_0-2,m=n_0$ [with account taken of the relation $a(n,m)=a(m,n+N)$]
923: we obtain
924: %
925: \begin{eqnarray}
926: \label{hybr_interrelation}
927: a(n_0+1, n_0+2)&=& Ce^{i\theta (n_0+1)} +C'e^{-i\theta (n_0+1)},\\
928: a(n_0-2, n_0-1)&=& Ce^{i\theta (n_0+N-1)} +C'e^{-i\theta (n_0+N-1)}\nonumber .
929: \end{eqnarray}
930:
931: The interrelation between $C$ and $C'$ and the equation for $\theta$
932: follow from Eqs.~(\ref{hybridization}), (\ref{hybr_interrelation})
933: and (\ref{a(n,m)}) written
934: for $n=n_0+1,m=n_0+ 2$ and $n=n_0-2,m=n_0-1$. They have the form
935: %
936: \begin{equation}
937: \label{hybr_cc'}
938: C' = -\frac{2(J \Delta -g) e^{i\theta } -J}
939: {2(J \Delta-g) e^{-i\theta } -J} Ce^{2i\theta n_0},
940: \end{equation}
941: %
942: and
943: %
944: \begin{eqnarray}
945: \label{hybr_theta}
946: &\tilde f(\theta)=\tilde f(-\theta),\\
947: &\tilde f(\theta)=e^{i \theta N} \left[2(J \Delta-g)e^{-i\theta} -
948: J \right]^2.\nonumber
949: \end{eqnarray}
950:
951: Equation (\ref{hybr_theta}) is an $2N$th order equation for
952: $\exp(i\theta)$. Its analysis is completely analogous to that of
953: Eq.~(\ref{theta_BP}). The roots $\theta=0,\pi$ are spurious, and the
954: roots $\theta$ of opposite signs describe one and the same wave
955: function. Therefore Eq.~(\ref{hybr_theta}) has $N-1$ physically
956: distinct roots $\theta$. Real $\theta$'s correspond to extended
957: states. Complex roots appear for $2|J \Delta-g|
958: >J$. These roots, $\theta_{LDP}^{(s)}$, describe localized states. In
959: the limit of a long chain, $N\to \infty$,
960: the imaginary part of one of them is
961: %
962: \begin{equation}
963: \label{hybr_decrem}
964: {\rm Im}\, \theta_{LDP}^{(s)} = \ln |2(J\Delta - g)/J|,
965: \end{equation}
966: whereas the other root has just opposite sign.
967:
968: The wave function of the localized state is maximal either on sites
969: $(n_0+1, n_0+2)$ or $(n_0-2,n_0-1)$ and exponentially decays into the
970: chain. For $J\Delta - g < 0$ this decay is accompanied by
971: oscillations, Re~$\theta_{LDP}^{(s)}=\pi$. The energy of the
972: localized state is
973: %
974: \begin{equation}
975: \label{hybr_s_energy}
976: E_{LDP}^{(s)} = 2{\varepsilon}_1 + J \Delta + \frac{(J/2)^2}{J\Delta - g}.
977: \end{equation}
978: %
979:
980: The localized state (\ref{hybridization}), (\ref{hybr_decrem}) is a
981: ``surface-type'' state induced by the defect. It is the resonant-region
982: analog of the localized next-to-the-defect state discussed in
983: Section~III.B3. The wave function of the latter state (\ref{BPfull}),
984: (\ref{BP_s}) was a linear combination of the wave functions of bound
985: pairs. In contrast, the state given by Eqs.~(\ref{hybridization}),
986: (\ref{hybr_decrem}) is a combination of the wave functions of the bound
987: pair located next to the defect and a localized-delocalized pair.
988:
989: The evolution of the surface-type state is controlled by the
990: difference between the excess energies of binding two excitations in a
991: pair or localizing one of them on the defect $|J\Delta - g|$. As
992: $|J\Delta - g|$ varies, the state changes in the following way. It
993: first splits off the BP band when $|J\Delta - g|$ becomes less than
994: $g$, see Fig.~\ref{fig:BP_s}. Its energy moves away from the band of
995: extended BP states with decreasing $|J\Delta - g|$ and the
996: localization length decreases [cf. Eq.~(\ref{BP_s})]. Well before
997: $|J\Delta - g|$ becomes of order $J$, the state becomes strongly
998: localized on the sites $(n_0+1, n_0+2)$ or $(n_0-2, n_0-1)$.
999:
1000: In the region $|J\Delta - g|\sim J$ the localized state becomes
1001: stronger hybridized with LDP states than with extended BP states. This
1002: hybridization occurs in first order in $J$, via a transition
1003: $(n_0+1,n_0+2) \to (n_0, n_0+2)$ [or $(n_0-2,n_0-1) \to
1004: (n_0-2,n_0)$]. In this region the localization length increases with
1005: decreasing $|J\Delta - g|$, cf. Eq.~(\ref{hybr_decrem}). Ultimately,
1006: for $|J\Delta - g|=J/2$ the localized surface-type state disappears,
1007: as seen in Fig.~\ref{fig:LDP}. The evolution of the energy of the
1008: localized state with $J\Delta - g$ is shown in Fig.~\ref{fig:LDP}.
1009:
1010: \begin{figure}[ht]
1011: \includegraphics[width=2.8in]{fig3_XXZ.eps}
1012: \caption{The energy difference between the localized BP state and the
1013: LDP band, $\varepsilon_{LDP} = (E_{LDP}^{(s)} - 2{\varepsilon}_1 -
1014: g)/(J/2)$, vs the energy mismatch $2(J\Delta -g)/J$.
1015: In the region between the dot-dashed lines the BP state is
1016: delocalized. The results refer to the case where the BP and LDP bands
1017: are close to each other or overlap, as sketched in the inset (the BP band
1018: is above the LDP band for $J\Delta > g$). The dashed line in the
1019: inset shows where the energy level of the localized BP
1020: state is located with respect to the bands.}
1021: \label{fig:LDP}
1022: \end{figure}
1023:
1024:
1025:
1026:
1027: The crossover from hybridization of the localized surface-type BP
1028: state with extended BP states to that with LDP states occurs in the
1029: region $|g|\gg |J\Delta - g| \gg |J|$. It is described using a
1030: different approach in Ref.~\onlinecite{us_DS}. We note that the
1031: expressions for the energy of the localized states (\ref{BP_s_energy})
1032: and (\ref{hybr_s_energy}) go over into each other for $|J\Delta -
1033: g|\ll |g|$.
1034:
1035: \subsection{Decoupling of bound pairs and LDP's
1036: in the resonant region}
1037:
1038: We are now in a position to consider resonant coupling between
1039: extended states of bound pairs and localized-delocalized pairs. To
1040: lowest order in $\Delta^{-1}$, a BP on sites $(n, n+1)$ far away from
1041: the defect can resonantly hop only to the nearest pair of sites, see
1042: Eqs.~(\ref{transfer_bound}), (\ref{transferBP}). As noted before, the
1043: hopping requires an intermediate virtual transition of the BP into a
1044: dissociated state, which differs in energy by $ J\Delta$.
1045:
1046: A different situation occurs for the BP on the sites $(n_0+2,n_0+3)$
1047: [or $(n_0-3,n_0-2)$]. Such BP can hop onto the sites $(n_0+1,n_0+2)$
1048: and $(n_0+3,n_0+4)$, as described by (\ref{transferBP}). But in
1049: addition, for $|J\Delta - g|\alt J$ it can make a
1050: transition into the LDP state on sites $(n_0,n_0+3)$
1051: [or $(n_0-3,n_0)$] . Indeed, such
1052: state has the energy $\approx 2{ \varepsilon}_1+g$, which is
1053: close to the BP energy $\approx 2{ \varepsilon}_1+J\Delta$.
1054:
1055: The transition $(n_0+2,n_0+3) \to (n_0,n_0+3)$ goes through the
1056: intermediate dissociated state $(n_0+1,n_0+3)$, which differs in
1057: energy by $\approx J\Delta$. It can be taken into account by adding
1058: the term $\delta\tilde H^{(t)}$ to the BP hopping Hamiltonian
1059: (\ref{transferBP}) for the sites $(n_0+2,n_0+3)$,
1060: %
1061: \begin{equation}
1062: \label{extra_H_t}
1063: \delta \tilde H^{(t)}a(n_0+2,n_0+3) = (J/4\Delta)a(n_0,n_0+3).
1064: \end{equation}
1065:
1066: Extended BP states are connected to LDP states only through a BP on
1067: the sites $(n_0+2,n_0+3)$. We are now in a position to analyze this
1068: connection. From Eqs.~(\ref{transferBP}) and (\ref{extra_H_t}), we see
1069: that
1070: %
1071: \begin{eqnarray}
1072: \label{coupling_resonant}
1073: &&[\tilde H^{(t)} + \delta \tilde H^{(t)}]a(n_0+2,n_0+3) =
1074: (J/4\Delta)\nonumber\\
1075: &&\times[a(n_0,n_0+3) + a(n_0+1,n_0+2)] + A,
1076: \end{eqnarray}
1077: %
1078: where $A$ is a linear combination of the amplitudes $a(n_0+2,n_0+3)$
1079: and $a(n_0+3,n_0+4)$.
1080:
1081: The sum $a(n_0,n_0+3)+ a(n_0+1,n_0+2)$ in
1082: Eq.~(\ref{coupling_resonant}) can be expressed in terms of the LDP
1083: wave functions (\ref{hybridization}). With account taken of the
1084: interrelation (\ref{hybr_cc'}) between the coefficients $C,C'$ in
1085: Eq.~(\ref{hybridization}), we have
1086: %
1087: \begin{eqnarray}
1088: \label{amplitude_resonant}
1089: &&a(n_0+1,n_0+2)+a(n_0,n_0+3)\nonumber\\
1090: &&\propto C \sin \theta \cos \theta [J\Delta -g -
1091: J \cos \theta ].
1092: \end{eqnarray}
1093: %
1094: Equations (\ref{coupling_resonant}), (\ref{amplitude_resonant}) describe
1095: the coupling between the BP on the sites $(n_0+2,n_0+3)$ and the LDP
1096: eigenstates with given $\theta$.
1097:
1098: An important conclusion can now be drawn regarding the behavior of BP
1099: and LDP states in the resonant region. The center of the BP band lies
1100: at $2{\varepsilon}_1 + J\Delta +(J/2\Delta)$ (\ref{bound}), and the BP
1101: band is parametrically narrower than the LDP band $2{\varepsilon}_1+g
1102: +J\cos\theta$ (\ref{energy_LDP}). When the BP band is inside the LDP
1103: band, this means that, for an appropriate wave number $\theta$ of the
1104: LDP magnon, $J\Delta = g+J\cos\theta$, to zeroth order in
1105: $\Delta^{-1}$ (this is the approximation used to obtain the LDP
1106: dispersion law). It follows from Eqs.~(\ref{coupling_resonant}) and
1107: (\ref{amplitude_resonant}) that, for such $J\Delta -g$ and $\theta$
1108: there is no coupling between the LDP and extended BP states.
1109:
1110: The above result means that LDP and extended BP states do not experience
1111: resonant scattering into each other, even though it is allowed by the
1112: energy-conservation law. Such scattering would correspond to the
1113: scattering of a magnon off the excitation localized on the
1114: defect, with both of them becoming a bound pair that moves away from
1115: the defect, or an inverse process.
1116:
1117: Physically, the antiresonant decoupling of BP and LDP states is a
1118: result of strong mixing of a BP on the sites $(n_0+1,n_0+2)$ and
1119: LDP's. Because of the mixing, the amplitudes of transitions of extended
1120: BP states to the sites $(n_0+1,n_0+2)$ and $(n_0,n_0+3)$ compensate
1121: each other, to lowest order in $\Delta^{-1}$.
1122:
1123: To illustrate the antiresonant decoupling we show in
1124: Fig.~\ref{fig:temporal_nonresonant} two types of time evolution of
1125: excitation pairs. In both cases the initial state of the system was
1126: chosen as a pair on the sites $(n_0+1,n_0+2)$, i.e., $a(n_0+1,n_0+2)=1$
1127: for $t=0$. The time-dependent Schr\"odinger equation was then solved
1128: with the boundary condition that corresponds to a closed chain.
1129:
1130: The solid lines refer to the case of nonoverlapping BP and LDP bands,
1131: $|J|\ll |g-J\Delta|$ and $|g|<|J\Delta|/2$. In this case there does
1132: not emerge a localized surface-type BP state next to the
1133: defect. Therefore an excitation pair placed initially on the sites
1134: $(n_0+1,n_0+2)$ resonantly transforms into extended BP states and
1135: propagates through the chain. This propagation is seen from the figure
1136: as oscillations of the return probability and the probability to find
1137: the BP on another arbitrarily chosen pair of neighboring sites
1138: $(n_0+2,n_0+3)$. The oscillation rate should be small, of the order
1139: of the bandwidth $J/\Delta$. This estimate agrees with the numerical
1140: data. It is seen that the BP state is not transformed into LDP
1141: states. The amplitude of LDP states on sites $(n_0,m\geq n_0+2)$
1142: remains extremely small, as illustrated for $m=n_0+2$.
1143:
1144: The dotted lines in Fig.~\ref{fig:temporal_nonresonant} show a
1145: completely different picture which arises when the BP band is inside
1146: the LDP band. In this case an excitation pair placed initially on
1147: $(n_0+1,n_0+2)$ hybridizes with LDP rather than BP states. A
1148: transformation of the pair $(n_0+1,n_0+2)$ into LDP's with increasing
1149: time is clearly seen. The period of oscillations is of the order of
1150: the reciprocal bandwidth of the LDP's $J^{-1}$, it is much shorter than in
1151: the previous case. Remarkably, as a consequence of the antiresonance,
1152: extended states of bound pairs are not excited to any appreciable
1153: extent, as seen from the amplitude of the pair on the sites
1154: $(n_0+2,n_0+3)$.
1155:
1156: \begin{figure}[ht]
1157: \includegraphics[width=3.2in]{fig4_XXZ.eps}
1158: \caption{Time evolution of a two-excitation wave packet in an $XXZ$
1159: chain with a defect; $|a(n,m)|^2$ is the occupation of sites
1160: $(n,m)$. Initially an excitation pair is placed next to the defect on
1161: sites $(n_0+1,n_0+2)$. The solid and dashed lines refer to the cases
1162: where the BP and LDP bands are, respectively, far away from each other
1163: ($g=J\Delta/4$) and overlapping ($g=J\Delta$). In the first case a
1164: bound pair slowly oscillates between neighboring sites $(n,n+1)$ and
1165: does not dissociate [$a(n_0,n_0+2)$ remains very small]. In the case
1166: of overlapping bands, the pair at $(n_0+1,n_0+2)$ is hybridized with
1167: LDP's, but practically does not mix with bound pairs on other sites
1168: [$a(n_0+2,n_0+3)$ remains very small]. The results refer to a ten-site
1169: closed chain with $\Delta =10$.}
1170: \label{fig:temporal_nonresonant}
1171: \end{figure}
1172:
1173:
1174: \section{Conclusions}
1175:
1176: We have analyzed the dynamics of a disordered spin chain with a
1177: strongly anisotropic coupling in a magnetic field. A defect in such a
1178: chain can lead to several localized states, depending on the number of
1179: excitations. This is a consequence of the interaction between
1180: excitations and its interplay with the disorder. We have studied
1181: chains with one and two excitations.
1182:
1183: The major results refer to the case of two excitations. Here, the
1184: physics is determined by the interrelation between the excess on-site
1185: energy of the defect $g$ and the anisotropic part of the exchange
1186: coupling $J\Delta$. Strong anisotropy leads to binding of excitations
1187: into nearest-neighbor pairs that freely propagate in an ideal
1188: chain. Because of the defect, BP's can localize. A simple
1189: type of a localized BP is a pair with one of the excitations located
1190: on a defect. A less obvious localized state corresponds to a pair
1191: localized next to a defect. It reminds of a surface state split off
1192: from the band of extended BP states, with the surface being the defect
1193: site. We specified the conditions where the localization occurs and
1194: found the characteristics of the localized states.
1195:
1196: Our most unexpected observation is the antiresonant decoupling of
1197: extended BP states from localized-delocalized pairs. The LDP's are
1198: formed by one excitation on the defect site and another in an extended
1199: state. The antiresonance occurs for $g\approx J\Delta$, when the BP and
1200: LDP bands overlap. It results from destructive quantum interference of
1201: the amplitudes of transitions of BP's into two types of resonant
1202: two-excitation states: one is an LDP, and the other is an excitation
1203: pair on the sites next to the defect. As a result of the
1204: antiresonant decoupling, extended BP's and LDP's do not \
1205: scatter into each other, even though the
1206: scattering is allowed by energy conservation. This means that an
1207: excitation localized on the defect does not delocalize as a result of
1208: coupling to other excitations.
1209:
1210: The occurrence of multiple localized states in the presence of other
1211: excitations is important for quantum computing. It shows that, even
1212: where the interaction between the qubits is ``on'' all the time, we
1213: may still have well-defined states of individual qubits that can be
1214: addressed and controlled. One can prepare entangled localized pairs of
1215: excitations, as we discussed in Sec.~III~B~2, or more complicated
1216: entangled excitation complexes. The results of the paper also provide
1217: an example of new many-body effects that can be studied using quantum
1218: computers with individually controlled qubit transition energies.
1219:
1220: \begin{acknowledgments}
1221: This research was supported in part by the NSF through Grant
1222: No. ITR-0085922 and by the Institute for Quantum Sciences at Michigan
1223: State University.
1224: \end{acknowledgments}
1225:
1226:
1227: \appendix
1228:
1229: \section{One excitation}
1230:
1231: In an infinite spin chain in a magnetic field, the anisotropy of the
1232: spin-spin interaction does not affect the spectrum and wave functions
1233: of one excitation. The matrix element of the term
1234: $\sum_n\sigma_n^z\sigma_{n+1}^z$ in the Hamiltonian
1235: (\ref{hamiltonian}) is just a constant. However, the situation becomes
1236: different for a chain of finite length, because the coupling
1237: anisotropy can lead to surface states. In the case of two excitations,
1238: analogs of surface-type states emerge near defects
1239: in an infinite chain, as discussed in Sec.~III.
1240: Here, for completeness and keeping
1241: in mind a reader with the background in quantum computing, we briefly
1242: outline the results of the standard analysis of a finite-length spin
1243: chain with one excitation.
1244:
1245: \subsection{Localized surface and defect-induced states in an open chain}
1246:
1247: The Schr\"odinger equation for an excitation in an open chain has the
1248: form (\ref{eq_a}), and its solution $a(n)$ to the left and to the
1249: right from the defect can be written in the form of a superposition of
1250: counterpropagating plane waves, Eq.~(\ref{left_right}). The relation
1251: between the amplitudes of these waves $C'_{l,r}$ and $C_{l,r}$ follows
1252: from the boundary conditions $a(0)=a(N+1)=0$. By substituting
1253: Eq.~(\ref{left_right}) into Eq.~(\ref{eq_a}) with $n=1$ and $n=N$, we
1254: obtain
1255: %
1256: \begin{eqnarray}
1257: \label{bound_cond}
1258: &&C'_l = - D C_l, \quad C'_r = -D^{-1} e^{2i\theta (N+1)} C_r,\\
1259: &&D = [1-\Delta \exp (i\theta )]/[1-\Delta \exp (-i \theta )].\nonumber
1260: \end{eqnarray}
1261: %
1262:
1263: The relations between $C'_{l,r}$, $C_{l,r}$, and the amplitude of the
1264: wave function on the defect site $a(n_0)$ follow from
1265: Eqs.~(\ref{eq_a}) and (\ref{left_right}) for $n=n_0\pm 1$,
1266: %
1267: \begin{eqnarray}
1268: \label{boundary_defect}
1269: a(n_0)&&= C_{l} e^{i\theta n_0}\left[1 - D e^{-2i\theta n_0}\right]\nonumber\\
1270: &&= C_{r} e^{i\theta n_0}\left[1 - D^{-1}
1271: e^{2i\theta (N+1-n_0)}\right] .
1272: \end{eqnarray}
1273: %
1274: With (\ref{bound_cond}) and (\ref{boundary_defect}), all coefficients in
1275: the wave function (\ref{left_right}) are expressed in terms of one
1276: number, $a(n_0)$. It can be obtained from normalization.
1277:
1278: In a finite chain, the values of $\theta$ are quantized. They can be
1279: found from Eq.~(\ref{eq_a}) with $n=n_0$. With account taken of
1280: (\ref{boundary_defect}), this equation can be written as
1281: %
1282: \begin{eqnarray}
1283: \label{theta}
1284: f_N (\theta ) - D^2 f_N (-\theta ) = -(ig/J\sin \theta )
1285: \qquad\qquad\qquad\nonumber\\
1286: \times [
1287: f_N (\theta ) + D^2 f_N (-\theta ) - 2D
1288: \cos \theta (N-2 n_0 +1) ],
1289: \end{eqnarray}
1290: %
1291: where
1292: %
1293: \begin{equation}
1294: f_N (\theta ) = \exp[i\theta (N+1)] .
1295: \end{equation}
1296:
1297: %
1298: The analysis of the roots of Eq.~(\ref{theta}) is standard. This is a
1299: $2(N+1)$-order equation for $\exp(i\theta)$, but its solutions for
1300: $\theta$ come in pairs $\theta$ and $-\theta$. Each pair gives one
1301: wave function, as seen from Eq.~(\ref{left_right}). In addition,
1302: Eq.~(\ref{theta}) has roots $\theta = 0,\pi$; they are spurious
1303: (unless $|\Delta| = 1$ and $g= 0$) and appear as a result of algebraic
1304: transformations. Therefore Eq.~(\ref{theta}) has $N$ physically
1305: distinct roots, as expected for a chain of $N$ spins. We note that,
1306: for $g=0$ the position of the impurity $n_0$ drops out from
1307: Eq.~(\ref{theta}), and then the equation goes over into the result for
1308: an ideal chain.
1309:
1310: Solutions of Eq.~(\ref{theta}) with real $\theta$ correspond, in the
1311: case of a long chain, to delocalized magnon-type excitations
1312: propagating in the chain. Their bandwidth is $2|J|$.
1313:
1314: Along with delocalized states, Eq.~(\ref{theta}) describes also
1315: localized states with complex $\theta$. Complex roots of
1316: Eq.~(\ref{theta}) can be found for a long chain, where $|{\rm
1317: Im}\,\theta|(N-n_0),\, |{\rm Im}\,\theta|n_0 \gg 1$. They describe
1318: surface states, which are localized on the chain boundaries, and a
1319: state localized on the defect. The localization length of the states
1320: is given by $|1/{\rm Im}\,\theta|$.
1321:
1322: The surface states arise only for the
1323: anisotropy parameter $|\Delta|>1$. The corresponding values of
1324: $\theta$ are
1325: %
1326: \begin{equation}
1327: \label{bound_local}
1328: \theta_s = \pm i\ln|\Delta| + \pi\Theta(-\Delta).
1329: \end{equation}
1330: %
1331: Here, the signs $+$ and $-$ refer to the states localized on the
1332: left and right boundaries, respectively, and $\Theta(x)$ is the step
1333: function.
1334:
1335: From Eq.~(\ref{energy_one}), the energy of the surface state is
1336: %
1337: \begin{equation}
1338: \label{E_s}
1339: E_s={\varepsilon}_1 + J(\Delta^2 +1)/2\Delta
1340: \end{equation}
1341: %
1342: It lies outside the energy band of delocalized excitations. We note
1343: that, for $\Delta > 1$ the surface states decay monotonically with the
1344: distance from the boundary (Re~$\theta=0$). For sufficiently large
1345: negative $\Delta$, on the other hand, the decay of the wave function
1346: is accompanied by oscillations, and $a(n)$ changes sign from site to site.
1347:
1348: A defect in a long chain gives rise to a localized one-spin excitation
1349: for an arbitrary excess energy $g$ \cite{Economou}. The amplitude
1350: $a(n)$ decays away from the defect as
1351: %
1352: \begin{eqnarray}
1353: \label{one_localized}
1354: &a(n)=a(n_0)\exp(i\theta_d|n-n_0|),\\
1355: &\theta_d=i\sinh^{-1}(|g/J|)+
1356: \pi\Theta(-g/J).\nonumber
1357: \end{eqnarray}
1358: %
1359: The energy of the localized state is
1360: %
1361: \begin{equation}
1362: \label{E_d}
1363: E_d=\varepsilon_1+
1364: (g^2+J^2)^{1/2}{\rm sgn}\,g.
1365: \end{equation}
1366:
1367: For small $|g/J|$, we have Im~$\theta_d \approx |g/J|$, i.e., the
1368: reciprocal localization length is simply proportional to the defect
1369: excess energy $|g|$. In the opposite case of large $|g/J|$ we have
1370: Im~$\theta_d = \ln |2g/J|$. In this case the amplitude of the
1371: localized state rapidly falls off with the distance from the defect,
1372: $a(n)\propto (2|g/J|)^{-|n-n_0|}$.
1373:
1374: If the localization length is comparable to the chain length, the
1375: notion of localization is not well defined. However,
1376: when discussing numerical results, one can formally call
1377: a state localized if its wave function exponentially decays away from
1378: the defect and is described by a solution of Eq.~(\ref{theta}) with
1379: complex $\theta$. This is equivalent to the statement that the state
1380: energy $E_d$ lies outside the band of magnons in the infinite
1381: chain. In an open finite chain such localized state may emerge
1382: provided the localization length is smaller than the distance from the
1383: defect to the boundaries. This means that the defect excess energy
1384: $|g|$ should exceed a minimal value that depends on the size of the
1385: chain. The comparison of Eq.~(\ref{one_localized}) with the numerical
1386: solutions of the full equation (\ref{theta}) for a finite chain is shown in
1387: Fig.~\ref{fig:loc_length}.
1388:
1389: \begin{figure}[ht]
1390: \includegraphics[width=3.2in]{fig5_XXZ.eps}
1391: \caption{ The reciprocal localization length Im~$\theta_d$ for an
1392: infinite chain [Eq.~(\protect\ref{one_localized})] as a function of
1393: the defect excess energy $g$ for $J=1$ (solid line). Also shown are
1394: the results for an open chain with
1395: the number of sites $N=6$ and 12 and $n_0=N/2$ (stars and crosses,
1396: respectively) and a closed chain with 6 and 12 sites (circles and
1397: squares, respectively). They are obtained from
1398: Eq.~(\protect\ref{theta}) with $\Delta = 10$ and from
1399: Eq.~(\protect\ref{dispersion}). In an open chain, solutions with
1400: nonzero Im~$\theta_d$ emerge starting with a certain
1401: $|g/J|>|g/J|_{\min}$. In a closed chain with even $N$ there is no
1402: threshold in $|g/J|$ for the onset of states with Im~$\theta_d\neq
1403: 0$.}
1404: \label{fig:loc_length}
1405: \end{figure}
1406:
1407: \subsection{One-excitation states in a closed chain}
1408:
1409: As pointed out in Sec.~II, for a closed chain the solution of the
1410: Schr\"odinger equation can be also sought in the form of
1411: counterpropagating waves with different amplitudes,
1412: Eq.~(\ref{right}). Clearly, the phases $\theta$ and $-\theta$ describe
1413: one and the same wave function. The one-excitation energy $E_1$ is
1414: given by Eq.~(\ref{energy_one}).
1415:
1416: The interrelation between the amplitudes of the waves $C, C'$ in
1417: Eq.~(\ref{right}) and the amplitude of the wave function on the defect
1418: site $a(n_0)$ can be obtained from Eq.~(\ref{eq_a}) with $n=n_0\pm
1419: 1$. This equation has two solutions,
1420: %
1421: \begin{equation}
1422: \label{untouched}
1423: e^{i\theta N}=1,\quad a(n_0)=Ce^{i\theta n_0}+C'e^{-i\theta n_0}
1424: \end{equation}
1425: %
1426: and
1427: %
1428: \begin{eqnarray}
1429: \label{touched}
1430: a(n_0)&=&Ce^{i\theta n_0}\left(1+e^{i\theta N}\right)\\
1431: &=&C'e^{-i\theta n_0}\left(1+e^{-i\theta N}\right)\quad [\exp(i\theta
1432: N)\neq 1].\nonumber
1433: \end{eqnarray}
1434: %
1435: In order to fully determine the wave function $a(n)$ (\ref{right}),
1436: Eqs.~(\ref{untouched}), (\ref{touched}) should be substituted into the
1437: Schr\"odinger equation (\ref{eq_a}) for $n=n_0$.
1438:
1439: Equations (\ref{eq_a}) and (\ref{untouched}) can be satisfied provided that
1440: either $g=0$, which means that there is no defect, or $a(n_0)=0$. The
1441: first condition describes excitations in an ideal closed chain and is
1442: not interesting for the present paper. The condition $a(n_0)=0$
1443: corresponds to the wave function $a(n)\propto \sin \theta (n-n_0)$,
1444: which has a simple physical meaning. It is a standing wave in an ideal
1445: chain with a node at the location of a defect. Because of the node,
1446: the corresponding state ``does not know'' about the defect, and
1447: therefore it is exactly the same as in an ideal chain.
1448:
1449: In the presence of a defect, the solutions of Eq.~(\ref{untouched}) in
1450: the range of interest $0<\theta < \pi$ are $\theta = 2\pi k/N$ with
1451: $k=1,2,\ldots,(N-1)/2$ for odd $N$, or $k=1,2,\ldots,N/2-1$ for even
1452: $N$.
1453:
1454: The equation for $\theta$ that follows from Eqs.~(\ref{eq_a}) and
1455: (\ref{touched}) has the form
1456: %
1457: \begin{equation}
1458: \label{dispersion}
1459: \exp (i\theta N) - 1 = -\frac{ig}{J \sin \theta} [\exp (i\theta N)+1] .
1460: \end{equation}
1461: %
1462: For $g\neq 0$ this equation has either $(N+1)/2$ (for odd $N$) or
1463: $N/2+1$ (for even $N$) solutions for $\pm \theta$. Therefore the total
1464: number of solutions for $\theta$ that follow from
1465: Eqs.~(\ref{untouched}) and (\ref{dispersion}) is $N$, as expected.
1466:
1467: By rewriting Eq.~(\ref{dispersion}) as $\tan(\theta N/2) =
1468: -(g/J\sin\theta)$ and plotting the left- and right-hand sides as
1469: functions of $\theta$ (cf. Ref.~\onlinecite{Economou}), one can see
1470: that all physically distinct roots of this equation but one are real
1471: and lie in the interval $0< \theta < \pi$ [except for one case, see
1472: below]. Such solutions describe delocalized states with
1473: sinusoidal wave functions.
1474:
1475: The complex root of Eq.~(\ref{dispersion}), $\theta=\theta_d$,
1476: describes a state localized on the defect. For a long chain,
1477: Im~$\theta_d N\gg 1$, the solution has the form (\ref{one_localized}),
1478: as expected. An interesting situation occurs for a shorter chain. If
1479: $N$ is even or if $g/J>0$, a localized solution with complex
1480: $\theta_d$ emerges for any defect excess energy $g$. Thresholdless
1481: localization does not happen in an open chain. In a closed chain, it
1482: arises because there is no reflection from boundaries. For small
1483: positive $g/J$ one obtains the complex solution of
1484: Eq.~(\ref{dispersion}) in the form $\theta_d\approx i(2g/NJ)^{1/2}$.
1485: The square-root dependence of Im~$\theta_d$ on $g$ is seen from
1486: Fig.~\ref{fig:loc_length}.
1487:
1488: Equation (\ref{dispersion}) has a complex
1489: solution also for even $N$ and small negative $g/J$. In this case
1490: $\theta_d\approx i(-2g/NJ)^{1/2} + \pi$, i.e., the decay of the wave
1491: function $a(n)$ is accompanied by sign flips, $a(n+1)/a(n) <
1492: 0$. Such oscillations cannot be reconciled with the
1493: periodicity condition for odd $N$. Therefore, for odd $N$ and negative
1494: $g/J$ a decaying solution arises only when $-g/J$ exceeds a threshold
1495: value. One can show from Eq.~(\ref{dispersion}) that this value is
1496: $|g/J|_{\min}=2/N$, which has also been confirmed numerically.
1497:
1498:
1499:
1500:
1501:
1502:
1503:
1504:
1505:
1506: \begin{thebibliography}{99}
1507:
1508:
1509: \bibitem{recent_reviews} M.A. Nielsen and I.L. Chuang,
1510: {\it Quantum Computation and Quantum Information} (Cambridge
1511: University Press, Cambridge, 2000).
1512:
1513: %\bibitem{Kane98} B. Kane, Nature {\bf 393}, 133 (1998).
1514:
1515: \bibitem{Makhlin99} Y. Makhlin, G. Sch\"on, and A. Shnirman,
1516: Rev. Mod. Phys. {\bf 73}, 357 (2001).
1517:
1518: \bibitem{JJ-all}
1519: %V.D. Averin, Solid State Commun. {\bf 105}, 659 (1998);
1520: %Y. Nakamura, Y.A. Pashkin, and H.S. Tsai, Nature {\bf 398}, 786 (1999);
1521: D. Vion, A. Aassime, A. Cottet, P. Joyez, H. Pothier,
1522: C. Urbina, D. Esteve, and M.H. Devoret, Science {\bf 296}, 886 (2002);
1523: Y. Yu, S.Y. Han, X. Chu, S.I. Chu, and Z. Wang, {\it ibid} {\bf 296}, 889
1524: (2002);
1525: Y.A.~Pashkin, T.~Yamamoto, O.~Astafiev, Y.~Nakamura, D.V.~Averin, and
1526: J.S.~Tsai, Nature (London) {\bf 421}, 823 (2002);
1527: A.J. Berkley, H. Xu, R.C. Ramos, M.A. Gubrud, F.W. Trauch,
1528: P.R. Johnson, J.R. Anderson, A.J. Dragt, C.J. Lobb, and
1529: F.C. Wellstood, Science {\bf 300}, 1548 (2003).
1530: %D.A. Lidar, L.A. Wu, and A. Blais,
1531: %Quant. Info. Proc. {\bf 1}, 155 (2002).
1532:
1533:
1534: \bibitem{mark} P.M. Platzman and M.I. Dykman, Science {\bf 284},
1535: 1967 (1999); M.I. Dykman and P.M. Platzman, Fortschr. Phys.
1536: {\bf 48}, 9 (2000);
1537: M.I. Dykman, P.M. Platzman, and P. Seddighrad, Phys. Rev. B {\bf 67},
1538: 155402 (2003).
1539:
1540: \bibitem{Chuang9802} I.L. Chuang, L.M.K. Vandersypen,
1541: X. Zhou, D.B. Leung, and S. Lloyd,
1542: Nature (London) {\bf 393}, 143 (1998);
1543: L.M.K. Vandersypen, M. Steffen, G. Breyta, C.S. Yannoni,
1544: M.H. Sherwood, and I.L. Chuang, {\it ibid} {\bf 414}, 883 (2001).
1545:
1546: \bibitem{Yamamoto02} T.D. Ladd, J.R. Goldman, F. Yamaguchi,
1547: Y. Yamamoto, E. Abe, and K.M. Itoh,
1548: Phys. Rev. Lett. {\bf 89}, 017901 (2002).
1549:
1550: \bibitem{Cory00} D.G. Cory, R. Laflamme, E. Knill, L. Viola, T.F. Havel,
1551: N. Bouland, G. Boutis, E. Fortunato, S. Lloyd, R. Martinez,
1552: C. Negrevergne, M. Pravia, Y. Sharf, G. Teklemariam,
1553: Y.S. Weinstein, and W.H. Zurek,
1554: Fortschr. Phys. {\bf 48}, 875 (2000).
1555:
1556: \bibitem{Piermarocchi02} P.~Chen, C.~Piermarocchi, and L.J.~Sham,
1557: Phys. Rev. Lett. {\bf 87}, 067401 (2001); X.~Li, Y.~Wu, D.G.~Steel,
1558: D.~Gammon, T.H.~Stievater, D.S.~Katzer, D.~Park, L.J.~Sham, and
1559: C.~Piermarocchi, Science {\bf 301}, 809 (2003).
1560:
1561: \bibitem{Demille02} D. DeMille, Phys. Rev. Lett. {\bf 88}, 067901 (2002).
1562:
1563: \bibitem{Silvestrov01} P.G. Silvestrov, H. Schomerus, and
1564: C.W.J. Beenakker, Phys. Rev. Lett. {\bf 86}, 5192 (2001);
1565:
1566: \bibitem{Kaminsky-Lloyd02} W.M. Kaminsky and S. Lloyd quant-ph/0211152.
1567:
1568: \bibitem{Bethe} H.A. Bethe, Z. Phys. {\bf 71}, 205 (1931);
1569: C.N. Yang and C.P. Yang, Phys. Rev. {\bf 150}, 321, 327 (1966).
1570:
1571: \bibitem{Baxter} F.C. Alcaraz, M.N. Barber, M.T. Batchelor, R.J.
1572: Baxter and G.R.W. Quispel, J. Phys. A. {\bf 20}, 6397 (1987).
1573:
1574: \bibitem{integrable} P. Schmitteckert, P. Schwab, and U. Eckern,
1575: Europhys. Lett. {\bf 30}, 543 (1995);
1576: H.-P. Eckle, A. Punnoose, and R.A. R\"omer,
1577: Europhys. Lett. {\bf 39}, 293 (1997);
1578: A. Zvyagin, J. Phys. A {\bf 34}, R21 (2001).
1579:
1580: %\bibitem{Golding_unpublished} B. Golding, unpublished.
1581:
1582: %\bibitem{Romer99}
1583:
1584:
1585: %\bibitem{Wellstood} P.R. Johnson, F.W. Strauch, A.J. Dragt,
1586: %R.C. Ramos, C. J. Lobb, J. R. Anderson, and F. C. Wellstood,
1587: %Phys. Rev. B {\bf 67}, 020509 (2003); A.J. Berkley, H. Xu, R.C. Ramos,
1588: %M.A. Gubrud, F.W. Trauch, P.R. Johnson, J.R. Anderson, A.J. Dragt,
1589: %C.J. Lobb, and F.C. Wellstood, Science {\bf 300}, 1548 (2003).
1590:
1591:
1592: \bibitem{Shepelyansky00} B. Georgeot and D.L. Shepelyansky,
1593: Phys. Rev. E {\bf 62}, 3504, 6366 (2000).
1594:
1595: \bibitem{Izrailev01} G.P. Berman, F. Borgonovi, F.M. Izrailev,
1596: and V.I. Tsifrinovich, Phys. Rev. E {\bf 64}, 056226 (2001);
1597: {\it ibid} {\bf 65}, 015204 (2002).
1598:
1599: \bibitem{Lloyd02} V.V. Flambaum, Aust. J. Phys. {\bf 53}, 489 (2000);
1600: %P.G. Silvestrov, H. Schomerus, and C.W.J. Beenakker, Phys. Rev. Lett.
1601: %{\bf 86}, 5192 (2001);
1602: J. Emerson, Y.S. Weinstein, S. Lloyd, and D.G. Cory,
1603: Phys. Rev. Lett. {\bf 89}, 284102 (2002).
1604:
1605: \bibitem{Economou} G.F. Koster and J.C. Slater,
1606: Phys. Rev. {\bf 95}, 1167 (1954); E.N. Economou,
1607: {\it Green's function in quantum physics}, (Springer-Verlag,
1608: Berlin, New York, 1979).
1609:
1610: \bibitem{Goodkind01} J.M. Goodkind and S. Pilla, Quantum
1611: Information and Computation {\bf 1}, 108 (2001).
1612:
1613: \bibitem{us_DS} M.I. Dykman and L.F. Santos, J. Phys. A
1614: {\bf 36}, L561 (2003).
1615:
1616:
1617: \end{thebibliography}
1618:
1619: \end{document}
1620: