1: \documentclass[aps,twocolumn,prl]{revtex4}
2: %\documentclass[aps,epsfig,float,twocolumn,prl]{revtex4}
3: %\documentstyle[aps,epsfig,float,twocolumn,prl]{revtex}
4: %\documentstyle[11pt]{article}
5: \newcommand{\be}{\begin{equation}}
6: \newcommand{\ee}{\end{equation}}
7: \newcommand{\sg}{\rm{sign}}
8:
9: \usepackage{graphicx}
10: \usepackage{amsmath}
11: \usepackage{amstext}
12: \usepackage{amssymb}
13: \usepackage{amsfonts}
14: \usepackage{amsbsy}
15: %\usepackage{showkeys}
16:
17:
18:
19:
20: \renewcommand{\v}[1]{\boldsymbol{#1}}
21: \newcommand{\si}{\sigma}
22: \newcommand{\<}{\langle}
23: \renewcommand{\>}{\rangle}
24: \renewcommand{\t}[1]{\tilde{#1}}
25: \renewcommand{\th}{\theta}
26: \newcommand{\cH}{{\cal H}}
27:
28:
29:
30:
31:
32: \begin{document}
33: \bibliographystyle{apsrev}
34:
35: %\twocolumn[\hsize\textwidth\columnwidth\hsize\csname @twocolumnfalse\endcsname
36:
37: \title{Quasi-adiabatic Continuation of Quantum States:
38: %Area, Perimeter, and Zero Laws
39: The Stability of Topological Ground State Degeneracy
40: and Emergent Gauge Invariance
41: }
42: \author{M. B. Hastings$^1$ }
43: \email{hastings@lanl.gov}
44: \author{Xiao-Gang Wen$^2$}
45: \homepage{http://dao.mit.edu/~wen}
46: %} \address{
47: \affiliation{$^1$Center for Nonlinear Studies and Theoretical Division,
48: Los Alamos National
49: Laboratory, Los Alamos, NM 87545}
50: \affiliation{
51: $^2$Department of Physics, Massachusetts Institute of Technology,
52: Cambridge, Massachusetts 02139
53: }
54:
55: \date{March 7, 2005}
56: \begin{abstract}
57: We define for quantum many-body systems a quasi-adiabatic continuation of
58: quantum states. The continuation is valid
59: when the Hamiltonian has a gap, or else has a sufficiently small
60: low-energy density of states, and thus is away from a quantum phase
61: transition. This continuation takes local operators into local operators,
62: while approximately
63: preserving the ground state expectation values. We apply this
64: continuation to the problem of gauge theories coupled to matter, and propose a
65: new distinction, perimeter law versus ``zero law" to identify confinement.
66: We also apply the continuation to local bosonic models with emergent gauge
67: theories. We show that local gauge invariance is topological and cannot
68: be broken by any local perturbations in the bosonic models in either
69: continuous or discrete gauge groups. We show that
70: the ground state degeneracy in emergent discrete gauge theories
71: is a robust property of the bosonic model, and we argue that the
72: robustness of local gauge invariance in the continuous case protects
73: the gapless gauge boson.
74: \vskip2mm
75: \end{abstract}
76: \maketitle
77: %]
78:
79: \section{Introduction}
80:
81: Traditionally,
82: gauge theory was described as a theory of a vector field $a_\mu$ that has
83: a local gauge symmetry. In a Lagrangian framework,
84: this means $L(a_\mu+\partial_\mu
85: \phi)=L(a_\mu)$. This gauge symmetry was believed to be
86: the defining property
87: of a gauge theory. It was believed that the gauge symmetry protects the gapless
88: gauge boson for continuous gauge groups and the topological ground state
89: degeneracy on
90: compact space for discrete gauge groups. Even the slightest gauge symmetry
91: breaking, such as a gauge potential term $(a_\mu)^2$ in the Lagrangian, gives a
92: finite mass to the gauge boson or lifts the topological degeneracy. In this
93: case we can no longer regard the theory as a gauge theory at low energies.
94:
95: However, the above standard picture for gauge symmetry protecting gapless
96: gauge bosons and ground state degeneracy is very formal, since the gauge
97: symmetry is not really a symmetry.
98: %is very different from the continuous global symmetry that protects the
99: %gaplessness of Goldstone bosons.
100: Within the Hamiltonian formulation of gauge theories, the gauge transformation
101: is simply a transformation between different labels that label the \emph{same}
102: physical state. It is a do-nothing transformation. It is very different from
103: the usual symmetry transformation that transforms a physical state to a
104: \emph{different} physical state.
105: Therefore, it is not clear what is the essence of
106: gauge theory and gauge symmetry.
107: %So the standard picture for gauge theory is too formal that may hide the
108: %essence of the gauge theory and gauge symmetry. It may hide the physical
109: %reason why gauge bosons are gapless.
110:
111: In last 15 years, it was shown, with increasing rigor, that \emph{deconfined}
112: gauge theories can emerge from certain local bosonic lattice models.
113: \cite{KL8795,WWZcsp,RS9173,Wsrvb,SF0050,MS0181,Wqoslpub,Wlight,SP0258,BFG0212,%
114: MS0204,Walight,MS0312,HFB0404,AFF0493,LWstrnet} If the emergent gauge theory
115: has a continuous gauge group, the local bosonic model will have gapless
116: excitations that behave just like gauge
117: bosons.\cite{Wlight,MS0204,Walight,MS0312,HFB0404} If the emergent gauge
118: theory has a discrete gauge group or has a Chern-Simons term, the local
119: bosonic model will have degenerate ground states on compact
120: space.\cite{Wrig,WNtop,Wsrvb} This raises a physical question: what protects
121: the gapless gauge bosons and the ground state degeneracy? According to the
122: standard picture for gauge theory, those properties are protected by gauge
123: symmetry, but from the point of view of local bosonic model, what is this
124: ``gauge symmetry''? How does ``gauge symmetry'' emerge at low energies?
125:
126: A close examination of those local bosonic models with emergent gauge theory
127: reveals that the emergence of gauge theory is intimately connected to
128: string condensation.\cite{Walight,LWstrnet} The ``gauge symmetry'' is
129: related to the integrity of the strings. If the strings are unbreakable, one
130: can show that the low energy states are gauge invariant. However, in general,
131: the strings in the boson model are not perfectly well defined. Strings may
132: break up momentarily and rejoin. One may wonder if this means that the
133: ``gauge symmetry'' becomes approximate. We know that a theory that loses its
134: ``gauge symmetry'' even slightly no longer behaves like a gauge theory at low
135: energies. This seems to suggest that breakable strings will give gauge bosons a
136: mass gap or lift the ground state degeneracy.
137:
138: On the other hand, it was believed that the gaplessness of the gauge bosons
139: and the degeneracy of the ground states in those bosonic models are
140: topological and are robust against any local
141: perturbations.\cite{Wrig,WNtop,Wsrvb}
142: % of the bosonic model.
143: A formal argument goes as the follows (see, for example, Ref. \cite{Wen04},
144: page 393 and page 435). We first derive the low energy effective gauge theory
145: of the bosonic model. We then argue that any generic perturbation of the
146: bosonic model cannot generate any terms that break the ``gauge symmetry'' in
147: the low energy effective gauge theory. Therefore, all the properties protected
148: by the ``gauge symmetry'' are robust against arbitrary perturbations of the
149: bosonic model.
150:
151: We see that to understand why ``gauge symmetry'' remains exact
152: even for a generic boson model with virtually breakable strings is vital in
153: our understanding why the degeneracy of the ground states is protected even
154: when the original boson model has no symmetry, and why the gaplessness of the
155: gauge bosons is protected even when the original boson model has only
156: translation symmetry. We would like to address some of these issues in this
157: paper.
158:
159: In the next section, we start by introducing some bosonic models with emergent
160: gauge theories, and raise the question of the ability of topological order and
161: gauge symmetry to survive local perturbations of the Hamiltonian, as well as
162: proposing the zero law to identify deconfined gauge theories. We then
163: introduce in the following section a quasi-adiabatic continuation which
164: enables us to identify appropriately dressed operators for the perturbed
165: Hamiltonian, such that the dressed operator has almost the same ground state
166: expectation value for the perturbed Hamiltonian as the original operator had
167: for the original Hamiltonian. The final section illustrates the application
168: of the continuation
169: via a series of examples, beginning with local models without
170: emergent gauge structure, such as quantum Ising models, and then going on to
171: emergent gauge theories. Many of these systems are theories for which the
172: unperturbed Hamiltonian has a ground state degeneracy and then a gap to the
173: rest of the spectrum, with no local operators connecting the degenerate ground
174: states. We are then able to show that, so long as the gap to the rest of the
175: spectrum remains open, the splitting between the low energy states of the
176: perturbed Hamiltonian remains exponentially small. This is illustrated in the
177: case of the ferromagnetic quantum Ising model; in the case of the fractional
178: quantum Hall effect where we are able to extend results on the insensitivity
179: of the topological degeneracy to disorder\cite{Wrig,WNtop}; and in the case of
180: an emergent $Z_2$ gauge theory. The case of gapless theories is also
181: discussed; it will turn out that gauge symmetry is much more robust
182: in compact than in non-compact theories.
183:
184: \section{Simple local bosonic models with emergent gauge theories}
185:
186: To make the above discussion more concrete, in this section we are going to
187: discuss two simple bosonic models that have emergent $Z_2$ and $U(1)$
188: gauge theories respectively.
189:
190: \subsection{A bosonic model with emergent $Z_2$ gauge theory}
191:
192: The first bosonic model is a spin-1/2 model on a $d$-dimensional cubic lattice.
193: \cite{W7159,K7959,K032}
194: The spins live on the links labeled by $\v i$.
195: The Hamiltonian is given by
196: \begin{eqnarray}
197: \label{HZ2}
198: H_0=U\sum_{\v I}\left( 1-W_{\v I}\right)
199: -g\sum_{\v p} \left(\prod_\text{edges of $\v p$} \si^x_{\v i}\right)
200: \end{eqnarray}
201: where $\v I$ labels the vertices and $\v p$ the squares of the lattice, and
202: $W_{\v I}$ is given by
203: \begin{equation}
204: \label{xf}
205: W_{\v I}=\prod_\text{legs of $\v I$} \si^z_{\v i}.
206: \end{equation}
207: The legs of a vertex are the links that connect to the vertex and the edges
208: of a square are the four links around the square.
209: $\si^{x,y,z}$ are the Pauli matrices.
210:
211: \begin{figure}[tb]
212: \centerline{
213: \includegraphics[scale=0.4]{z2str.eps}
214: }
215: \caption{
216: A closed-string state. The up-spins are represented by open dots
217: and the down-spin by filled dots.
218: }
219: \label{z2str}
220: \end{figure}
221:
222: \begin{figure}[tb]
223: \centerline{
224: \includegraphics[scale=0.7]{specH0.eps}
225: }
226: \caption{
227: The energy levels of $N$ spin-1/2 spins described by $H_0$.
228: }
229: \label{specH0}
230: \end{figure}
231:
232: When $U$ is very large, the low energy sector of the model is formed by closed
233: string states. What are the closed string states?
234: First, the state with all spins up is defined as the no string state.
235: A closed string state is a state where the down spins form closed loops
236: (see Fig. \ref{z2str}).
237: We note that the close-string states are the states that satisfy
238: $ W_{\v I} |\text{closed strings}\> = |\text{closed strings}\>$,
239: Since $[W_{\v I}, H_0]=0$, the closed-string states and the open-string
240: states\footnote{The open-string states are states which are not closed-string
241: states.} do not mix.
242: This allows us to plot the
243: spectrum of $H_0$ separately for closed-string states and open-string states
244: in Fig. \ref{specH0}.
245:
246: When
247: $g>0$, the ground state $|\Phi_0\>$ of the model is the equal weight
248: superposition of all closed string states. Such a state is called a
249: closed-string condensed state since the closed-string creation operator
250: $S(C_\text{closed})$ has a non-zero expectation value \cite{Wqoexct}
251: \begin{eqnarray}
252: \label{zlaw}
253: \<\Phi_0|S(C_\text{closed})|\Phi_0\>=1
254: \end{eqnarray}
255: regardless of the size of the string $C_\text{closed}$.
256: Here a string creation operator that creates a string $C$ is given by
257: \begin{eqnarray*}
258: S(C)=\prod_{\v i\text{ on }C}\si^x_{\v i}.
259: \end{eqnarray*}
260: We can show (\ref{zlaw}) to be true by noting that
261: \begin{equation}
262: \label{SHcom}
263: [S(C_\text{closed}), H_0]=0.
264: \end{equation}
265:
266:
267: What is the physical character of the closed-string condensed state? It is
268: believed that the closed-string condensed state contain a new kind of order
269: -- topological order -- which cannot be characterized by symmetry breaking and
270: long range order.\cite{Wrig,LWstrnet} So we need a new way to characterize such
271: an order. One way to characterize the topological order is through the robust
272: ground state degeneracy on torus.\cite{WNtop,Wsrvb} The closed-string condensed
273: state in our spin-1/2 model is characterized by a four-fold ground state
274: degeneracy.
275:
276: To show the four-fold ground state degeneracy,
277: we would like to first point out that the $S(C_\text{close})$ are not the only
278: closed string operators that commute with the Hamiltonian.
279: We can define dual string operators that also commute with the
280: Hamiltonian.\cite{K032}
281:
282: \begin{figure}[tb]
283: \centerline{
284: \includegraphics[scale=0.4]{z2dstr.eps}
285: }
286: \caption{
287: A dual closed-string.
288: }
289: \label{z2dstr}
290: \end{figure}
291:
292: While strings are formed by segments that connect nearest neighbor vertices,
293: dual strings are formed by segments that connect the centers of
294: nearest neighbor squares (see Fig. \ref{z2dstr}).
295: A dual string operator $\t S(\t C)$ for a dual
296: string $\t C$ is defined as
297: \begin{equation}
298: \label{ds}
299: \t S(\t C)=\prod_{\v i\text{ cross }\t C}\si^z_{\v i}.
300: \end{equation}
301: One can check that
302: \begin{equation}
303: \label{SHcomD}
304: [\t S(\t C_\text{closed}), H_0]=0,
305: \end{equation}
306: which implies that the dual closed-strings also condense
307: \begin{eqnarray}
308: \label{zlawD}
309: \<\Phi_0|\t S(\t C_\text{closed})|\Phi_0\>=1.
310: \end{eqnarray}
311:
312: Now we are ready to show that the ground states of $H_0$ on torus have at
313: least four-fold degeneracy. Let $C_x$ and $C_y$ ($\t C_x$ and $\t C_y$)
314: be the closed (dual) strings that wrap around the torus once in $x$ and $y$
315: directions. We find that the four large-string operators
316: $(S(C_x), S(C_y),\t S(\t C_x), \t S(\t C_y))$ commute with each other
317: except
318: \begin{eqnarray}
319: \label{acr}
320: \{S(C_x), \t S(\t C_y)\}=0, \ \ \ \ \ \
321: \{\t S(\t C_x), S(C_y)\}=0.
322: \end{eqnarray}
323: The above algebra has only one four-dimensional irreducible representation.
324: Since the large-closed-string operators $(S(C_x), S(C_y),\t S(\t C_x), \t S(\t
325: C_y))$ act within the degenerate ground states, the ground state degeneracy
326: must be a multiple of four.
327:
328: When $U= \infty$, the model becomes the $Z_2$ gauge theory.\cite{W7159}
329: One way to see
330: this is to note that the only states with finite energies are
331: closed-string states and closed-string states are gauge invariant states under
332: local $Z_2$ gauge transformations. The local $Z_2$ gauge transformations are
333: generated by the unitary operators $W_{\v I}$ in (\ref{xf}):
334: \begin{eqnarray}
335: \label{gaugeinv}
336: W_{\v I} |\text{closed strings}\> = |\text{closed strings}\> \ \ \
337: \text{for any }\v I.
338: \end{eqnarray}
339: A generic $Z_2$ gauge transformation is given by
340: \begin{eqnarray}
341: \prod_{\v I} (W_{\v I})^{n_{\v I}} |\text{closed strings}\>
342: = |\text{closed strings}\> .
343: \end{eqnarray}
344: %%where $n_{\v I}=0,1$ and a local $Z_2$ transformation $W_{\v I}$
345: %is given by (\ref{xf}).
346: In the gauge theory language the closed-string operator
347: $S(C_\text{closed})$ turns out to be the Wilson-loop operator,\cite{W7159,W7445}
348: which is gauge invariant
349: \begin{equation*}
350: W_{\v I} S(C_\text{closed}) W_{\v I}^\dag = S(C_\text{closed}) .
351: \end{equation*}
352:
353: (\ref{zlaw}) implies that the expectation values
354: of the Wilson loop satisfies the perimeter law
355: \begin{eqnarray}
356: \label{plaw}
357: \<\Phi_0|S(C_\text{closed})|\Phi_0\>\sim e^{-\alpha |C_\text{closed}|}
358: \end{eqnarray}
359: with zero coefficient $\alpha=0$. Here $|C_\text{closed}|$ is the length
360: of the string $C_\text{close}$.
361: In this case we will call (\ref{zlaw}) the zero law.
362:
363: The perimeter law indicates that the $Z_2$ gauge theory is in the deconfined
364: phase. The $Z_2$ deconfined phase has four nearly degenerate ground states on
365: the torus which is consistent with our previous direct calculation on the
366: spin model. The energy separation between the four nearly degenerate ground
367: states is of order $e^{-L/\xi}$ where $L$ is the linear size of the torus and
368: $\xi$ a finite length scale.
369:
370: Now let us add a term
371: \begin{equation*}
372: H_1=-J_1\sum_{\v i} \si^z_{\v i}
373: \end{equation*}
374: to our spin model Hamiltonian $H_0$ and assume $U$ is finite.
375: $S(C_x)$ and $S(C_y)$ no longer commute
376: with the modified Hamiltonian $H_0+H_1$. So it is not clear if $H_0+H_1$ still
377: has four degenerate ground states on a torus.
378:
379: %Certainly, if $J_1\gg g$, the
380: %ground state is given by a state where almost all spins pointing up, which is
381: %non degenerate. But when $J_1\ll g$, do we still have the four-fold degeneracy
382: %in the ground states?
383:
384: To understand the properties of modified spin system $H_0+H_1$, we note that
385: $H_0+H_1$ still does not mix the closed-string and open-string states. So if
386: $U\gg g,J_1$, the low lying states are still closed string states. One can
387: check that if we restrict $H_0+H_1$ to the close-string subspace, the system
388: is identical to a pure lattice $Z_2$ gauge theory.
389: %On the other hand, the low energy sector of $H_0+H_1$ is still formed by
390: %closed strings if $U$ is very large. Those low energy closed-string states
391: %are still invariant under the $Z_2$ gauge transformation. This leads us to
392: %think that low energy effective theory is still a pure $Z_2$ gauge theory.
393: A pure $Z_2$ gauge theory has two phases in d+1 dimensions if $d > 1$: a
394: deconfined phase where the expectation value of the Wilson loop satisfies the
395: perimeter law, and a confined phase where the expectation value of the Wilson
396: loop satisfies the area law: \begin{eqnarray} \label{alaw}
397: \<\Phi_0|S(C_\text{closed})|\Phi_0\>\sim e^{-\gamma A(C_\text{closed})},
398: \end{eqnarray} where $A(C_\text{closed})$ is the area enclosed by the loop
399: $C_\text{closed}$ and $\gamma >0$. So we expect our spin model $H_0+H_1$ also
400: has two phases. If $|J_1|\ll |g|$, the ground state of the spin model is
401: filled with large closed strings which correspond to the $Z_2$ deconfined
402: phase. If $J_1 \gg |g|$, the ground state of the spin model hardly has any
403: strings (i.e. almost all spins point up) which correspond to deconfined phase.
404:
405: Base on this picture, we guess that when $|J_1|\ll |g|$, the ground states of
406: $H_0+H_1$ are (nearly) four-fold degenerate, while when $|J_1|\gg |g|$ the
407: ground state is not degenerate. But the result for $|J_1|\ll |g|$ clearly is
408: just a guess. Can we provide a more rigorous proof?
409:
410: \begin{figure}[tb]
411: \centerline{
412: \includegraphics[scale=0.7]{specH012.eps}
413: }
414: \caption{
415: The energy levels of $N$ spin-1/2 spins described by $H_0+H_1+H_2$, assuming
416: $|J_1|,|J_2| \ll |g| \ll U \ll N|g|$.
417: }
418: \label{specH012}
419: \end{figure}
420:
421: The situation gets even more complicated if we add another term
422: \begin{equation*}
423: H_2=-J_2\sum_{\v i} \si^x_{\v i}.
424: \end{equation*}
425: Such a term can mix the closed-string and open-string states. The low energy
426: sector of $H_0+H_1+H_2$ is not formed by simple closed-string states, although
427: the mixing with the open string states may be small for small $J_2$ (see Fig.
428: \ref{specH012}). It appears that $H_2$ breaks the $Z_2$ ``gauge symmetry''
429: since the low energy states no longer satisfy (\ref{gaugeinv}).
430:
431:
432: Also when $J_2 \neq 0$ and $U<\infty$, the expectation
433: value of a closed string operators satisfies the perimeter law in both
434: $|J_1|\ll |g|$ limit and $J_1\gg |g|$ limit, which give no sign of two phases.
435: All of those suggest that the four-fold ground state degeneracy is lifted
436: by a finite $J_2$.
437:
438: \begin{figure}[tb]
439: \centerline{
440: \includegraphics[scale=0.7]{z2phs.eps}
441: }
442: \caption{
443: A likely quantum phase diagram for the spin-1/2 system $H_0+H_1+H_2$.
444: The deconfined phase is characterized by
445: four nearly degenerate ground states on torus
446: with energy splitting of order $e^{-L/\xi}$ where $L$ is the linear size of the
447: torus and $\xi$ a length scale.
448: The confined phase is characterized by a non degenerate ground state on torus.
449: In general, the confined phase and the deconfined phase are distinguished by
450: the zero law and the perimeter law of certain loop operators.
451: }
452: \label{z2phs}
453: \end{figure}
454:
455: But this suggestion is incorrect. Ref. \cite{Wsrvb} argues that any local
456: perturbations of $H_0$ cannot break the $Z_2$ ``gauge symmetry''. As a
457: result, $H_0+H_1+H_2$ will have four nearly degenerate ground states as long
458: as $J_1$ and $J_2$ are not too large. Such a phase contains a non-trivial
459: topological order. However, Ref. \cite{Wsrvb} only provides a formal
460: argument. A more rigorous understanding is needed.
461:
462: Certainly, large $J_1$ and $J_2$ will polarize the spins and lift the ground
463: state degeneracy. Such a phase has a trivial topological order. This
464: suggests a phase diagram in Fig. \ref{z2phs}. We note that the two phases in
465: Fig. \ref{z2phs} have the same symmetry and are distinguished only by
466: topological orders.
467:
468: {}From the phase diagram, we see that when $J_2\neq 0$ the perimeter/area laws
469: of the closed string operators (or Wilson loop) cannot determine if the ground
470: state is topologically ordered or not. Thus for a generic bosonic model,
471: perimeter/area laws of the closed string operators is not the proper way to
472: test if the ground state has closed-string condensation (or non-trivial
473: topological order).
474:
475: In the examples
476: section, we will show that the topological phase is characterized
477: by dressed closed-string operators which satisfy the zero law. The trivial
478: phase does not contain any such closed string operator. The closed string
479: operators in the trivial phase all satisfy the perimeter law. So it is the
480: zero/perimeter laws that distinguish topological/trivial (deconfined/confined)
481: phases, instead of perimeter/area laws.
482:
483: For $J_1=J_2=0$, the system has an exact four-fold degeneracy of the ground
484: state on the torus, followed by a gap of order $g$ to the next lowest state.
485: The closed-string operators satisfy the zero law. In the examples section, we
486: will show that, for small but non-zero $J_1$ and $J_2$, a deformed or dressed
487: closed-string operator can still satisfy the zero law. The zero law of the
488: dressed string operator allows us to show the four-fold degeneracy of the
489: ground states in the small $J_{1,2}$ limit. More precisely, we assume that
490: the gap, from the four lowest states to the rest of the spectrum, remains
491: open, and then we show the four-fold ground state degeneracy up to an
492: exponentially small splitting. We will also show that, in the small $J_{1,2}$
493: limit, the low energy sector of the model is still formed by $Z_2$ gauge
494: invariant states. However, for $J_2\neq 0$, the gauge invariance is under a
495: deformed $Z_2$ gauge transformation. So in this sense, none of the small
496: perturbations in the spin model can break the ``gauge symmetry'' in the low
497: energy effective gauge theory.
498:
499: \subsection{A bosonic model with emergent $U(1)$ gauge theory}
500:
501: In our second bosonic model, we consider rotors on the links of
502: $d$-dimensional cubic lattice.
503: A rotor can be viewed as a particle moving on a circle.
504: The position of the particle is given by an angle $\th$,
505: and the angular momentum of the particle by $L^z=-i \partial_\th$.
506: The Hamiltonian of the rotor model is given by
507: %(see Fig. \ref{})
508: % \cite{Walight,Wen04,HFB0404}
509: \begin{align}
510: \label{strnetH}
511: H_\text{rotor} &= U\sum_{\v I} Q^2_{\v I}
512: - g \sum_{\v p} (B_{\v p}+h.c.)
513: \nonumber\\
514: &\ \ \ \ + J_1 \sum_{\v i} (L^z_{\v i})^2
515: + J_2 \sum_{\v i} (L^+_{\v i}+L^-_{\v i})
516: \\
517: B_{\v p} &= L^+_{1}L^-_{2}L^+_{3}L^-_{4},\ \ \ \ \ \ \ \ \
518: Q_{\v I} = (-)^{\v I} \sum_{\text{legs of }\v I} L^z_{\v i} ,
519: \nonumber
520: \end{align}
521: where $\v I$
522: labels the vertices, $\v i$ labels the links and $\v p$
523: labels the squares of the cubic lattice.
524: %The legs of $\v I$ are the six links that are attached to the vertex $\v I$
525: %(see Fig. \ref{cubbos}b), and
526: 1, 2, 3, 4 label the four links that form
527: the edges of the square $\v p$.
528: $L^+=e^{i\th}$ is the raising operator of $L^z$, $L^-=(L^+)^\dag$, and
529: $(-)^{\v I}=1$ for the even vertices and.
530: $(-)^{\v I}=-1$ for the odd vertices.
531:
532: %In this paper, we mainly consider the case where $S$ is a large integer
533: %(or more precisely $S\to \infty$ limit). However, due to the stability of the
534: %string-net condensed state in 3D, we expect many results about the condensed
535: %state remain to be valid for relatively small $S$ \cite{MS0312} or even
536: %half-integer $S$ \cite{HFB0404}.
537:
538:
539: \begin{figure}[tb]
540: \centerline{
541: \includegraphics[scale=0.7]{u1str.eps}
542: }
543: \caption{
544: The empty dots represent rotors with $L^z=0$ -- the no string states.
545: A closed string is formed by a loop connecting
546: neighboring vertices.
547: A closed-string state is obtained by
548: alternately increase or decrease $L^z$
549: by 1 along the closed string.
550: The filled dots represent rotors with $L^z=\pm 1$.
551: The arrows on the links all point from even vertices to odd vertices.
552: }
553: \label{u1str}
554: \end{figure}
555:
556: When $g=J_1=J_2=0$ and $U>0$,
557: the ground states are highly degenerate and form a low
558: energy subspace. One of the ground states is the state with $L^z_{\v i}=0$
559: for every rotor. Other ground states can be constructed from the first ground
560: state by drawing an loop in the cubic lattice, and then alternately increasing
561: or decreasing $L^z$ by 1 along the loop (see Fig.~\ref{u1str}). The sum
562: $\sum_{\text{legs of }\v I} L^z_{\v i}$ vanishes on every vertices for such a
563: closed-string state. Such a process can be repeated to construct all of the
564: degenerate ground states. We see that the degenerate ground states are formed
565: by loops, or more precisely string-nets, since loops can overlap and form
566: branched strings.
567:
568: %We may define the orientation of the string on a link
569: %as pointing from the even vertex (where $I_x+I_y+I_z=$even) to odd vertex if
570: %$L^z>0$ and from the odd vertex to even vertex if $L^z<0$. Under this
571: %definition, the loops in the degenerate ground state are all oriented strings.
572:
573: The $J_1$-term gives the strings in the degenerate ground states a fine
574: energy and represents string tension. The $B_{\v p}$ operator creates a small
575: loop of closed string or deforms the existing strings.
576: %in the degenerate ground states.
577: Thus the $g$-term generates string ``hopping'' or string fluctuations.
578:
579: The $J_{1,2}$-term and the $g$-term lift the degeneracy of the ground states.
580: In the $U\gg J_{1} \gg \text{max}(|g|,|J_2|)$ limit, the true ground state correspond to a
581: state with almost no strings (i.e., a state with $L^z_{\v i}=0$ on every
582: link). The excitations above such a state have finite energy gaps. In the
583: $U\gg |g| \gg J_{1,2}>0$ limit, the true ground state is a superposition of
584: many large closed strings \cite{Walight,Wen04}. Such a state is a string-net
585: condensed state.
586:
587: When $J_2=0$, the Hamiltonian $H_\text{rotor}$ does not mix the closed-string
588: states and open-string states.
589: When restricted to the closed-string
590: subspace, $H_\text{rotor}$ is identical to the Hamiltonian of lattice $U(1)$
591: gauge theory.
592: The closed-string states can be viewed as gauge invariant states
593: since they are invariant under local $U(1)$ gauge transformations
594: \begin{equation}
595: \label{u1gauge}
596: e^{i \phi Q_{\v I}}
597: |\text{closed-string}\> = |\text{closed-string}\>,\ \ \ \
598: \text{for any } \v I.
599: \end{equation}
600: A general $U(1)$ gauge transformation is generated by
601: $e^{i\sum_{\v I} \phi_{\v I} Q_{\v I}}$.
602:
603: In the limit $|J_1| \gg |g|$, the lattice $U(1)$ gauge theory is in the strong
604: coupling limit and is in a confined phase. In the limit $|J_1| \ll |g|$, the
605: lattice $U(1)$ gauge theory is in the weak coupling limit and has gapless
606: $U(1)$ gauge bosons as its only low lying excitations if $d>2$. So when
607: $J_2=0$ and when $U\gg |g| \gg |J_1|$, the rotor model $H_\text{rotor}$
608: contains emergent gapless $U(1)$ gauge
609: bosons.\cite{MS0204,Walight,MS0312,HFB0404,Wen04}
610:
611: When $J_2\neq 0$, $H_\text{rotor}$ mixes the closed-string and open-string
612: states. The low energy states are no longer pure closed-string states and is
613: not invariant under the local $U(1)$ gauge transformation (\ref{u1gauge}). It
614: appears that a non-zero $J_2$ will break the $U(1)$ gauge symmetry. We may
615: conclude that even a small $J_2$ will give the $U(1)$ gauge boson a gap and
616: the rotor model $H_\text{rotor}$ cease to have emergent $U(1)$ gauge bosons at
617: low energies. In the section of examples,
618: we will show that this line of argument
619: is incorrect. For a small $J_2$ (or any other small perturbations to
620: $H_\text{rotor}$), we can define deformed local $U(1)$ gauge transformations
621: so that the low energy states of $H_\text{rotor}$ are invariant under the
622: deformed local $U(1)$ gauge transformations. Thus the local ``$U(1)$ gauge
623: symmetry'' cannot be broken by any small perturbations if $U\gg |J_{1,2}|,
624: |g|$. Thus far, we will prove these results for $U(1)$ theories;
625: we do not prove, but strongly conjecture, that the stability of
626: the $U(1)$ gauge symmetry protects the gaplessness of the
627: gauge boson.
628: As a result, no translation invariant perturbation can give the gapless
629: $U(1)$ gauge boson a mass gap. The gaplessness of the emergent $U(1)$ gauge
630: boson is topologically stable!
631:
632:
633:
634:
635:
636:
637:
638: \section{Quasi-Adiabatic Continuation}
639:
640: In this section we define the quasi-adiabatic continuation. We consider a
641: family of Hamiltonians, ${\cal H}_s$, depending on a continuous parameter $s$,
642: where we wish to define a continuation from $s=0$ to $s=1$. We denote
643: eigenstates of ${\cal H}_{s}$ by $\Psi_{a,s}$, where a state $\Psi_{a,0}$
644: evolves into state $\Psi_{a,s}$ under an {\it adiabatic} evolution of ${\cal
645: H}_s$. In the event of a level crossing as a function of $s$, any
646: arbitrary continuation of the states through the level crossing is allowed.
647:
648: Let us begin with some motivation and discussion. We define the unitary
649: operator $V(s)=\sum_a |\Psi_{a,s}\rangle \langle\Psi_{a,0}|$. Then, for any
650: operator $O$, we could define a corresponding dressed operator by $O_{\rm
651: adiab}(s)=V(s)OV(s)^{\dagger}$ so that $O_{\rm adiab}(s)$ would have exactly
652: the same expectation value in state $\Psi_{a,s}$ as $O$ does in state
653: $\Psi_{a,0}$. Using such a definition of dressed operators, we can show that
654: the dressed string operators and the dressed gauge transformations will have
655: the same properties in the deformed model as the bare string operators and the
656: bare gauge transformations in the exactly soluble model. However, this
657: definition of a dressed operator would not suffice for our purposes at all!
658: In particular, we do not have any reason to believe that the resulting $O_{\rm
659: adiab}(s)$ would still be a local operator.
660:
661: Indeed, even if $O$ only involves operators on a few sites, in general, the
662: continued $O_{\rm adiab}(s)$ will contain operators on every site of the
663: system. However, in this section, we will show that, under certain
664: conditions, we can continue $O$ into an local operator that acts only on a
665: finite number of sites.
666:
667: To state the result more precisely, let us first assume that $\cH_s$ has a gap
668: $\Delta E$
669: separating a low energy sector and a high energy sector for all $0<s<1$.
670: For
671: any operator $O$ which acts on a set of sites $S_O$, we will construct
672: an approximate dressed operator $O(s)$ which only acts on sites within
673: a distance $l$ from the sites $S_O$. Then, there is a unitary matrix
674: $Q_0(s)$ which acts only within the low energy sector of $\cH_s$, such that
675: \begin{equation}
676: \label{boundGap}
677: \<\Psi_\text{low,s}|Q_0(s)^{\dagger} O(s) Q_0(s)
678: - O_{\rm adiab}(s)
679: |\Psi_\text{low,s}\> < N_{S_{O,l}} e^{- l/\xi}
680: \end{equation}
681: in the large $l$ limit for a certain
682: fixed length scale $\xi$, where $N_{S_{O,l}}$ is the number of
683: sites within distance $l$ of a site in $S_O$. This number grows
684: only as a power of $l$, so is easily overwhelmed by the exponential.
685: The state $|\Psi_\text{low,s}\>$ is any state in the low energy sector
686: of $\cH_s$.
687:
688: If $\cH_s$
689: has no gap between the low energy and high energy
690: sectors, we will define a density of states $\rho(E)$ in the high
691: energy sector. If $\rho(E)$ is
692: bounded by $\rho(E)< E^{\alpha-1}$, then the
693: bound in (\ref{boundGap}) is weakened to
694: \begin{equation}
695: \label{boundNoGap}
696: \<\Psi_\text{low,s}|Q_0(s)^{\dagger} O(s) Q_0(s)-O_{\rm adiab}(s)
697: |\Psi_\text{low,s}\> <
698: N_{S_{O,l}} l^{1-\alpha/2}.
699: \end{equation}
700: For a local operator $O$, $N_{S_{O,l}}$ grows as $l^d$, where
701: $d$ is the dimension of the lattice, and so if $d+1-\alpha/2<0$ then
702: the error decays for large $l$. For a string-like operator,
703: $N_{S_{O,l}}$ is proportional to $l^{d-1} |O|$, where $|O|$ is the
704: length of the strength, for $l<|O|$, and is proportional to $l^d$ for
705: $l>|O|$.
706: (A more rigorous statement of the results will be given later.)
707:
708: The key in obtaining the above result is to
709: adopt a {\it different} definition of the dressed operator by
710: $O(s)=\t V(s)O \t V(s)^{\dagger}$, where the unitary operator $\t V(s)$ is defined
711: following Eq.~(\ref{weq}). Physically, the definition of $V$ corresponds to
712: adiabatically changing the Hamiltonian from ${\cal H}_0$ to ${\cal H}_s$,
713: while the definition (\ref{weq}) of $\t V(s)$ corresponds to a quasi-adiabatic
714: change of the Hamiltonian. So $\t V(s)$ can be viewed as an approximation of
715: $V(s)$. The operator $\t V(s)$ will be chosen to achieve the goal of defining
716: dressed operators $O(s)$ which have approximately the same ground state
717: expectation values in the perturbed Hamiltonian ${\cal H}_s$ as the original
718: operators did in the unperturbed Hamiltonian, while preserving the locality of
719: the operators.
720:
721: The locality of the operators is very important. Suppose, for example that
722: the ground state of ${\cal H}_0$ has long-range correlations. That is, there
723: exist two local operators, $O_1,O_2$ which are correlated even though the
724: two operators are far separated from each other in space. For example, if
725: this is a spin system with long-range spin correlations these
726: may be spin operators acting on two different sites which are far separated from
727: each other. Then, however, the operators $O_1(s),O_2(s)$ will also
728: be correlated in the ground state of ${\cal H}_s$, and since the operators
729: remain local under the continuation this implies the existence of long-range
730: correlations in the ground state of ${\cal H}_s$.
731:
732: The major result will be an explicit definition of $\t V(s)$ in terms of
733: derivatives of the Hamiltonian which accomplishes these goals. In order to
734: show that ground state expectation values remain approximately unchanged under
735: the continuation, we will make some assumptions on the existence of a gap,
736: though some extensions to gapless systems with sufficiently small low-energy
737: density of states will be discussed. The dressed operators remain local under
738: this evolution: local operators are spread out over a length scale of order
739: the correlation length in a gapped system, while string-like operators are
740: spread out over a length scale which grows only logarithmically with the
741: string length. Detailed proofs will be given in an Appendix.
742:
743: The importance of the low energy density of states in what follows can be
744: understood physically by analogy to another continuation that should be much
745: more familiar, namely Fermi liquid theory. As discussed by
746: Anderson\cite{pwa}, the correct way to think of Fermi liquid theory is to
747: think of starting with a non-interacting system and turning the interactions
748: on slowly, but not infinitely slowly; that is, physically exactly the same
749: procedure we imagine here. Anderson's discussion of how fast the interactions
750: need to be turned on is based on considerations of the quasi-particle states,
751: and corresponds very closely to our two criteria: maintaining both the
752: expectation values and the locality of the operators. To show that this is
753: possible, the analysis in the Fermi liquid case relies on the low density of
754: particle-hole excitations near the Fermi surface. Here, we rely on something
755: similar, namely a low density of states at low energy. Our continuation is
756: very general, and thus valid for a much wider range of systems than Fermi
757: systems, but this generality can in some cases limit what we can prove on
758: specific systems.
759:
760: \subsection{Definition of System}
761:
762: We consider a family of Hamiltonians, ${\cal H}_s$, which obey the
763: finite-range conditions\cite{LR7251,H0402}: ${\cal H}_s=\sum_i {\cal H}_s^i$,
764: where letters $i,j,...$ labels different lattice sites of the system; each
765: ${\cal H}_s^i$ acts only on sites $j$ with $d(i,j)\leq R$ where $R$ is the
766: interaction
767: range and $d(i,j)$ is some metric on the lattice; and $||{\cal H}_s^i||\leq J$
768: for some constant $J$ for all $i,s$. We further assume that
769: $||\partial_s{\cal H}_s^i|| \leq K$ for some constant $K$ for
770: all $i,s$. It is possible to slightly
771: weaken the finite-range conditions and consider exponentially decaying
772: interactions as well\cite{LR7251}.
773:
774: To define the concept of ``density of states'', let us focus on a state
775: $\Psi_{\rm low,s}$
776: that we continue. We will consider a $\Psi_{\rm low,s}$ which is a
777: ground state of $\cH_s$, but in general we could continue any eigenstate of
778: $\cH_s$. The state $\Psi_{a,s}$ has energy $E_a>0$ compared to $\Psi_{\rm
779: low,s}$ if $a$ is not in the low energy sector.
780: We assume that, for $0\leq s \leq 1$, the
781: density of states of the local operator $u^i_s=\partial_s {\cal H}^i_s$ is
782: bounded as follows. We assume that there is a $D(E)$ such that $\sum_{a\in
783: high,
784: |E_a| \leq E}|\langle \Psi_{a,s}|u^i_s| \Psi_{\rm low,s}\rangle|^2\leq D(E)
785: ||u^i_s||$, where $||...||$ denotes the operator norm. The sum is restricted
786: to states $\Psi_a$ in the high energy sector.
787: Note that $D(E)\leq 1$
788: for all $E$. We define $\partial_E D(E)=\rho(E)$ to be the density of states
789: at energy $E$ produced by operator $u^i_s$.
790:
791: There are a number of systems for which this density of states bound is
792: relevant. For a discrete gauge theory, there will be a set of topological
793: excitations below a gap. These topological excitations
794: form the low energy sector, and thus $D(E)$ vanishes below the gap.
795: For a transverse-field Ising system, ${\cal H}^i=J
796: \sum_{<i,j>} \si^z_i \si^z_j+B \si^y_i$, in the ferromagnetic phase, there
797: will again be two almost degenerate states below a finite gap. At $B=0$, these
798: states correspond to symmetric and anti-symmetric combinations of all spins up
799: or all spins down.
800:
801: The bound $D(E)$ implies a locality bound\cite{H0402}. Suppose some operator
802: $O$ acts only on some set of sites, $S_O$. We define the distance between a
803: site $j$ and the operator $O$ to be equal to $d(j,O)={\rm min}_{i\in
804: S_O}(d(i,j))$, with the minimum is taken over sites $i \in S_O$, and we define
805: the distance between two operators $O_1,O_2$ to be $d(O_1,O_2)={\rm min}_{i\in
806: S_{O_1},j \in S_{O_2}}(d(i,j))$. Suppose the system has a unique ground
807: state. Then if the system is gapped, so that $D(E)=0$ for $E\leq \Delta E$,
808: the connected expectation value $|\langle O u^i_s \rangle_s -\langle O
809: \rangle_s \langle u^i_s \rangle_s|$ is exponentially decaying in $l$, where
810: $\langle ... \rangle_s$ denotes the ground state expectation value with
811: Hamiltonian $\cH_s$. On the other hand, if $D(E)\propto E^{\alpha}$, then
812: $|\langle O u^i_s \rangle_s -\langle O \rangle_s \langle u^i_s \rangle_s|$ is
813: bounded by some constant times $d(i,O)^{-\alpha/2}$.
814:
815: \subsection{Definition of Quasi-Adiabatic Continuation}
816:
817: We introduce the unitary operator
818: \begin{eqnarray}
819: \label{weq}
820: &&\t V(s) \\
821: &=&{\cal S}'\exp\left\{-\int_0^{s}
822: {\rm d}{s'}
823: \int_0^{\infty} {\rm d}\tau
824: e^{-(\tau/t_q)^2/2}
825: %\times \\ \nonumber
826: [\tilde u_{s'}^+(i\tau)-
827: h.c.]
828: \right\}, \nonumber
829: \end{eqnarray}
830: where the symbol ${\cal S}'$ denotes that the exponential is $S'$-ordered, in
831: analogy to the usual time ordered or path ordered exponentials. We define
832: $u_{s}=\partial_s{\cal H}_{s}=\sum_i u^i_s$, and define $\tilde
833: u_{s}^+(i\tau)$ following Ref. \cite{H0402}: for any operator $A$
834: \begin{eqnarray}
835: {\tilde A}(t)\equiv A(t)
836: \exp[-(t/t_q)^2/2]
837: , \\
838: \label{tdef}
839: {\tilde A}^{\pm}(\pm i\tau)
840: =\frac{1}{2\pi}
841: \int {\rm d}t \,
842: {\tilde A}(t)\frac{1}{\pm it+\tau}.
843: \end{eqnarray}
844: The time $t_q$ will be chosen later. For comparison with previous
845: work\cite{H0431,H0402}, $t_q$
846: is what was previously called $\sqrt{q}/\Delta E$; in this work, we also
847: consider the possibility of gapless theories where there is no scale $\Delta
848: E$. The time evolution of operators is defined by $A(t)=\exp[i{\cal H}_{s'}
849: t]A\exp[-i{\cal H}_{s'}t]$. The Hermitian conjugate in Eq.~(\ref{weq}) of
850: $\tilde u_{s}^+(i\tau)$ is $\tilde u_{s}^-(-i\tau)$, and $\t V(s)$ is a unitary
851: operator.
852:
853: At $t_q=\infty$, the operator $\t V(s)$ becomes equal to
854: $V(s)=\sum_a\Psi_{a,s}\rangle \langle\Psi_{\rm low,s}$. To see this, note that at
855: $t_q=\infty$, we have $\tilde A^+(i\tau)=A^+(i\tau)$, where $A^+(i\tau)$ is
856: the positive energy part of $A$ taken at imaginary time $i\tau$. That is, in
857: a basis of eigenstates of ${\cal H}_s$ with energies $E_a,E_b$, we have matrix
858: elements $A^+(i\tau)_{ab}= A_{ab}\Theta(E_a-E_b)\exp[-\tau(E_a-E_b)]$, where
859: $\Theta(x)$ is the step function. Then, $-\int_0^{\infty} {\rm d}\tau
860: [u^+(i\tau)-u^-(-i\tau)]= -(E_a-E_b)^{-1} \partial_s{\cal H}_s$, which gives
861: the result of linear perturbation theory for the adiabatic evolution of
862: quantum states with a change in the Hamiltonian. We instead keep $t_q$ finite
863: to define a ``quasi-adiabatic" evolution, which will transform local operators
864: into local operators. To show that keeping $t_q$ finite maintains the
865: locality we will rely on finite group velocity results, while we will use the
866: gap to show that we get only a small error in the ground state expectation
867: values by taking a finite $t_q$; detailed proofs of this are in the Appendix,
868: while the physical discussion is given in the next section.
869: We will relate the time $t_q$ to the scale $l$ by Eq.~(\ref{lscale}) below.
870:
871: \subsection{Results}
872: \label{res}
873:
874: For any operator $O$, we define $O(s)=\t V(s) O \t V(s)^{\dagger}$, where
875: $O(s)$ has been ``smeared out" over a scale $l$ given by Eq.~(\ref{lscale})
876: below. For a gapped theory, we only need to take a length $l$ of order the
877: correlation length $\xi$ to get a small error in Eq.~(\ref{boundGap}). To
878: understand how this works, define $Q(s)=\tilde V(s) V(s)^{\dagger}$. Then,
879: for a state $| \Psi_{\rm low,s} \rangle$ in a low energy subspace, $\langle
880: \Psi_{\rm low,s} | O(s) | \Psi_{\rm low,s} \rangle= \langle \Psi_{\rm low,0} |
881: Q(s) O Q^{\dagger}(s) | \Psi_{\rm low,0} \rangle$.
882: If we can show that $Q(s)$ has nearly vanishing matrix elements
883: between the low energy and high energy states,
884: %If we can show that $Q(s)$
885: %is close to unity, outside the low energy subspace,
886: we can bound the difference in expectation values in Eq.~(\ref{boundGap}).
887: The gap between the states in the low energy subspace and the high energy
888: subspace enables us to show this; for a given gap $\Delta E$, if we perform
889: the quasi-adiabatic continuation sufficiently slowly, we can show that
890: (loosely speaking) the matrix elements of $Q(s)$ between the low energy and
891: high energy states almost vanishes and $Q(s)$ almost acts within the low
892: energy subspace as an unitary operator.
893: %Thus, acting on a low energy state, $Q(s)$ produces
894: %excitations above the gap, but we can bound the matrix elements between the
895: %ground and excited states.
896:
897: However, there is one complication: in a
898: macroscopic system, $Q(s)$ produces excitations throughout the sample. Thus,
899: for infinitesimal $s$, $Q(s)=1+s \sum_i e^i + ...$, where the
900: anti-Hermitian operator $e^i$ creates
901: local excitations near site $i$. The probability that $Q(s)$ produces some
902: excitation acting on the ground state diverges with the system size.
903: Strictly speaking, the matrix elements of $Q(s)$ between the low energy and
904: high energy states are not small.
905: However,
906: the terms $e^i$ in $Q$ with $i$ sufficiently far from $O$ can be commuted
907: through $O$ and do not affect the final expectation value.
908: We will show that
909: the error from the terms $e^i$ with $i$ near $O$ is
910: exponentially small in $l$ divided by the correlation length.
911: The proof in the appendix follows this argument, using a triangle inequality
912: to make the bound precise.
913:
914: If $D(E)$ is not gapped, but has a power law behavior, $D(E)\propto
915: E^{\alpha}$ for small $\alpha$, then in a $d$-dimensional system we can still
916: find a local $O(s)$ if $\alpha/2>d+1$ according to
917: Eq.~(\ref{boundNoGap}). This requirement on the exponent can
918: be physically understood as follows: if we keep $t_q$ finite so that $O(s)$ is
919: spread out over a length scale $l$ under the quasi-adiabatic evolution, then
920: we have to worry about any error in $Q(s)$ on the length scale $l$.
921: The correlation
922: function of $u_s$ with $O$ decays as $l^{-\alpha/2}$, and this
923: also is how the error terms decay, while the space-time
924: volume at scale $l$ is of order $l^{d+1}$. Thus, this
925: requirement can be understood in terms of the relevance or irrelevance of
926: $u_s$ at large scales in the given correlation function.
927:
928: This physical description is based on two technical results. First, we
929: claim that $O(s)$ is {\it local} up to a scale $l$. Specifically, we claim
930: that for any operator $O_j$ which acts only on site $j$,
931: \begin{eqnarray}
932: \label{le}
933: ||[O_j,O(s)]||\leq X_1 ||O_j|| ||O|| \times \\ \nonumber
934: {\rm max}(\exp[-d(j,O)/\xi_C],\exp[-(d(j,O)/l_q)^2/2]),
935: \end{eqnarray}
936: for some constant $X_1$ of order unity and constant $l_q$ of order $t_q/c_1$,
937: where $c_1$ is a characteristic inverse velocity of the system. The length
938: $\xi_C$ is a microscopic length of order the interaction range of the
939: Hamiltonian and is defined later. Eq.~(\ref{le}) implies that, while
940: $O(s)$ does involve sites more than a distance $l_q$ from $S_O$,
941: the commutator becomes exponentially small once $d(j,O)$ becomes larger than
942: $l_q$. To relate $l_q$ and $l$, we ask for the commutator to
943: be smaller than the error in Eq.~(\ref{boundNoGap}) or Eq.~(\ref{boundGap})
944: once $d(j,O)$ becomes larger than $l$.
945: For a gapless theory, we only need to take
946: $d(j,O)$ logarithmically larger than $l_q$ before the exponential decay becomes
947: much smaller than $l^{d+1-\alpha/2}$; in this case, $l$ is of order $l_q$.
948: For a gapped theory, we need that $\exp[-(l/l_q)^2]\sim \exp[-l/\xi]$, so
949: that
950: \be
951: \label{lscale}
952: l\sim l_q^2/\xi \sim t_q^2 \Delta E/c_1.
953: \ee
954: The correlation length $\xi$ is at most of order $1/(c_1 \Delta E)$.
955:
956: If one prefers to have an operator which involves only sites
957: within a distance $l$ of $S_O$,
958: and {\it exactly} commutes with sites more
959: than a distance $l$ from $S_O$, one may
960: define
961: an operator $O_{trunc}(s)$ such that
962: \begin{widetext}
963: \begin{eqnarray}
964: \label{otb}
965: ||O_{trunc}(s)-O(s)||\leq X_2
966: \sum_{j, d(i,j)\geq l}
967: {\rm max}(\exp[-d(j,O)/\xi_C],\exp[-(d(j,O)/l_q)^2/2]),
968: \end{eqnarray}
969: \end{widetext}
970: for some constant $X_2$ of order unity. In order for the error
971: in Eq.~(\ref{otb}) to be of order the error in Eq.~(\ref{boundNoGap}) or
972: Eq.~(\ref{boundGap}), we need to pick $l$ as above: $l\sim l_q^2/\xi$ in
973: the gapped case, up to logarithmic corrections.
974:
975: Second, we claim that $O(s)$ has almost the same expectation value as
976: $O_{\rm adiab}(s)$, up to the unitary matrix $Q_0$.
977: Specifically, we show that for $0\leq s \leq 1$
978: \begin{eqnarray}
979: \label{ce}
980: |\langle \Psi_{\rm low,s}|
981: Q_0(s)^{\dagger} O(s) Q_0(s) -O_{\rm adiab}|\Psi_{\rm low,s}\rangle|
982: \\ \nonumber
983: \leq
984: 2 ||O|| [c_2(s)+c_3(s)],
985: \end{eqnarray}
986: where $c_2(s),c_3(s)$ are given by Eqs.~(\ref{teq},\ref{difb}) in
987: the Appendix.
988: One may verify from the calculation in the Appendix that
989: the error term, $c_2(s)+c_3(s)$, gives the error described above in
990: Eqs~(\ref{boundGap},\ref{boundNoGap}).
991: Note that if
992: $O=O_1 O_2 ... O_n$, then $O(s)=O_1(s) O_2(s) ... O_n(s)$, so the expectation
993: values of products of operators are also approximately preserved under this
994: quasi-adiabatic evolution.
995: Thus,
996: $|\langle \Psi_{\rm low,s}|Q_0(s)^{\dagger} O_1(s) O_2(s) ... Q_0(s)-
997: O_{1,{\rm adiab}}
998: O_{2,{\rm adiab}} ... |\Psi_{\rm low,s} \rangle| \leq
999: 2 ||O_1|| ||O_2|| ... [c_2(s)+c_3(s)]$, where here the $c_2(s),c_3(s)$
1000: are the error terms appropriate for the product operator $O_1O_2...$
1001:
1002: \section{Examples}
1003: We illustrate the quasi-adiabatic continuation by a series of
1004: examples. We start with local operators, considering a system with
1005: $Z_2$ (Ising) symmetry, a fractional quantum Hall system,
1006: and then a system with $U(1)$ symmetry. We then
1007: repeat the process in the case of non-local, string operators.
1008:
1009: \subsection{Local Ising Model}
1010: The first and simplest example is a quantum Ising
1011: ferromagnet
1012: in a transverse field. Let the Hamiltonian be
1013: ${\cal H}=J \sum_{<i,j>} \si^z_i \si^z_j+B \sum_i \si^y_i$
1014: where each site has a spin-$1/2$ and
1015: the ferromagnetic interaction $J$ couples nearest neighbor spins on the
1016: lattice.
1017:
1018: For $B=0$, the system has two exact ground states, one state which
1019: we denote $\Psi_{\uparrow}$ with all spins
1020: up and one state $\Psi_{\downarrow}$ with all spins down.
1021: The gap to the lowest excited state above these two states
1022: is $2Jq$ where $q$ is the coordination
1023: number of the lattice; this state has one flipped spin.
1024: We define symmetric and anti-symmetric combinations
1025: $\Psi_{S,A}=(1/\sqrt{2})(\Psi_{\uparrow}\pm \Psi_{\downarrow})$. The states
1026: $\Psi_{S,A}$
1027: are eigenvectors of the operator
1028: $\prod_i \si^y_i$, with eigenvalues $\pm 1$. This operator,
1029: which flips the spin on every site, commutes with the Hamiltonian for
1030: all $B$.
1031:
1032: We now consider the quasi-adiabatic continuation with
1033: ${\cal H}_s=
1034: J \sum_{<i,j>} \si^z_i \si^z_j+s B \sum_i \si^y_i$. At $s=0$, this
1035: has the exact ground states $\Psi_{S,A}$.
1036: For $sB$ small enough, the two lowest eigenstates of ${\cal H}_s$
1037: are adiabatic continuations of
1038: $\Psi_{S,A}$. Hence,
1039: the matrix element of the operator $\partial_s {\cal H}_s=
1040: B \sum_i \si^y_i$
1041: between these two states vanishes for all $B$, since
1042: $\partial_s {\cal H}_s$ commutes with $\prod_i \si^y_i$ and these
1043: two states are eigenstates of $\prod_i \si^y_i$ with different eigenvalues.
1044:
1045: Above these two states, there is a gap to the rest of the spectrum. It is
1046: known that for small enough $B$, the gap will remain open for all $s$ with
1047: $0\leq s\leq 1$. Thus, we can perform the continuation. The
1048: vanishing of the matrix elements of $\partial_s \cH_s$ between the
1049: two low energy states implies that $Q_0(s)$ is equal to the identity
1050: matrix. We consider the
1051: continuation of the operator $\si^z_i$. The ground state for $B=0$ may be
1052: taken to be $\Psi_S$. For $B=0$, we have $\langle \Psi_S |
1053: \si^z_{i_1} \si^z_{i_2} ...
1054: \si^z_{i_n} | \Psi_S \rangle=1$ for $n$ even and $-1$ for $n$ odd.
1055: The quasi-adiabatic
1056: continuation $\si^z_i \to \si^z_i(s)$ spreads
1057: out the $\si^z_i$ operators over a distance $l$. Certainly $l$ is less
1058: than the linear size of the system.
1059: % of order
1060: %the inverse gap, which is of order the correlation length $\xi$.
1061: Then $\langle \Psi_{S,s} |
1062: \si^z_{i_1}(s) \si^z_{i_2}(s) ... \si^z_{i_n}(s) |\Psi_{S,s}
1063: \rangle$,
1064: again equal to $1$ or $-1$, up to some error, depending on whether $n$ is even
1065: or odd. Here, $\Psi_{S,s}=V(s)\Psi_S$.
1066: The error is exponentially small in $l/\xi$. We conjecture that
1067: the operators $\si^z_{i_1}(s)$ correspond to {\it block spin} operators: they
1068: are equal to plus or minus $1$ depending on whether the average spin over a
1069: correlation volume is positive or negative.
1070:
1071: If we instead consider the Hamiltonian
1072: ${\cal H}_s=J \sum_{<i,j>} \si^z_i \si^z_j+
1073: s B \sum_i \si^z_i$, the operator $\partial_s {\cal H}_s$ will have
1074: non-vanishing
1075: matrix elements between the states $\Psi_{S,A}$, and thus we must start
1076: with states $\Psi_{\uparrow},\Psi_{\downarrow}$ to perform the continuation
1077: if we want to have $Q_0(s)$ be equal to the identity.
1078: In this case, $\si^z_i(s)=\si^z_i$ and $\langle \Psi_{\uparrow,s} |
1079: \si^z_{i_1}(s)
1080: \si^z_{i_2}(s) ... \si^z_{i_n}(s) | \Psi_{\uparrow,s} \rangle=(+1)^n$,
1081: where we have assumed that the sign of $B$ is such that $\Psi_{\uparrow,s}$
1082: is the ground state, rather than $\Psi_{\downarrow,s}$. For the opposite
1083: sign of $B$, the correlation function is instead $(-1)^n$.
1084:
1085: The two cases, depending on the different ways to add the magnetic field
1086: transverse or parallel to the $z$-axis, lead to different ground
1087: state correlation functions. In the second case, the correlation
1088: function of the continued operator
1089: is not a very interesting result; the unit operator would have
1090: the same correlation function. However, in the first case, the ability
1091: to find a continuation of the operator with the given correlation functions
1092: is a much more interesting result. There are long-range correlations
1093: in the operator, since $\langle \Psi_{S,s}|
1094: \si^z_{i_1}(s) \si^z_{i_2}(s) |\Psi_{S,s} \rangle-
1095: \langle \Psi_{S,s}|
1096: \si^z_{i_1}(s) |\Psi_{S,s} \rangle \langle \Psi_{S,s} |
1097: \si^z_{i_2}(s) |\Psi_{S,s}\rangle$ is
1098: non-vanishing even when sites $i_1,i_2$ are far from each other, and
1099: in particular even when
1100: the distance between $i_1,i_2$ much larger than $l$. This
1101: implies\cite{H0402} the presence of another state close in energy to the
1102: ground state when the magnetic field is added in the transverse
1103: direction.
1104:
1105: However, we would like to show the ground state degeneracy
1106: in another way, based directly on continuing the states
1107: $\Psi_{S,A}$. This way will be very valuable for
1108: more complicated systems such as the quantum Hall system.
1109: We will consider a more general class of Hamiltonians ${\cal H}_s$:
1110: we consider arbitrary Hamiltonians which are sums of local
1111: terms, and which commute with $\prod_i \si^y_i$.
1112: We show that, under the assumption that the gap between the
1113: two lowest states and
1114: the rest of the spectrum remains open, the energy difference
1115: between the continuation of the two lowest
1116: states is exponentially small. As long
1117: as that gap remains open, it is possible to adiabatically continue
1118: these two states, giving states $\Psi_{S,s}=V(s)\Psi_{S}$ and
1119: $\Psi_{A,s}=V(s)\Psi_{A}$ as eigenstates
1120: of ${\cal H}_s$. Here, we rely on the fact that ${\cal H}_s$
1121: commutes with $\prod_i \si^y_i$ and thus has vanishing
1122: matrix elements between the states.
1123:
1124: To compute
1125: the difference in energies for given $s$, we compute
1126: $\langle \Psi_{A,s} |{\cal H}_s |\Psi_{A,s} \rangle-
1127: \langle \Psi_{S,s} |{\cal H}_s |\Psi_{S,s} \rangle=
1128: \langle \Psi_{A} |V(s)^{\dagger} {\cal H}_s V(s) |\Psi_{A} \rangle-
1129: \langle \Psi_{S} |V(s)^{\dagger} {\cal H}_s V(s) |\Psi_{S} \rangle$.
1130: Thus, $V(s)^{\dagger} {\cal H}_s V(s)$ defines
1131: the continuation of the operator $\cH_s$ from $s$ back to $0$ and
1132: $\t V(s)^{\dagger} {\cal H}_s \t V(s)$ defines a quasi-adiabatic
1133: continuation from $s$ back to $0$:
1134: $\langle \Psi_{S,s} | {\cal H}_s | \Psi_{S,s} \rangle$
1135: is approximately equal to
1136: $\langle \Psi_{S} | \t V(s)^{\dagger} {\cal H}_s \t V(s) | \Psi_{S} \rangle$.
1137: The error in this continuation
1138: is exponentially small in $l/\xi$. The operator
1139: ${\cal H}_s$ is a sum of local operators, while the operator
1140: $\t V(s)^{\dagger} {\cal H}_s \t V(s)$ is a sum of terms spread
1141: out over length scale $l$. The only operators $O$ such that
1142: $\langle \Psi_{A}| O |\Psi_{A} \rangle-
1143: \langle \Psi_{S}| O |\Psi_{S} \rangle\neq 0$ are operators which
1144: flip every spin in the system and thus have non-vanishing
1145: matrix elements between $\Psi_{\uparrow}$ and $\Psi_{\downarrow}$.
1146: Thus all {\it local} operators have vanishing matrix elements
1147: between the two states $\Psi_{\uparrow}$ and $\Psi_{\downarrow}$.
1148: In particular, if the length scale
1149: $l$ is smaller than the system size $L$, then
1150: $\langle \Psi_{A}| \t V(s)^{\dagger} {\cal H}_s \t V(s) |\Psi_{A} \rangle=
1151: \langle \Psi_{S}| \t V(s)^{\dagger} {\cal H}_s \t V(s) |\Psi_{S} \rangle$.
1152: Thus we can pick $l$ just smaller then the system size to
1153: show that
1154: $\langle \Psi_{A}| V(s)^{\dagger} {\cal H}_s V(s) |\Psi_{A} \rangle-
1155: \langle \Psi_{S}| V(s)^{\dagger} {\cal H}_s V(s) |\Psi_{S} \rangle$
1156: is of order $||{\cal H}_s|| \exp(-L/\xi)\sim L^d \exp(-L/\xi)$.
1157: Here, the bound for this system is not a very tight bound: one
1158: expects the level splitting to be exponentially small in
1159: $(L/\xi)^d$ instead.
1160:
1161: The key steps in this argument were that $(1)$: matrix
1162: elements of operators which commute with $\prod_i \si^y_i$ vanish
1163: between $\Psi_{S}$ and $\Psi_{A}$; and $(2)$: all
1164: local operators have the same expectation values in state
1165: $\Psi_A$ as in state $\Psi_S$.
1166:
1167: \subsection{Quantum Hall Effect}
1168:
1169: We now turn to the case of the fractional quantum Hall effect.
1170: Consider a system with no disorder on a torus at filling factor
1171: $p/q$, with $p$ and $q$ coprime. The magnetic translation
1172: group implies at least a $q$-fold degeneracy of the ground
1173: state\cite{WNtop}. Assume that in the absence of disorder
1174: there is a $q$-fold degenerate ground state, with a gap
1175: to the rest of the spectrum. Now, consider adding disorder to the
1176: system, defining ${\cal H}_s={\cal H}_0+s \int dx dy
1177: U(x,y) \Psi^{\dagger}(x,y) \Psi(x,y)$, where $U(x,y)$ is
1178: a disorder potential and ${\cal H}_0$ is the clean Hamiltonian.
1179: Wen and Niu argued\cite{WNtop} that to first order in $s U(x,y)$ the
1180: splitting between the $q$-fold degenerate states was exponentially
1181: small. We will use the continuation to show that the splitting
1182: is exponentially small for non-vanishing disorder strength
1183: under the sole assumptions that
1184: the gap to the rest of the spectrum remains open and
1185: that at $s=0$ all local operators have the same expectation value
1186: up to exponentially small terms in the $q$ lowest states
1187: and that at $s=0$ all local operators have exponentially small
1188: matrix elements between the $q$ lowest states. We note
1189: that since these last statements involve only $s=0$ they
1190: can be checked in specific model systems without disorder, and
1191: in fact were checked
1192: in \cite{WNtop} when they showed the exponentially small
1193: splitting of the $q$ lowest states in linear perturbation theory.
1194:
1195: Suppose, then that for $0\leq s \leq 1$ there are
1196: $q$ states, $\Psi_{n,s}$ for $n=0...q-1$,
1197: and a gap to the rest of the
1198: spectrum. We wish to show that these states are close
1199: in energy. Thus, we compute
1200: $ \langle \Psi_{m,s} | {\cal H}_s |\Psi_{m,s} \rangle-
1201: \langle \Psi_{n,s} | {\cal H}_s |\Psi_{n,s} \rangle$,
1202: for some $m,n=0...q-1$.
1203: As above, $V(s)^{\dagger} {\cal H}_s V(s)$ defines
1204: the continuation of the operator $\cH_s$ from $s$ back to $0$, and
1205: $\t V(s)^{\dagger} {\cal H}_s \t V(s)$ defines a quasi-adiabatic
1206: continuation from $s$ back to $0$.
1207:
1208: However, unlike the case of the
1209: quantum Ising model, we do not have any symmetries to make
1210: matrix elements of ${\cal H}_s$ and $\partial_s {\cal H}_s$ vanish
1211: between the low lying states. Thus, we do not have any control on
1212: the matrix $Q_0(s)$.
1213: Then,
1214: $\langle \Psi_{m,s} | {\cal H}_s | \Psi_{m,s} \rangle$
1215: is equal to
1216: $\langle \Psi_{m,0} | Q_0(s)^{\dagger} \t V(s)^{\dagger} {\cal H}_s \t V(s) Q_0(s) | \Psi_{m,0} \rangle$,
1217: up to an error of order $||{\cal H}_s|| \exp(-l/\xi)$. Since we
1218: are continuing from non-zero $s$ to $s=0$, now the matrix $Q_0$ acts
1219: within the low energy sector of $\cH_0$.
1220:
1221: However, this matrix $Q_0(s)$ causes no problem. As before, for $l<L$,
1222: $\t V(s)^{\dagger} {\cal H}_s \t V(s)$ is a local operator, and then
1223: under the assumptions above, the expectation value
1224: $\langle \Psi_{m,0} | Q_0(s)^{\dagger} \t V(s)^{\dagger} {\cal H}_s \t V(s) Q_0(s) | \Psi_{m,0} \rangle$ is independent of $Q_0(s)$, up to
1225: exponentially small corrections, thus giving the desired result.
1226:
1227: \subsection{Local Rotor Model}
1228: The next example is a system with a continuous symmetry. We take the
1229: Hamiltonian ${\cal H}_s= k(s)^{-1}\sum_i \Pi_i^2 + k(s)\sum_{<i,j>}
1230: z_i \overline{z_j}$, where $z_i$ is a continuous complex field with
1231: $|z_i|=1$, and $\Pi_i$ is a
1232: momentum with $[z_i,\Pi_i]=i z_i$. The parameter $k(s)$ is an $s$-dependent
1233: stiffness of the field $z$.
1234:
1235: We pick a quasi-adiabatic continuation using $k(s)=k_0 (k_1/k_0)^s$, so
1236: $k(0)=k_0$ and $k(1)=k_1$, thus $\partial_s {\cal H}_s=
1237: \ln(k_1/k_0)
1238: [-k(s)^{-1} \sum_i \Pi_i^2
1239: +k(s)\sum_{<i,j>} z_i\overline{z_j}]$. We choose the initial
1240: $k_0$ to be large compared to the final $k_1$.
1241: We assume that we $k_1$ is sufficiently large that the system
1242: is still in a phase with gapless excitations and algebraic
1243: correlations. In this phase, we can compute the
1244: density of states by writing $z=\exp(i\phi)$ for some $\phi$
1245: with a Gaussian action for $\phi$.
1246: At low energy, the system acquires relativistic
1247: invariance, and thus in $d$ dimensions the density of single particle
1248: states at energy $E$ is of order $E^{d-1}$. The
1249: matrix element of $\phi$ between the ground state and such a
1250: single particle state is of order $E^{-1/2}$, and thus the
1251: integrated density of states created by $\phi$ is
1252: $D(E)\propto E^{d-1}$.
1253: The integrated density
1254: of states below energy $E$ created by $\partial_s {\cal H}_s$ instead
1255: is $KD(E)\sim \ln(k_0/k_1) E^{2d+2}$. Then,
1256: $\alpha/2=d+1$ and we are in a marginal case: the error in
1257: the continuation is of order $\log(L/l)$.
1258:
1259: \subsection{Emergent Discrete Gauge Theories}
1260:
1261: We now turn from these theories with local operators to emergent
1262: gauge theories. We consider the Hamiltonian of the emergent $Z_2$
1263: gauge theory in the introduction.
1264:
1265: For $J_1=J_2=0$, we have 4 exactly degenerate ground states on a torus, and a
1266: gap to the rest of the spectrum. For non-zero $J_1,J_2$, the deformed
1267: Hamiltonian is still local $\cH_s=\sum_{\v i} \cH_s^{\v i}$, where $\cH_s^{\v
1268: i}$ is a local operator defined near the site $\v i$. We can use the same
1269: reasoning used in FQH states to show that $\cH_s$ still has 4 exactly
1270: degenerate ground states on a torus.
1271: However, here we will use a slightly different approach. Continuing $\cH_s$
1272: from $s$ back to 0, we find that
1273: \begin{equation*}
1274: \<\Psi_{m,s}|\cH^{\v i}_s|\Psi_{n,s}\> =
1275: \<\Psi_{m,0}|V^\dagger(s)\cH^{\v i}_sV(s)|\Psi_{n,0}\>,
1276: \end{equation*}
1277: where $|\Psi_{n,s}\>$, $n=1,...,4$, are the 4 low lying states of $\cH_s$.
1278: Due to the energy gap separating the 4 low lying states with rest of states,
1279: the operator $V^\dagger(s)\cH^{\v i}_sV(s)$ is almost a local operator. More
1280: precisely, up to an error of order $e^{-l/\xi}$ and a unitary rotation $Q_0(s)$
1281: between the 4 low lying states $|\Psi_{n,0}\>$,
1282: $V^\dagger(s)\cH^{\v i}_sV(s)$ can be replaced
1283: by a truncated operator $H^{\v i}$ which only acts on sites within a distance
1284: $l$ from the site $\v i$. Thus we have
1285: \begin{equation*}
1286: \<\Psi_{m,s}|\cH^{\v i}_s|\Psi_{n,s}\> =
1287: \<\Psi_{m,0}|Q_0(s)^\dagger H^{\v i} Q_0(s)|\Psi_{n,0}\>
1288: +O(e^{-l/\xi}).
1289: \end{equation*}
1290: Since $Q_0(s)|\Psi_{n,0}\>$ are the ground states of the exactly soluble model
1291: (\ref{HZ2}), they from a irreducible representation of the algebra of the
1292: large-closed-string operators (\ref{acr}). We can choose the length
1293: $l$ over which the operators are smeared to be one
1294: quarter of the linear size $L$ of the system. In this case $H^{\v i}$ will be
1295: local enough that we can choose the positions large-closed-string operators to
1296: avoid any overlap between $H^{\v i}$ and the closed-string operators. So
1297: $H^{\v i}$ commutes with the closed-string operators. As a result, $H^{\v i}$
1298: must be proportional to the identity operator within the irreducible
1299: representation. This way, we have shown that $\<\Psi_{m,s}|\cH^{\v
1300: i}_s|\Psi_{n,s}\>\propto \delta_{mn}$ up to an error $e^{-L/4\xi}$, which
1301: implies that $\<\Psi_{m,s}|\cH_s|\Psi_{n,s}\>\propto \delta_{mn}$ up to an
1302: error $L^2e^{-L/4\xi}$. So the energy splitting between the 4 low lying
1303: states is less than $L^2e^{-L/4\xi}$ for the deformed Hamiltonian $\cH_s$.
1304:
1305: %using the same arguments as for the quantum Hall case, it is
1306: %possible to show that the splitting between the 4 lowest energy
1307: %states on a torus is exponentially small, under the
1308: %assumption that the gap to the rest of the spectrum
1309: %remains open.
1310:
1311: We can also continue any closed-string or dual closed-string
1312: operators $S(C_\text{closed})$ and
1313: obtain
1314: \begin{equation*}
1315: \<\Psi_{m,0}|S(C_\text{closed})|\Psi_{n,0}\> =
1316: \<\Psi_{m,s}|V(s)S(C_\text{closed})V^\dagger(s)|\Psi_{n,s}\>,
1317: \end{equation*}
1318: Then there exist dressed
1319: closed string operators $S^\text{dre}(C_\text{closed})$
1320: that have a width $l$ such that
1321: \begin{align*}
1322: &\<\Psi_{m,s}|Q_0(s)^\dagger S^\text{dre}(C_\text{closed}) Q_0(s)|\Psi_{n,s}\>
1323: \nonumber\\
1324: =&
1325: \<\Psi_{m,s}|V(s)S(C_\text{closed})V^\dagger(s)|\Psi_{n,s}\>
1326: +O(e^{-l/\xi}) ,
1327: \end{align*}
1328: Here $Q_0(s)$ is a unitary rotation between the 4 low lying states
1329: $|\Psi_{n,s}\>$ and $Q_0(s)$ is independent of the closed-string operators.
1330: This implies that the dressed string and dual string operators
1331: $S^\text{dre}(C_\text{closed})$ have the same algebra among the low energy
1332: states of the perturbed Hamiltonian as the original operators
1333: $S(C_\text{closed})$ do for the original Hamiltonian (up to an error
1334: $\exp(-l/\xi)$).
1335:
1336: The above also implies that the expectation of the dressed closed-string
1337: operators in the ground state of the perturbed Hamiltonian satisfies the zero
1338: law (\ref{zlaw}), up to an error of order $|S| \exp(-l/\xi)$, where $|S|$ is
1339: the string length. We see that the error is exponentially small for long
1340: strings since $l$ can be chosen to be a fraction of $|S|$.
1341:
1342:
1343: %Finally, since the continuation is a unitary transformation, it preserves the
1344: %commutation and anti-commutation relations (\ref{acr}). Thus, the continued
1345: %string and dual string operators have the same algebra among the ground states
1346: %of the perturbed Hamiltonian as the original operators do for the original
1347: %Hamiltonian.
1348:
1349: Our final task is to show the invariance of certain states under a deformed
1350: gauge transformation for $J_2\neq 0$. First consider $J_1,J_2=0$. The
1351: Hamiltonian has the 4 ground states which are invariant under the local gauge
1352: transformation $W_{\v I}$ given by Eq.~(\ref{xf}). We can create gauge
1353: invariant excited states by acting on the 4 ground states by open dual string
1354: operators. Following Eq.~(\ref{ds}), we define an open dual string operator
1355: as $\t S(\t C)=\prod_{\v i\text{ cross }\t C}\si^z_{\v i}$, where now the dual
1356: string
1357: $\t C$ is open, with two endpoints. The states created by acting with $\t S(\t
1358: C)$ are still invariant under the local $Z_2$ gauge transformations, since
1359: $[\t S(\t C),W_{\v I}]=0$. They are excited states which introduce $Z_2$
1360: gauge flux at the endpoints of the open string. By acting on the ground state
1361: with the open dual strings, we can create all gauge invariant states, and for
1362: $U\gg g$, these are the lowest energy excited states. Continuing to non-zero
1363: $J_1$ at $J_2=0$ still leaves these states invariant under the local $Z_2$ gauge
1364: transformation. Continuing to non-zero $J_2$ breaks the local $Z_2$ gauge
1365: invariance. However, these states are invariant under the deformed
1366: local $Z_2$ gauge transformation $W_{\v I}(s)$.
1367: To see this, we use the continuation to show that $\langle
1368: \Psi_{\rm low,s}| W_{\v I}(s) |\Psi_{\rm low,s} \rangle=1$, up to exponentially small error in $l/\xi$. Since
1369: $W_{\v I}(s)$ is unitary, this implies the gauge invariance of the ground
1370: state under the deformed gauge transformation up to exponentially small error.
1371: The gauge invariance, up to the same small error, of the states created by
1372: acting on the ground state with operators $\t S(\t C,s)=\t V(s) \t S(\t C,s)
1373: \t V(s)^{\dagger}$ then follows from the exact commutator $[\t S(\t C,s),W_{\v
1374: I}(s)]=0$. The size of the deformed gauge group is, however, much smaller
1375: than the original: if the linear system size is $L$, there are $(L/l)^d$
1376: different local gauge transformations rather than $L^d$ as is the case at
1377: $J_2=0$.
1378:
1379: \subsection{Emergent Continuous Gauge Theories}
1380:
1381: Our final problem is the theory with the $U(1)$ gauge symmetry,
1382: Eq~(\ref{strnetH}). The absence of a gap makes it much more difficult
1383: to obtain results on this system. We will obtain only one result,
1384: the existence of a deformed $U(1)$ gauge invariance of the system.
1385:
1386: We consider some gauge transformation acting on a site,
1387: $W_{\v I,\phi}=e^{i \phi Q_{\v I}}$. When $J_2=0$, the ground state
1388: has an exact gauge invariance: $\langle \Psi_{0}| W_{\v I,\phi} |\Psi_0
1389: \rangle=1$. If
1390: we continue to non-zero $J_2$, the ground state breaks the gauge invariance.
1391: However, we claim that the expectation
1392: value of the continued operator, $\langle \Psi_{0,s}|W_{\v I,\phi}(s) |
1393: \Psi_{0,s} \rangle$,
1394: is still close to unity.
1395: Unlike the $Z_2$ case, this does {\it not} follow simply from the results
1396: derived previously as there is no gap, and the low energy density of states
1397: is too high to use the results for gapless systems. However, it is still
1398: possible to show this result. We only very briefly sketch the argument,
1399: leaving a more detailed presentation for future work.
1400: The idea is as follows: at $J_2=0$, the system
1401: has gauge non-invariant states at an energy of order $U$ above the ground state.
1402: Under the continuation, at finite $t_q$, the operator $\t V(s)$ will take
1403: the ground state at $s=0$ into some state which is a mix of ground and
1404: excited states. That is, the expectation value
1405: $\langle \Psi_{0,s} | W_{\v I,\phi}(s) | \Psi_{0,s}
1406: \rangle=\langle \Psi_0 | Q(s)^{\dagger} W_{\v I,\phi}
1407: Q(s)| \Psi_0
1408: \rangle$, where the unitary matrix $Q(s)=\tilde V(s)^{\dagger}V(s)$
1409: mixes the ground state with the excited states. However, since all the
1410: low-lying excited states at $J_2=0$ are gauge-invariant, it is no longer
1411: necessary to show that $Q(s)^{\dagger} W_{\v I,\phi} Q(s)$ is close to
1412: $W_{\v I,\phi}$;
1413: it suffices to show that we do not mix in the states which
1414: are not gauge invariant under $W_{\v I,\phi}$ and
1415: all such states are at energy of order $U$ above the ground state. At
1416: infinitesimal $s$, the mixing into such states is exponentially small, since
1417: the energy $U$ acts like a gap; however, one has some mixing into
1418: the low-lying gauge-invariant states. As $s$ increases, one mixes
1419: into progressively higher energy states, until eventually one
1420: begins to excite the gauge non-invariant states. However, if $t_q$
1421: is sufficiently big (roughly of order $U^{-1}$) the mixing into the
1422: gauge non-invariant states can be bounded at $s=1$, thus obtaining
1423: the desired result.
1424:
1425: Now consider the excited states. At $J_2=0$,
1426: the gauge invariant
1427: excited states of the system can be obtained by acting on the
1428: ground state with operators of the form $O=\prod_{i} e^{i \phi_{i}
1429: L^z_i}$, where the phase $\phi_{i}$ is some arbitrary function of the
1430: leg. These operators commute with $W_{\v I,\phi}$ and therefore at
1431: non-zero $J_2$ the
1432: continued operators commute with the continued gauge transformation:
1433: $[O(s),W_{\v I,\phi}(s)]=0$. Therefore, there are a class of excited
1434: states for non-zero $J_2$, namely the excited states created by
1435: acting on the ground state with the operators $O(s)$, which are
1436: also gauge invariant under the deformed
1437: gauge transformation, up to the same error.
1438:
1439: As in the $Z_2$ case, the size of the deformed gauge group is much smaller
1440: than the original gauge group. The original gauge group had
1441: $L^d$ different generators, while the deformed group has only
1442: $(L/l)^d$ such generators.
1443: A more careful analysis should be able to then use this deformed
1444: gauge invariance to show that the gapless photon is protected. This is also
1445: a job for the future.
1446:
1447: The compactness of the gauge group was important here. Consider a non-compact
1448: $U(1)$ theory, with Lagrangian $(1/2)(\partial_{\mu}A_{\nu}- \partial_{\nu}
1449: A_{\mu})^2+(\lambda/2) (\partial_\mu A^\mu)^2+M A_\mu A^\mu$. The term in
1450: $\lambda$ is a gauge fixing term and $M$ is a term which breaks the gauge
1451: symmetry. With this quadratic action, it is easy to verify that a non-zero
1452: $M$ opens a gap. Why doesn't the continuation work here? The reason is that
1453: this theory has no gap to the gauge non-invariant states, unlike the compact
1454: cases before, despite the gauge fixing term. The gauge fixing term does not
1455: open a gap to the gauge non-invariant states in this case, instead it adds
1456: {\it gapless} longitudinal and scalar photons to the theory. To say it
1457: differently, in a compact theory, the charge is quantized and the open string
1458: states have a minimum possible energy because they necessarily terminate in an
1459: end with a charge that is a multiple of the charge quantum. In a non-compact
1460: theory, the charge may be arbitrarily small. So the energy cost is also
1461: arbitrarily small.
1462:
1463: \section{Discussion}
1464: The standard wisdom is that as long as gaps remain open, a system
1465: does not have a quantum phase transition and thus the long-distance
1466: structure of correlation functions remains the same. We have shown
1467: a precise form of this statement. We have found that, by appropriately
1468: dressing operators, the long-distance structure can in fact be
1469: preserved to a much greater degree than one might have expected.
1470: In particular, we have shown the presence of the zero law (\ref{zlaw})
1471: for gauge theories in the deconfined phase, and we propose this
1472: as a test of confinement. Further, we have considered the stability
1473: of topological order under perturbations of the Hamiltonian, and shown
1474: that the order is robust unless the gap to the rest of the spectrum
1475: (local excitations) closes.
1476:
1477: Topologically ordered states are described by emergent gauge theories at low
1478: energies. The topological order is closely tied to the emergent gauge
1479: invariance of the low energy gauge theories. From the point of view of the low
1480: energy gauge theory, our result shows that the emergent low energy gauge
1481: invariance is topological. It cannot be broken by any local perturbations in
1482: the parent bosonic model. We hope this result will shed light on the true
1483: meaning of gauge invariance and gauge theory.
1484:
1485: The continuation is also useful for systems without emergent local gauge
1486: structure, and may have a wider applicability. For example, an outstanding
1487: question is to prove that in some neighborhood of the AKLT
1488: point\cite{AKL8799,AKL8877}, a spin-$1$ chain remains gapped\cite{haldane}.
1489: The continuation might be useful in doing this, or at least in showing, under
1490: the assumption of the existence of a gap, the persistence of string
1491: order\cite{string} throughout the Haldane phase.
1492:
1493:
1494: {\it Acknowledgments---}
1495: MBH was supported by DOE contract W-7405-ENG-36.
1496: XGW was supported by NSF Grant No. DMR--04--33632,
1497: NSF-MRSEC Grant No. DMR--02--13282, and NFSC no. 10228408.
1498:
1499: \appendix
1500: \section{Proof of Locality and Approximation Results}
1501:
1502: \subsection{Locality Result}
1503: To show Eq.~(\ref{le}),
1504: we use the finite group velocity result, proven in \cite{LR7251,H0431}.
1505: This result uses
1506: the finite-range conditions on the Hamiltonian above to
1507: bound the commutator
1508: $||[A(t),B(0)]||$, where $A(t)=\exp(i{\cal H}_s t)A\exp(-i{\cal H}_s t)$. One
1509: can
1510: show that this commutator is exponentially small for times $t$ less than
1511: $c_1 l$ where $l$ is the distance between $A$ and $B$ and $c_1$ is
1512: some characteristic inverse velocity which depends on $J$, $R$, and
1513: the lattice structure.
1514: The specific bound is that
1515: $||[A(t),B(0)|| \leq ||A|| ||B|| \sum_j g(t,d(A,j))$,
1516: where the sum ranges over sites $j$ which
1517: appear in operator $B$ and where the function $g$ has the property that
1518: for $|t|\leq c_1 l$,
1519: $g(c_1 l,l)$ is exponentially decaying in $l$ for large $l$ with
1520: decay length $\xi_C$ for some constant $\xi_C$ which is of order $R$.
1521: Recall that $d(A,j)$ is the minimum over sites $i$ acted on by
1522: $A$ of the distance $d(i,j)$.
1523:
1524: Before giving the proof of Eq.~(\ref{le}), we give a physical description.
1525: The finite group
1526: velocity result has a very simple interpretation. Consider a local operator
1527: $A$. Under time evolution, we get an operator $A(t)$ which ``spreads out"
1528: over space as time passes. The finite group velocity result implies that
1529: for finite $t$ the operator $A$ is still local up to some
1530: length $t/c_1$ (here $c_1^{-1}$ is some characteristic velocity of
1531: the system) in the following sense: the
1532: commutator of $A(t)$ with any operator $B$ is exponentially small
1533: if $B$ is at least distance $t/c_1$ from $A(0)$. This applies
1534: in particular for $A=u_s$. Thus, in the definition
1535: (\ref{weq}) of $U$, the operators $u_{s'}^+(i\tau)$ are local in the same
1536: sense: Eq.~(\ref{tdef}) gives $\tilde u_{s'}^+(i\tau)$ as an integral over
1537: $t$ of $u_{s'}(t)$ and for $t>>t_q$ the integral is cut off exponentially,
1538: while for $t\sim t_q$ the $u_{s'}(t)$ are local up to length of order
1539: $t_q/c_1$. Then, we define $O(s)$ by the unitary transformation
1540: $\t V(s)$. We view this unitary transformation as defining a fictitious
1541: time evolution with time parameter $s$ and Hamiltonian given by
1542: the exponent of Eq.~(\ref{weq}). We have just established that this
1543: exponent is local and we can then apply the finite group velocity
1544: result to this evolution to show that $O(s)$ is also local. The rest
1545: of this subsection consists of a few precise error bounds following
1546: these statements.
1547:
1548: We compute the
1549: commutator $[\tilde u^{i+}_s(i\tau),O_j]$ where $O_j$ is some operator
1550: which acts only on site $j$, and
1551: where $\tilde u^{i+}_s(i\tau)$ is defined following Eq.~(\ref{tdef})
1552: taking $A=u^i_s$.
1553: We separate the integral over times $t$ into
1554: times with $t<c_1 l$, where $l=d(i,j)-R$, and times with
1555: $|t|>c_1 l$. For the first set of times, we have
1556: $(2\pi)^{-1} \int_{|t|\leq c_1 l} {\rm d}t (it+\tau) ||[u_s(t),O_j]||\leq
1557: (2\pi)^{-1} ||u^i_s|| ||O|| \exp(-l/\xi_C)$. For the
1558: second set of times, we have
1559: $(2\pi)^{-1} \int {\rm d}t (it+\tau) ||[u_s(t),O_j]||\leq
1560: \pi^{-1} ||u^i_s|| ||O|| (\sqrt{2 \pi }t_q/ c_1 l)
1561: \exp[-(c_1 l/t_q)^2/2]$. Thus, for large $l$,
1562: we find $||[\tilde u^{i+}_s(i\tau),O_j]||$ is exponentially decaying in $d(i,j)$
1563: with decay length $\xi_C$. Note that
1564: $\exp(-l/\xi_C)\approx \exp[-(c_1 l /t_q)^2/2]$ for
1565: $l\approx 2 (t_q/c_1)^{2}/\xi_C$.
1566:
1567: Now, here is the trick. We regard Eq.~(\ref{weq}) as defining the
1568: ``time" evolution of states, where the parameter $s'$ is an effective
1569: time parameter and
1570: $D=i\int_0^{\infty} {\rm d}\tau
1571: \exp[-(\tau/t_q)^2/2]
1572: [\tilde d_{s'}^+(i\tau)-
1573: h.c.]$ is some effective $s'$-dependent ``Hamiltonian", so
1574: that $\t V(s)={\cal S}'\exp[-i\int_{0}^s {\rm d}s' D]$.
1575: We write
1576: $D=\sum_i D^i$, where $D^i=\int_0^{\infty} {\rm d}\tau
1577: \exp[-(\tau/t_q)^2/2]
1578: [\tilde u_{s'}^{i+}(i\tau)-
1579: h.c.]$.
1580: Then, $||[D^i,O_j]||\leq ||O_j|| F(d(i,j))$, where
1581: the function $F(l)$ is
1582: exponentially decaying as $\exp[-l/\xi_C]$ for
1583: large $l$ and decaying as $(\sqrt{2\pi}t_q/c_1 l)\exp[-(c_1 l/ t_q)^2/2]$
1584: for small $l$.
1585:
1586: This exponential decay is in fact good enough to prove the finite group
1587: velocity result\cite{LR7251} using $D$ as an effective
1588: Hamiltonian; it is not necessary that $D^i$ act only
1589: on sites within some finite range, but an exponential decay also suffices.
1590: Following\cite{H0431}, define $G_i$ by the differential equations, for $s>0$,
1591: $\partial_s G_i(s)=\sum_j G_j(s) F(d(i,j))$ with initial conditions
1592: $G_i(0)=2$ if $i \in S_O$ and $G_i(0)=0$ otherwise. Then, one can show that
1593: $||[O(s),O_j]||\leq ||O|| ||O_j|| G_j(s)$. Solving the equations for $G_j$, one
1594: arrives at the bound (\ref{le}).
1595:
1596: One can define the operator $O_{trunc}$ by
1597: setting $\t V_{\rm trunc}(s)={\cal S}'\exp[-i\int_{0}^s {\rm d}s'
1598: {\rm Tr}_{j,d(O,j)\geq l}(D)]$, where the trace is the trace of operator
1599: $D$
1600: over sites $j$ with $d(O,j)\geq l$. Then, set $O_{trunc}(s)=\t V_{trunc}(s)
1601: O \t V_{trunc}(s)^{\dagger}$, getting Eq.~(\ref{otb}).
1602:
1603: \subsection{Approximation Result}
1604: The idea behind the proof of the approximation result is that,
1605: for any site $i$, the difference between $\tilde u_s^{i+}(i\tau)$ and
1606: $u_s^{i+}(i\tau)$ can be made small by taking large enough $t_q$.
1607: Here, $u_s^{i+}$ is the positive energy part of $u_s$ and $\tilde u_s^{i+}$
1608: is defined following Eq.~(\ref{tdef}) with $A=u_s^i.$ The difference
1609: between these two is closely related to $e^i$, as given below.
1610: Eq.~(\ref{weq}) involves summing over all sites $i$, but sites $i$ which are
1611: sufficiently far from $O$ will turn out to have little effect on defining
1612: $O(s)$. Thus, the task in this subsection is to figure out how much difference
1613: there is between
1614: $\tilde u_s^{i+}(i\tau)$ and $u_s^{i+}(i\tau)$, and then sum that error over
1615: sites near enough to $O$, giving the difference between the quasi-adiabatic
1616: continuation and the adiabatic continuation. Since the adiabatic continuation
1617: preserves expectation values, this will give an estimate in the error in
1618: the expectation values. We now do this carefully.
1619:
1620: The proof of Eq.~(\ref{ce})
1621: involves defining an additional operator and using triangle inequalities.
1622: We define
1623: \begin{eqnarray}
1624: \label{lqo}
1625: \t V_l(s)
1626: ={\cal S}'\exp\{-\int_0^{s}
1627: {\rm d}{s'}
1628: \int_0^{\infty} {\rm d}\tau
1629: \sum_i \exp[-(\tau/t_{q_i})^2/2]
1630: \times \\ \nonumber
1631: [\tilde u_{s'}^{i+}(i\tau)-
1632: h.c.]
1633: \},
1634: \end{eqnarray}
1635: where now $t_{q_i}$ may depend on $i$ and we define $\tilde u_{s}^{i+}(i\tau)$ by
1636: $\tilde u_s^{i}(t)=u_s^i(t) \exp[-(t/t_{q_i})^2/2]$, again using the
1637: $t_{q_i}$ which depend on $i$.
1638:
1639: We then pick $t_{q_i}=t_q$ for $d(i,O)\leq 2 l$, and
1640: $t_{q_i}=t_q+c_1[d(i,O)-2 l]$ otherwise. Define
1641: $O_l(s)=\t V_l(s) O \t V_l(s)^{\dagger}$.
1642: Thus, $\t V_l(s)$ has a $t_{q_i}$ which increases the further one
1643: gets from operator $O$. While the operator $\t V(s)$ would create
1644: local excitations everywhere acting on a ground state, $\t V_l(s)$ only
1645: creates local excitations near $O$.
1646:
1647: Using a triangle inequality,
1648: $|\langle \Psi_{\rm low,s} |Q_0(s) O(s) Q_0(s)^{\dagger}-O_{\rm adiab}(s)|
1649: \Psi_{\rm low,s} \rangle|
1650: \leq
1651: |\langle \Psi_{\rm low,s} |Q_0(s) O(s) Q_0(s)^{\dagger}-
1652: Q_0(s) O_l(s) Q_0(s)^{\dagger} |\Psi_{\rm low,s} \rangle|
1653: +|\langle \Psi_{\rm low,s} |
1654: Q_0(s) O_l(s) Q_0(s)^{\dagger} -O_{\rm adiab}(s)| \Psi_{\rm low,s} \rangle|$.
1655: The difference between $\t V_l(s)$ and $\t V(s)$ is the excitations
1656: far from $O$. That is, if for small $s$ $\t V(s) V(s)^{\dagger}=
1657: 1+s \sum_i e^i+...$ and $\t V_l(s) V(s)^{\dagger}=1+s \sum_i e^i_l+...$, then $e^i_l$ and
1658: $e^i$ differ only for $i$ far from $O$.
1659: However, for $i$ far from $O$, the $e^i$
1660: commute through $O$ (as discussed physically
1661: before) and so it is possible to bound the difference
1662: $|\langle \Psi_{\rm low,s} |Q_0(s) O(s) Q_0(s)^{\dagger}-
1663: Q_0(s) O_l(s) Q_0(s)^{\dagger} |\Psi_{\rm low,s} \rangle|$.
1664: Precisely,
1665: to bound the first difference, we note that the difference between the definition
1666: of $O_l(s)$ and $O(s)$ has to do terms $u_s^i$
1667: with sites $i$ which are at least a distance
1668: $2 l$ from $O$. Using the locality bound, one can bound
1669: the commutator of $O(s)$ and $O_l(s)$ with $\tilde u_s^i$ for these sites.
1670: This gives
1671: $||O(s)-O_l(s)||\leq c_3(s)$, where
1672: \begin{eqnarray}
1673: \label{difb}
1674: c_3(s)=X_3 \sum_{j,d(i,j)\geq 2 l}
1675: {\rm max}(\exp[-d(j,O)/\xi_C],\\ \nonumber
1676: \exp[-(d(j,O)/l_{q_j})^2/2]),
1677: \end{eqnarray}
1678: for some constant $X_3$ and where $l_{q_j}$ is of order $c_1 t_{q_j}$.
1679:
1680: We now bound the difference
1681: $|\langle \Psi_{\rm low,s} |
1682: Q_0(s) O_l(s) Q_0(s)^{\dagger} - O_{\rm adiab}(s)| \Psi_{\rm low,s} \rangle|$.
1683: To do this, it suffices to bound
1684: $|\langle \Psi_{\rm low,s} Q_0(s) \t V_l(s) -\langle\Psi_{\rm low,s} V(s)|$.
1685: This is equal to
1686: $|\langle \Psi_{\rm low,0} V(s)^{\dagger} Q_0(s) \t V_l(s) -
1687: \langle \Psi_{\rm low,0}|$.
1688: The operator $\t V_l(s)$ is equal to
1689: ${\cal S'} \exp\{-\int_0^s {\rm d}s'
1690: (\partial_{s'} \t V_l(s')) \t V_l(s')^{\dagger} \}$.
1691: This equals
1692: $V(s)
1693: {\cal S'} \exp\{-\int_0^s {\rm d}s'
1694: V(s')^{\dagger} [ (\partial_{s'} \t V_l(s'))
1695: \t V_l(s')^{\dagger} -
1696: (\partial_{s'} V(s')) V(s')^{\dagger} ] V(s) \}$.
1697: Thus,
1698: \begin{eqnarray}
1699: \label{grp}
1700: V(s)^{\dagger} Q_0(s) \t V_l(s) \\ \nonumber
1701: =[V(s)^{\dagger} Q_0(s) V(s)] \times \\ \nonumber
1702: {\cal S'} \exp\Bigl\{-\int_0^s {\rm d}s'
1703: V(s')^{\dagger} e_l(s')
1704: V(s') \Bigr\},
1705: \end{eqnarray}
1706: where
1707: \be
1708: e_l(s')=(\partial_{s'} \t V_l(s'))
1709: \t V_l(s')^{\dagger} V(s') -
1710: (\partial_{s'}
1711: V(s')) V(s')^{\dagger}.
1712: \ee
1713:
1714: We have grouped the
1715: operators $V(s)^{\dagger} Q_0(s) V(s)$ together in (\ref{grp}) for a reason:
1716: the matrix $Q_0(s)$ is an operator between the low energy states of
1717: $\cH_s$ so
1718: therefore the operator $V(s)^{\dagger} Q_0(s) V(s)$ is an operator
1719: between the low energy states of $\cH_0$. Now we turn to the exponential
1720: ${\cal S'} \exp\{-\int_0^s {\rm d}s'
1721: V(s')^{\dagger} e_l(s')
1722: V(s') \}$. We want to show that
1723: this operator is also equal to, up to some bounded error, an operator between
1724: the low energy states of $\cH_0$.
1725:
1726: Define $P_{high}$ to project onto the high energy states of $\cH_0$.
1727: Then, we can pick $Q_0(s)$ such that
1728: \begin{eqnarray}
1729: \label{ndd}
1730: |\langle \Psi_{\rm low,0} V(s)^{\dagger} Q_0(s) \t V_l(s) - \langle
1731: \Psi_{\rm low,0}| \\ \nonumber \leq
1732: \int_0^s {\rm d}s' |
1733: \langle \Psi_{\rm low,0} V(s')^{\dagger} e^i_l(s')
1734: V(s') P_{high} |,
1735: \end{eqnarray}
1736: We will bound the integral of Eq.~(\ref{ndd}) below; combining this bound
1737: with Eq.~(\ref{difb}) will give Eq.~(\ref{ce}).
1738:
1739: Using linear perturbation theory,
1740: \be
1741: \label{lpt}
1742: (\partial_{s'}
1743: V(s')) V(s')^{\dagger}
1744: =-\int_0^{\infty}{\rm d}\tau
1745: [u_s^+(i\tau)-h.c.]+P,
1746: \ee
1747: where
1748: $P$ only has non-vanishing matrix elements between states of the same
1749: energy: $P_{ab}=0$ if $E_a\neq E_b$.
1750:
1751: The error $e_l(s)=-P+\sum_i e^i_l(s)$ where
1752: \begin{eqnarray}
1753: e^i_l(s)=
1754: -\int_0^{\infty}{\rm d}\tau
1755: [\exp[-(\tau/t_{q_i})^2/2]\tilde u_s^{i+}(i\tau)-u_s^{i+}(i\tau)-h.c.].
1756: \end{eqnarray}
1757: For $s=0,
1758: e^i_l(s)=e^i_l$ defined above.
1759:
1760: We now bound the projection into the high energy sector
1761: $|\langle \Psi_{\rm low,s'} e^i_l(s') P_{\rm high} |$.
1762: We can show by performing
1763: some elementary integrations\cite{H0402} that, for any eigenstate $\Psi_{a,s}$
1764: with energy $E_a>\tau/t_{q_i}^2$,
1765: $|\langle \Psi_{\rm low,s}|
1766: \exp[-(\tau/t_{q_i})^2/2]\tilde u_s^{i+}(i\tau)-u_s^{i+}(i\tau)|\Psi_{\rm a,s}\rangle|
1767: \leq \exp[-(\tau/t_{q_i})^2/2]\exp[-(t_{q_i}E_a)^2/2] |(u_s^i)_{0a}|$,
1768: where
1769: $|(u_s^i)_{0a}|$ is the absolute value of the matrix element of
1770: $u_s^i$ between state $\Psi_{\rm low,s}$ and $\Psi_{a,s}$.
1771: Similarly, for $E_a <\tau/t_{q_i}^2$,
1772: $|\langle \Psi_{\rm low,s}
1773: |\exp[-(\tau/t_{q_i})^2/2]\tilde u_s^{i+}(i\tau)-u_s^{i+}(i\tau)|\Psi_{a,s}\rangle|
1774: \leq \exp[-\tau E_a]
1775: |(u_s^i)_{0a}|$. Integrating over $\tau$ and summing over states $\Psi_a$
1776: outside the sector of ground states,
1777: using the bound on density of states, we have
1778: \begin{eqnarray}
1779: |\langle \Psi_{\rm low,s} e^i_l(s)
1780: P_{high}
1781: |^2\leq \int {\rm d}E \rho(E) \times
1782: \\ \nonumber
1783: \{(K/E)2\exp[-(t_{q_i}E)^2]
1784: +K t_{q_i}
1785: \exp[-(t_{q_i}E)^2/2]\sqrt{2\pi})\}^2.
1786: \end{eqnarray}
1787:
1788: Summing over sites $i$ and using Eq.~(\ref{ndd})
1789: we obtain the bound
1790: $|\langle \Psi_{\rm low,0} V(s)^{\dagger} Q_0(s) \t V_l(s) - \langle
1791: \Psi_{\rm low,0}| \leq
1792: c_2(s)$ where we define
1793: \begin{eqnarray}
1794: \label{teq}
1795: c_2(s)=s \sum_i
1796: \Bigl( \int {\rm d}E \rho(E) \{
1797: (K/E)2\exp[-(t_{q_i}E)^2]+ \\ \nonumber K t_{q_i}
1798: \exp[-(t_{q_i}E)^2/2]\sqrt{2\pi}\}^2\Bigr)^{1/2}.
1799: \end{eqnarray}
1800: This completes the calculation.
1801:
1802: % Create the reference section using BibTeX:
1803: %\bibliography{/home/wen/bib/wencross,/home/wen/bib/all,/home/wen/bib/publst}
1804:
1805:
1806: \begin{thebibliography}{29}
1807: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1808: \expandafter\ifx\csname bibnamefont\endcsname\relax
1809: \def\bibnamefont#1{#1}\fi
1810: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1811: \def\bibfnamefont#1{#1}\fi
1812: \expandafter\ifx\csname citenamefont\endcsname\relax
1813: \def\citenamefont#1{#1}\fi
1814: \expandafter\ifx\csname url\endcsname\relax
1815: \def\url#1{\texttt{#1}}\fi
1816: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1817: \providecommand{\bibinfo}[2]{#2}
1818: \providecommand{\eprint}[2][]{\url{#2}}
1819:
1820: \bibitem[{\citenamefont{Kalmeyer and Laughlin}(1987)}]{KL8795}
1821: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Kalmeyer}} \bibnamefont{and}
1822: \bibinfo{author}{\bibfnamefont{R.~B.} \bibnamefont{Laughlin}},
1823: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{59}},
1824: \bibinfo{pages}{2095} (\bibinfo{year}{1987}).
1825:
1826: \bibitem[{\citenamefont{Wen et~al.}(1989)\citenamefont{Wen, Wilczek, and
1827: Zee}}]{WWZcsp}
1828: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1829: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Wilczek}}, \bibnamefont{and}
1830: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Zee}},
1831: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{39}},
1832: \bibinfo{pages}{11413} (\bibinfo{year}{1989}).
1833:
1834: \bibitem[{\citenamefont{Read and Sachdev}(1991)}]{RS9173}
1835: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Read}} \bibnamefont{and}
1836: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sachdev}},
1837: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{66}},
1838: \bibinfo{pages}{1773} (\bibinfo{year}{1991}).
1839:
1840: \bibitem[{\citenamefont{Wen}(1991)}]{Wsrvb}
1841: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1842: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{44}},
1843: \bibinfo{pages}{2664} (\bibinfo{year}{1991}).
1844:
1845: \bibitem[{\citenamefont{Senthil and Fisher}(2000)}]{SF0050}
1846: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Senthil}} \bibnamefont{and}
1847: \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
1848: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{62}},
1849: \bibinfo{pages}{7850} (\bibinfo{year}{2000}).
1850:
1851: \bibitem[{\citenamefont{Moessner and Sondhi}(2001)}]{MS0181}
1852: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Moessner}} \bibnamefont{and}
1853: \bibinfo{author}{\bibfnamefont{S.~L.} \bibnamefont{Sondhi}},
1854: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{86}},
1855: \bibinfo{pages}{1881} (\bibinfo{year}{2001}).
1856:
1857: \bibitem[{\citenamefont{Wen}(2002{\natexlab{a}})}]{Wqoslpub}
1858: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1859: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{65}},
1860: \bibinfo{pages}{165113} (\bibinfo{year}{2002}{\natexlab{a}}).
1861:
1862: \bibitem[{\citenamefont{Wen}(2002{\natexlab{b}})}]{Wlight}
1863: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1864: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{88}},
1865: \bibinfo{pages}{11602} (\bibinfo{year}{2002}{\natexlab{b}}).
1866:
1867: \bibitem[{\citenamefont{Sachdev and Park}(2002)}]{SP0258}
1868: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Sachdev}} \bibnamefont{and}
1869: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Park}},
1870: \bibinfo{journal}{Annals of Physics (N.Y.)} \textbf{\bibinfo{volume}{298}},
1871: \bibinfo{pages}{58} (\bibinfo{year}{2002}).
1872:
1873: \bibitem[{\citenamefont{Balents et~al.}(2002)\citenamefont{Balents, Fisher, and
1874: Girvin}}]{BFG0212}
1875: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Balents}},
1876: \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
1877: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~M.}
1878: \bibnamefont{Girvin}}, \bibinfo{journal}{Phys. Rev. B}
1879: \textbf{\bibinfo{volume}{65}}, \bibinfo{pages}{224412}
1880: (\bibinfo{year}{2002}).
1881:
1882: \bibitem[{\citenamefont{Wen}(2003{\natexlab{a}})}]{Walight}
1883: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1884: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{68}},
1885: \bibinfo{pages}{115413} (\bibinfo{year}{2003}{\natexlab{a}}).
1886:
1887: \bibitem[{\citenamefont{Ardonne et~al.}(2004)\citenamefont{Ardonne, Fendley,
1888: and Fradkin}}]{AFF0493}
1889: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Ardonne}},
1890: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Fendley}}, \bibnamefont{and}
1891: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Fradkin}},
1892: \bibinfo{journal}{Annals Phys.} \textbf{\bibinfo{volume}{310}},
1893: \bibinfo{pages}{493} (\bibinfo{year}{2004}).
1894:
1895: \bibitem[{\citenamefont{Levin and Wen}(2004)}]{LWstrnet}
1896: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Levin}} \bibnamefont{and}
1897: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1898: \bibinfo{journal}{cond-mat/0404617} (\bibinfo{year}{2004}).
1899:
1900: \bibitem[{\citenamefont{Motrunich and Senthil}(2002)}]{MS0204}
1901: \bibinfo{author}{\bibfnamefont{O.~I.} \bibnamefont{Motrunich}}
1902: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Senthil}},
1903: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{89}},
1904: \bibinfo{pages}{277004} (\bibinfo{year}{2002}).
1905:
1906: \bibitem[{\citenamefont{Moessner and Sondhi}(2003)}]{MS0312}
1907: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Moessner}} \bibnamefont{and}
1908: \bibinfo{author}{\bibfnamefont{S.~L.} \bibnamefont{Sondhi}},
1909: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{68}},
1910: \bibinfo{pages}{184512} (\bibinfo{year}{2003}).
1911:
1912: \bibitem[{\citenamefont{Hermele et~al.}(2004)\citenamefont{Hermele, Fisher, and
1913: Balents}}]{HFB0404}
1914: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Hermele}},
1915: \bibinfo{author}{\bibfnamefont{M.~P.~A.} \bibnamefont{Fisher}},
1916: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Balents}},
1917: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{69}},
1918: \bibinfo{pages}{064404} (\bibinfo{year}{2004}).
1919:
1920: \bibitem[{\citenamefont{Wen}(1990)}]{Wrig}
1921: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1922: \bibinfo{journal}{Int. J. Mod. Phys. B} \textbf{\bibinfo{volume}{4}},
1923: \bibinfo{pages}{239} (\bibinfo{year}{1990}).
1924:
1925: \bibitem[{\citenamefont{Wen and Niu}(1990)}]{WNtop}
1926: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}} \bibnamefont{and}
1927: \bibinfo{author}{\bibfnamefont{Q.}~\bibnamefont{Niu}},
1928: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{41}},
1929: \bibinfo{pages}{9377} (\bibinfo{year}{1990}).
1930:
1931: \bibitem[{\citenamefont{Wen}(2004)}]{Wen04}
1932: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1933: \emph{\bibinfo{title}{Quantum Field Theory of Many-Body Systems -- From the
1934: Origin of Sound to an Origin of Light and Electrons}}
1935: (\bibinfo{publisher}{Oxford Univ. Press}, \bibinfo{address}{Oxford},
1936: \bibinfo{year}{2004}).
1937:
1938: \bibitem[{\citenamefont{Wegner}(1971)}]{W7159}
1939: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Wegner}}, \bibinfo{journal}{J.
1940: Math. Phys.} \textbf{\bibinfo{volume}{12}}, \bibinfo{pages}{2259}
1941: (\bibinfo{year}{1971}).
1942:
1943: \bibitem[{\citenamefont{Kogut}(1979)}]{K7959}
1944: \bibinfo{author}{\bibfnamefont{J.~B.} \bibnamefont{Kogut}},
1945: \bibinfo{journal}{Rev. Mod. Phys.} \textbf{\bibinfo{volume}{51}},
1946: \bibinfo{pages}{659} (\bibinfo{year}{1979}).
1947:
1948: \bibitem[{\citenamefont{Kitaev}(2003)}]{K032}
1949: \bibinfo{author}{\bibfnamefont{A.~Y.} \bibnamefont{Kitaev}},
1950: \bibinfo{journal}{Ann. Phys. (N.Y.)} \textbf{\bibinfo{volume}{303}},
1951: \bibinfo{pages}{2} (\bibinfo{year}{2003}).
1952:
1953: \bibitem[{\citenamefont{Wen}(2003{\natexlab{b}})}]{Wqoexct}
1954: \bibinfo{author}{\bibfnamefont{X.-G.} \bibnamefont{Wen}},
1955: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{90}},
1956: \bibinfo{pages}{016803} (\bibinfo{year}{2003}{\natexlab{b}}).
1957:
1958: \bibitem[{\citenamefont{Wilson}(1974)}]{W7445}
1959: \bibinfo{author}{\bibfnamefont{K.~G.} \bibnamefont{Wilson}},
1960: \bibinfo{journal}{Phys. Rev. D} \textbf{\bibinfo{volume}{10}},
1961: \bibinfo{pages}{2445} (\bibinfo{year}{1974}).
1962:
1963: \bibitem[{\citenamefont{Anderson}(1984)}]{pwa}
1964: \bibinfo{author}{\bibfnamefont{P.~W.}\bibnamefont{Anderson}},
1965: \emph{\bibinfo{title}{Basic Notions of Condensed Matter Physics, Chapter 3}}
1966: (\bibinfo{publisher}{Addison-Wesley}, \bibinfo{address}{Reading, Mass.},
1967: \bibinfo{year}{1997}).
1968:
1969: \bibitem[{\citenamefont{Lieb and Robinson}(1972)}]{LR7251}
1970: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Lieb}} \bibnamefont{and}
1971: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Robinson}},
1972: \bibinfo{journal}{Commun. Math. Phys.} \textbf{\bibinfo{volume}{28}},
1973: \bibinfo{pages}{251} (\bibinfo{year}{1972}).
1974:
1975: \bibitem[{\citenamefont{Hastings}(2004{\natexlab{a}})}]{H0402}
1976: \bibinfo{author}{\bibfnamefont{M.~B.} \bibnamefont{Hastings}},
1977: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{93}},
1978: \bibinfo{pages}{140402} (\bibinfo{year}{2004}{\natexlab{a}}).
1979:
1980: \bibitem[{\citenamefont{Hastings}(2004{\natexlab{b}})}]{H0431}
1981: \bibinfo{author}{\bibfnamefont{M.~B.} \bibnamefont{Hastings}},
1982: \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{69}}
1983: (\bibinfo{year}{2004}{\natexlab{b}}).
1984:
1985: \bibitem[{\citenamefont{Affleck et~al.}(1987)\citenamefont{Affleck, Kennedy,
1986: Lieb, and Tasaki}}]{AKL8799}
1987: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Affleck}},
1988: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Kennedy}},
1989: \bibinfo{author}{\bibfnamefont{E.~H.} \bibnamefont{Lieb}}, \bibnamefont{and}
1990: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Tasaki}},
1991: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{59}},
1992: \bibinfo{pages}{799} (\bibinfo{year}{1987}).
1993:
1994: \bibitem[{\citenamefont{Affleck et~al.}(1988)\citenamefont{Affleck, Kennedy,
1995: Lieb, and Tasaki}}]{AKL8877}
1996: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Affleck}},
1997: \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Kennedy}},
1998: \bibinfo{author}{\bibfnamefont{E.~H.} \bibnamefont{Lieb}}, \bibnamefont{and}
1999: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Tasaki}},
2000: \bibinfo{journal}{Comm. Math. Phys.} \textbf{\bibinfo{volume}{115}},
2001: \bibinfo{pages}{477} (\bibinfo{year}{1988}).
2002:
2003: \bibitem{haldane} F. D. M. Haldane, Phys. Lett. {\bf 93A}, 464 (1983).
2004:
2005:
2006: \bibitem{string} M. den Nijs and K. Rommelse, Phys. Rev. B {\bf 40},
2007: 4709 (1989); S. M. Girvin and D. P. Arovas, Phys. Scr. T {\bf 27}, 156
2008: (1989); T. Kennedy, J. Phys.: Cond. Matt. {\bf 2}, 5737 (1990).
2009:
2010:
2011: \end{thebibliography}
2012:
2013: \end{document}
2014:
2015:
2016:
2017:
2018: