1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: %%%%%%%%%%%% VERSION: January 15, 2006 %%%%%%%%%%%%
4: %%%%%%%%%%%% %%%%%%%%%%%%
5: %%%%%%%%%%%% revtex/latex version %%%%%%%%%%%%
6: %%%%%%%%%%%% %%%%%%%%%%%%
7: %%%%%%%%%%%% Authors: C. M. Bender and B. Tan %%%%%%%%%%%%
8: %%%%%%%%%%%% %%%%%%%%%%%%
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11: %%%%%%%%%%%% %%%%%%%%%%%%
12: %%%%%%%%%%%% We are submitting this slightly revised %%%%%%%%%%%%
13: %%%%%%%%%%%% manuscript for publication in J. Phys. A %%%%%%%%%%%%
14: %%%%%%%%%%%% %%%%%%%%%%%%
15: %%%%%%%%%%%% Contact Information: %%%%%%%%%%%%
16: %%%%%%%%%%%% %%%%%%%%%%%%
17: %%%%%%%%%%%% ADDRESS: Prof. Carl M. Bender %%%%%%%%%%%%
18: %%%%%%%%%%%% Dept. of Phys. %%%%%%%%%%%%
19: %%%%%%%%%%%% Campus Box 1105 %%%%%%%%%%%%
20: %%%%%%%%%%%% Washington University %%%%%%%%%%%%
21: %%%%%%%%%%%% St. Louis, MO 63130 %%%%%%%%%%%%
22: %%%%%%%%%%%% %%%%%%%%%%%%
23: %%%%%%%%%%%% PHONE: 314-935-6216 %%%%%%%%%%%%
24: %%%%%%%%%%%% %%%%%%%%%%%%
25: %%%%%%%%%%%% FAX: 314-935-6219 %%%%%%%%%%%%
26: %%%%%%%%%%%% %%%%%%%%%%%%
27: %%%%%%%%%%%% E-MAIL: cmb@wuphys.wustl.edu %%%%%%%%%%%%
28: %%%%%%%%%%%% %%%%%%%%%%%%
29: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
31:
32: \documentclass[12pt]{iopart}
33: \usepackage{iopams}
34: \usepackage{setstack}
35: \usepackage{psfig}
36:
37: \newcommand{\half}{\mbox{$\textstyle \frac{1}{2}$}}
38: \newcommand{\cP}{\mathcal P}
39: \newcommand{\cC}{\mathcal C}
40: \newcommand{\cT}{\mathcal T}
41: \begin{document}
42:
43: \title[Calculation of the Hidden Symmetry Operator for a $\cP\cT$-Symmetric
44: Square Well]{Calculation of the Hidden Symmetry Operator for a
45: $\cP\cT$-Symmetric Square Well}
46:
47: \author[Bender and Tan]{Carl~M~Bender$^{*}$ and Barnabas~Tan$^{\dagger}$}
48:
49: \address{${}^{*}$Department of Physics, Washington University, St. Louis MO
50: 63130, USA}
51:
52: \address{${}^{\dagger}$Blackett Laboratory, Imperial College, London SW7 2BZ,
53: UK}
54:
55: \begin{abstract}
56: It has been shown that a Hamiltonian with an unbroken $\cP\cT$ symmetry also
57: possesses a hidden symmetry that is represented by the linear operator $\cC$.
58: This symmetry operator $\cC$ guarantees that the Hamiltonian acts on a Hilbert
59: space with an inner product that is both positive definite and conserved in
60: time, thereby ensuring that the Hamiltonian can be used to define a unitary
61: theory of quantum mechanics. In this paper it is shown how to construct the
62: operator $\cC$ for the $\cP\cT$-symmetric square well using perturbative
63: techniques.
64: % (\today)
65: \end{abstract}
66:
67: \submitto{\JPA}
68:
69: \pacs{11.30.Er, 11.25.Db, 11.10.Gh}
70: %\maketitle
71:
72: \section{Introduction}
73: \label{s1}
74:
75: The discovery \cite{r1,r2,r3} that there were huge classes of $\cP\cT$-symmetric
76: non-Hermitian Hamiltonians of the form $H=p^2+x^2(ix)^\epsilon$ ($\epsilon\geq
77: 0$) whose spectra were real and positive led to the investigation of many new
78: kinds of $\cP\cT$-symmetric model Hamiltonians. One particularly elegant model
79: is the $\cP\cT$-symmetric square well, whose Hamiltonian on the domain $0<x<\pi$
80: is given by
81: \begin{equation}
82: H=p^2+V(x),
83: \label{e1}
84: \end{equation}
85: where $V(x)=\infty$ for $x<0$ and $x>\pi$ and
86: \begin{equation}
87: V(x) = \left\{
88: \begin{array}{cl}
89: i\epsilon \quad & \mbox{for $\frac{\pi}{2}<x<\pi,$} \\
90: -i\epsilon \quad & \mbox{for $0<x<\frac{\pi}{2}.$} \\
91: \end{array}
92: \right.
93: \label{e2}
94: \end{equation}
95: This Hamiltonian reduces to the conventional Hermitian square well in the limit
96: as $\epsilon\to0$. For $H$ in (\ref{e1}) the parity operator $\cP$ performs a
97: reflection about $x=\frac{\pi}{2}$: $\cP:x\to\pi-x$. The $\cP\cT$-symmetric
98: square-well Hamiltonian was invented and first examined by Znojil \cite{r4} and
99: it has been heavily studied by many other researchers \cite{r5,r6,r7,r8}.
100:
101: The principal challenge in understanding non-Hermitian $\cP\cT$-symmetric
102: Hamiltonians was to show that they describe unitary time evolution. This was
103: accomplished by the discovery of a hidden symmetry operator called $\cC$. This
104: operator is used to define the Hilbert-space inner product with respect to which
105: the Hamiltonian is self-adjoint \cite{r9}. In Ref.~\cite{r9} the $\cC$ operator
106: in coordinate space was shown to have a representation as a sum over the
107: eigenfunctions $\phi_n(x)$ of the Hamiltonian:
108: \begin{equation}
109: \cC(x,y)=\sum_{n=0}^\infty\phi_n(x)\phi_n(y),
110: \label{e3}
111: \end{equation}
112: where the eigenfunctions are normalized so that they are eigenstates of the
113: $\cP\cT$ operator with eigenvalue 1,
114: \begin{equation}
115: \cP\cT\phi_n(x)=\phi_n(x),
116: \label{e4}
117: \end{equation}
118: and the integral of the square of the $n$th eigenfunction oscillates in sign:
119: \begin{equation}
120: \int dx\,[\phi_n(x)]^2=(-1)^n.
121: \label{e5}
122: \end{equation}
123:
124: The discovery of the $\cC$ operator led immediately to attempts to calculate it
125: for various model Hamiltonians. For the elementary $\cP\cT$-symmetric
126: non-Hermitian Hamiltonian $H=\half p^2+\half x^2+ix$, the exact $\cC$ operator
127: is given by \cite{r10}
128: \begin{equation}
129: \cC=e^{-2p}\cP.
130: \label{e6}
131: \end{equation}
132: However, for more complicated Hamiltonians the $\cC$ operator cannot be obtained
133: in closed form. It was shown in Ref.~\cite{r11} how to use perturbative methods
134: to evaluate the sum in (\ref{e3}) for the Hamiltonian
135: \begin{equation}
136: H=p^2+x^2+i\epsilon x^3.
137: \label{e7}
138: \end{equation}
139: In Ref.~\cite{r12} this perturbative procedure was extended to
140: quantum-mechanical Hamiltonians having several degrees of freedom.
141:
142: The perturbative methods used in Refs.~\cite{r11} and \cite{r12} were not
143: powerful enough to be used in quantum field theory, so a simple recipe for
144: finding $\cC$ was devised that can be used in systems having an infinite number
145: of degrees of freedom \cite{r10}. The procedure was to solve the three
146: simultaneous algebraic equations satisfied by $\cC$:
147: \begin{equation}
148: \cC^2=1,\quad [\cC,\cP\cT]=0,\quad [\cC,H]=0.
149: \label{e8}
150: \end{equation}
151: This recipe gives the $\cC$ operator as a product of the exponential of an
152: antisymmetric Hermitian operator $Q$ and the parity operator $\cP$:
153: \begin{equation}
154: \cC=e^Q\cP.
155: \label{e9}
156: \end{equation}
157: Note that $Q=-2p$ for the $\cC$ operator in (\ref{e6}). Mostafazadeh has shown
158: that the square root of the positive operator $e^Q$ can be used to construct a
159: similarity transformation that converts a non-Hermitian $\cP\cT$-symmetric
160: Hamiltonian $H$ to an equivalent Hermitian Hamiltonian $h$ \cite{r13}: $h=e^{-Q/
161: 2}He^{Q/2}$.
162:
163: In all the examples studied so far the $\cC$ operator is a combination of
164: integer powers of $x$ and integer numbers of derivatives multiplying the parity
165: operator $\cP$. Hence, the $Q$ operator is a polynomial in the operators $x$ and
166: $p=-i\frac{d}{dx}$. The novelty of the $\cP\cT$-symmetric square-well
167: Hamiltonian (\ref{e1}-\ref{e2}) is that $\cC$ contains {\it integrals} of $\cP$
168: and thus the $Q$ operator, while it is a simple function, is {\it not} a
169: polynomial in $x$ and $p$ and therefore cannot be found easily by the algebraic
170: perturbative methods that were introduced in Ref.~\cite{r10}. Thus, in
171: Sec.~\ref{s2} we calculate $\cC$ for this Hamiltonian by using the perturbative
172: techniques that were devised in Ref.~\cite{r11}. In Sec.~\ref{s3} we make some
173: concluding remarks.
174:
175: \section{Perturbative calculation of the $\cC$ operator}
176: \label{s2}
177:
178: The procedure we use here is as follows: First, we solve the Schr\"odinger
179: equation
180: \begin{equation}
181: -\phi_n''(x)+V(x)\phi_n(x)=E_n\phi_n(x)\quad(n=0,\,1,\,2,\,3,\,\dots)
182: \label{e10}
183: \end{equation}
184: subject to the boundary conditions $\phi_n(0)=\phi_n(\pi)=0$. We obtain the
185: eigenfunction $\phi_n(x)$ as a perturbation series to second order in powers of
186: $\epsilon$. The eigenfunctions are then normalized according to (\ref{e4}) and
187: (\ref{e5}). Next, we substitute the eigenfunctions into the formula (\ref{e3})
188: and evaluate the sum. The advantage of the domain of the square well being $0<x<
189: \pi$ is that this sum reduces to a set of Fourier sine and cosine series that
190: can be evaluated in closed form. After evaluating the sum, it is convenient to
191: translate the domain of the square well to the more symmetric region $-\frac{\pi
192: }{2}<x<\frac{\pi}{2}$. On this domain the parity operator in coordinate space is
193: $\cP(x,y)=\delta(x+y)$. Finally, we show that the $\cC$ operator to order
194: $\epsilon^2$ has the form in (\ref{e9}), and we evaluate the function $Q$ to
195: order $\epsilon^2$. Our final result for $Q(x,y)$ on the domain $-\frac{\pi}{2}<
196: x<\frac{\pi}{2}$ is
197: \begin{equation}
198: Q(x,y)=\textstyle{\frac{1}{4}}i\epsilon[x-y+\varepsilon(x-y)\,(|\,x+y\,|-\pi)]+
199: \mathcal{O}(\epsilon^3),
200: \label{e11}
201: \end{equation}
202: where $\varepsilon(x)$ is the standard step function
203: \begin{equation}
204: \varepsilon(x)=\left\{
205: \begin{array}{cl}
206: 1\quad&\mbox{($x>0$),}\\
207: 0\quad&\mbox{($x=0$),}\\
208: -1\quad&\mbox{($x<0$).}\\
209: \end{array}
210: \right.
211: \label{e12}
212: \end{equation}
213:
214: \subsection{Solution of the Schr\"odinger Equation}
215: We begin our analysis by solving the Schr\"odinger equation (\ref{e10}) in the
216: right ($x>\frac{\pi}{2}$) and left ($x<\frac{\pi}{2}$) regions of the square
217: well:
218: \begin{eqnarray}
219: &&\phi_{n,\,{\rm R}}(x)=a_n\,\bigg\{i^{\frac{1}{2}(1-(-1)^{n})}\sin (n+1)x\nonumber\\
220: &&\hspace{-2.4cm}+\left[i^{\frac{1}{2}(1+(-1)^{n})}\left(\frac{\pi}{2}-\frac{x}{2}
221: \right)\frac{(-1)^n\cos(n+1)x}{(n+1)}-\frac{1}{2}(1-(-1)^{n})\frac{\sin(n+1)x}
222: {2(n+1)^{2}}\right]\epsilon\nonumber\\
223: &&\hspace{-2.4cm}+\,i^{\half(1-(-1)^{n})}\left[\half(1+(-1)^{n})\left(\frac{x}{4
224: }-\frac{\pi}{4}\right)\frac{\cos(n+1)x}{(n+1)^3}+\left(\frac{x^2}{8}-\frac{\pi x
225: }{4}+\frac{\pi^2}{16}\right)\frac{\sin(n+1)x}{(n+1)^{2}}\right]\epsilon^2
226: \nonumber\\
227: &&\hspace{-2.4cm}+\mathcal{O}(\epsilon^3)\bigg\}\qquad\mbox{($x>\frac{\pi}{2}$
228: ),}
229: \label{e13}\\
230: \nonumber\\
231: &&\phi_{n,\,{\rm L}}(x)=a_n\,\bigg\{i^{\half(1-(-1)^{n})}\sin(n+1)x\nonumber\\
232: &&\hspace{-2.4cm}+\left[i^{\half(1+(-1)^{n})}\frac{x}{2}\frac{(-1)^n\cos(n+1)x}
233: {(n+1)}+\,\half(1-(-1)^{n})\frac{\sin(n+1)x}{2(n+1)^{2}}\right]\epsilon
234: \nonumber\\
235: &&\hspace{-2.4cm}+\,i^{\half(1-(-1)^{n})}\left[\frac{1}{2}(1+(-1)^{n})\frac{x}{4
236: }\frac{\cos(n+1)x}{(n+1)^3}+\left(\frac{x^2}{8}-\frac{\pi^2}{16}\right)\frac{
237: \sin(n+1)x}{(n+1)^2}\right]\epsilon^2\nonumber\\
238: &&\hspace{-2.4cm}+\mathcal{O}(\epsilon^3)\bigg\}\qquad\mbox{($x<\frac{\pi}{2}$
239: ).}
240: \label{e14}
241: \end{eqnarray}
242: These eigenfunctions and their first derivatives are continuous at $x=\frac{\pi}
243: {2}$.
244:
245: Having found the eigenfunction $\phi_n(x)$ to second order in $\epsilon$, we
246: can give the formula for the corresponding eigenvalues:
247: $$E_n=(n+1)^2+\frac{(-1)^n[2-(-1)^n]}{4(n+1)^2}\epsilon^2+\mathcal{O}
248: (\epsilon^4).$$
249: However, these eigenvalues are not needed for calculating the $\cC$ operator.
250:
251: \subsection{Normalization of the Eigenfunctions}
252: The normalization requirements in (\ref{e4}-\ref{e5}) give the value of the
253: coefficient $a_n$ in (\ref{e13}-\ref{e14}):
254: \begin{equation}
255: \hspace{-1.2cm}a_n=\sqrt{\frac{2}{\pi}}\left[1-(-1)^n\left(\frac{(2-(-1)^n)}
256: {(6-2(-1)^n)(n+1)^4}-\frac{(-1)^n\pi^2}{16(n+1)^2}\right)\epsilon^2
257: +\mathcal{O}(\epsilon^4)\right].
258: \label{e15}
259: \end{equation}
260: With this normalization, the $\cP\cT$ inner product between $\phi_m(x)$ and
261: $\phi_n(x)$ is $(-1)^n\delta_{mn}+\mathcal{O}(\epsilon^4)$.
262:
263: \subsection{Calculation of $\cC(x,y)$ to Leading Order (Zeroth Order) in
264: $\epsilon$}
265: The next step is to construct the operator $\cC(x,y)$, which is given in
266: (\ref{e3}) as a sum, by directly substituting the eigenfunctions $\phi_n(x)$
267: from (\ref{e13}-\ref{e14}). In general, there are four different regions of $x$
268: and $y$ to consider:
269: \begin{enumerate}
270: \item $x>\frac{\pi}{2}$, $y>\frac{\pi}{2},$
271: \item $x>\frac{\pi}{2}$, $y<\frac{\pi}{2},$
272: \item $x<\frac{\pi}{2}$, $y<\frac{\pi}{2},$
273: \item $x<\frac{\pi}{2}$, $y>\frac{\pi}{2}.$
274: \end{enumerate}
275: However, to zeroth-order in $\epsilon$, $\phi_n$ is common to all four regions
276: and the calculation is easy. We find that
277: \begin{equation}
278: \cC^{(0)}(x,y)=\frac{2}{\pi}\sum_{n=0}^{\infty}\,(-1)^n\sin(n+1)x\,\,\sin(n+1)y.
279: \label{e16}
280: \end{equation}
281: This is just the Fourier sine series for the parity operator in the range
282: $0<x<\pi$:
283: \begin{equation}
284: \cC^{(0)}(x,y)=\delta(x+y-\pi).
285: \label{e17}
286: \end{equation}
287: On the symmetric domain $-\frac{\pi}{2}<x<\frac{\pi}{2}$ this formula becomes
288: \begin{equation}
289: \cC^{(0)}(x,y)=\delta(x+y),
290: \label{e18}
291: \end{equation}
292: which is equivalent to the coordinate-space condition of completeness.
293:
294: \subsection{Calculation of $\cC(x,y)$ to First Order in $\epsilon$}
295: The calculation of $\cC(x,y)$ to first order in $\epsilon$ requires the
296: evaluation of Fourier sine and cosine series. These are expressed in terms of
297: single and double integrals of delta functions. Here, we describe the
298: calculation for the region $x>\frac{\pi}{2}$, $y>\frac{\pi}{2}$. The calculation
299: for the other three regions is similar.
300:
301: From (\ref{e13}) and (\ref{e15}) the first-order contribution to $\cC(x,y)$ is
302: \begin{eqnarray}
303: \hspace{-0.3cm}\cC^{(1)}(x,y)&=&\frac{1}{\pi}\sum_{n=0}^\infty\left[(\pi-x)\frac
304: {i(-1)^n}{n+1}\cos[(n+1)x]\,\sin[(n+1)y]\right]\nonumber\\
305: &&+\frac{1}{\pi}\sum_{n=0}^\infty\left[(\pi-y)\frac{i(-1)^n}{n+1}\sin[(n+1)
306: x]\,\cos[(n+1)y]\right]\nonumber\\
307: &&-\frac{1-(-1)^n}{\pi}\sum_{n=0}^\infty\left[i^{\frac{1}{2}(1-(-1)^n)}
308: \frac{\sin[(n+1)x]\,\sin[(n+1)y]}{(n+1)^2}\right].
309: \label{e19}
310: \end{eqnarray}
311: The first two terms of $\cC^{(1)}$ can be expressed as single integrals of the
312: parity operator, with the first having the upper limit $x$,
313: \begin{eqnarray}
314: \int_{\frac{\pi}{2}}^x dt\,\delta(t+y-\pi)&=&-\frac{2}{\pi}\sum_{n=0}^\infty
315: \left[\frac{(-1)^n}{n+1}\cos[(n+1)x]\,\sin[(n+1)y]\right]\nonumber\\
316: &&+\frac{1}{\pi}\sum_{n=0}^\infty\left[\frac{(-1)^n}{n+1}\sin[(2n+2)y]\right],
317: \label{e20}
318: \end{eqnarray}
319: and the second having the upper limit $y$,
320: \begin{eqnarray}
321: \int_{\frac{\pi}{2}}^y dt\,\delta(x+t-\pi)&=&-\frac{2}{\pi}\sum_{n=0}^\infty
322: \left[\frac{(-1)^n}{n+1}\sin[(n+1)x]\,\cos[(n+1)y]\right]\nonumber\\
323: &&+\frac{1}{\pi}\sum_{n=0}^\infty\left[\frac{(-1)^n}{n+1}\sin[(2n+2)x]\right].
324: \label{e21}
325: \end{eqnarray}
326:
327: The third term of $\cC^{(1)}$ involves sines of even $x$ and $y$. To obtain a
328: series of sines of even $x$ and $y$, we subtract the series represented by the
329: parity operator $\delta(x+y-\pi)$ from the series represented by $\delta(x-y)$.
330: The factor $(n+1)^2$ in the denominator implies that the third term of $\cC^{(1)
331: }$ may be expressed as a double integral. To maintain the symmetry, we note that
332: a double integral with respect to $x$ contributes half of the third term of
333: $\cC^{(1)}$, while a double integral with respect to $y$ contributes the other
334: half.
335:
336: This analysis allows us to express $\cC^{(1)}$ in the region $x>\frac{\pi}{2}$,
337: $y>\frac{\pi}{2}$ in terms of single integrals of the parity operator and double
338: integrals in $x$ and $y$:
339: \begin{eqnarray}
340: \cC^{(1)}(x,y)&=&i\,\Bigg\{\left(\frac{x}{2}-\frac{\pi}{2}\right)\int_{\pi/2}^x
341: dt\,\delta(t+y-\pi)\nonumber\\
342: &&+\frac{1}{4}\int_{\pi/2}^x dt\int_{\pi/2}^t ds\,\Big[\delta(s-y)-\delta(s+y-
343: \pi)\Big]\nonumber\\
344: &&+\frac{x}{4}-\frac{\pi}{4}+(x\leftrightarrow y)\Bigg\}.
345: \label{e22}
346: \end{eqnarray}
347: We simplify this result by evaluating integrals over delta functions and obtain
348: \begin{equation}
349: \cC^{(1)}(x,y)={\textstyle\frac{1}{4}}i(|x-y|+x+y-2\pi)\quad
350: \left(x>\textstyle{\frac{\pi}{2}},\,y>\textstyle{\frac{\pi}{2}}\right).
351: \label{e23}
352: \end{equation}
353:
354: The calculation of $\cC^{(1)}(x,y)$ for the remaining three regions follows a
355: similar procedure and we get
356: \begin{eqnarray}
357: \cC^{(1)}(x,y)&=&\nonumber\\
358: &&\hspace{-1.9cm}\left\{
359: \begin{array}{ll}
360: \half i[(x-\pi)\,\theta(x+y-\pi)+y\,\theta(\pi-x-y)]\quad&\mbox{$\left(x>\frac{
361: \pi}{2},\,y<\frac{\pi}{2}\right),$}\\
362: {\textstyle\frac{1}{4}}i[-|x-y|+(x+y)]\quad&\mbox{$\left(x<\frac{\pi}{2},\,y<
363: \frac{\pi}{2}\right),$}\\
364: \half i[(y-\pi)\,\theta(x+y-\pi)+x\,\theta(\pi-x-y)]\quad&\mbox{$
365: \left(x<\frac{\pi}{2},\,y>\frac{\pi}{2}\right),$}\\
366: \end{array}\right.
367: \label{e24}
368: \end{eqnarray}
369: where $\theta(x)$ is the Heaviside step function,
370: \begin{equation}
371: \theta(x)=\left\{
372: \begin{array}{cl}
373: 1\quad &\mbox{($x\geq0$),} \\ 0\quad & \mbox{($x<0$).}\\
374: \end{array}\right.
375: \label{e25}
376: \end{equation}
377: Finally, we condense the four expressions for $\cC^{(1)}(x,y)$ in the four
378: different regions into a single expression:
379: \begin{eqnarray}
380: \cC^{(1)}(x,y)&=&{\textstyle\frac{1}{4}}i[x+y-\pi-\theta(\pi-x-y)\,(|x-y|-\pi)
381: \nonumber\\
382: &&+\theta(x+y-\pi)\,(|x-y|-\pi)].
383: \label{e26}
384: \end{eqnarray}
385: On the symmetric region $-\frac{\pi}{2}<(x,y)<\frac{\pi}{2}$, this expression
386: becomes
387: \begin{eqnarray}
388: \cC^{(1)}(x,y)&=&{\textstyle\frac{1}{4}}i[x+y+\varepsilon(x+y)\,(|x-y|-\pi)].
389: \label{e27}
390: \end{eqnarray}
391: We plot the imaginary part of $\cC^{(1)}$ in Fig.~\ref{f1} as a function of
392: $x$ and $y$.
393:
394: \begin{figure}[th]\vspace{3.9in}
395: \special{psfile=Fig1.ps angle=0 hoffset=80 voffset=-33 hscale=80 vscale=80}
396: \caption{Three-dimensional plot of the imaginary part of $\cC^{(1)}(x,y)$, the
397: first-order perturbative contribution in (\ref{e27}) to the $\cC$ operator in
398: coordinate space. The plot is on the symmetric square domain $-\frac{\pi}{2}<(x,
399: y)<\frac{\pi}{2}$. Note that $\cC^{(1)}(x,y)$ vanishes on the boundary of this
400: square domain because the eigenfunctions $\phi_n(x)$ in (\ref{e3}) are required
401: to vanish at $x=0$ and $x=\pi$.}
402: \label{f1}
403: \end{figure}
404:
405: \subsection{Calculation of $\cC(x,y)$ to Second Order in $\epsilon$}
406:
407: The procedure for calculating $\cC^{(2)}(x,y)$ is similar to that used for
408: calculating $\cC^{(1)}(x,y)$, albeit more tedious. We must calculate sums of
409: products of sines and cosines, but this time the presence of factors of $(n+1
410: )^4$, $(n+1)^3$ and $(n+1)^2$ in the denominator requires the use of quadruple,
411: triple, and double integrals of delta functions to simplify the expression for
412: $\cC^{(2)}(x,y)$. We discuss the calculation of $\cC^{(2)}(x,y)$ explicitly for
413: the region $x<\frac{\pi}{2}$, $y<\frac{\pi}{2}$. The calculation of $\cC^{(2)}
414: (x,y)$ for the other three regions is similar.
415:
416: From (\ref{e13}) and (\ref{e15}) we see that the second-order calculation of
417: $\cC(x, y)$ gives
418: \begin{eqnarray}
419: \cC^{(2)}(x,y)&=&\frac{2}{\pi}\sum_{n=0}^\infty\left[\frac{x}{4}\,\frac{\cos
420: [(2n+1)x]\,\sin[(2n+1)y]}{(2n+1)^3}\right]\nonumber\\
421: &&+\frac{2}{\pi}\sum_{n=0}^\infty\left[\frac{y}{4}\,\frac{\sin[(2n+1)x]\,\cos[
422: (2n+1)y]}{(2n+1)^3}\right]\nonumber\\
423: &&+\frac{2}{\pi}\sum_{n=0}^\infty\left[\left(\frac{x^2}{8}+\frac{y^2}{8}\right)
424: \frac{(-1)^n\sin[(n+1)x]\,\sin[(n+1)y]}{(n+1)^2}\right]\nonumber\\
425: &&-\frac{2}{\pi}\sum_{n=0}^\infty\left[\frac{x}{4}\,\frac{\cos[(2n+2)x]
426: \,\sin[(2n+2)y]}{(2n+2)^3}\right]\nonumber\\
427: &&-\frac{2}{\pi}\sum_{n=0}^\infty\left[\frac{y}{4}\,\frac{\sin[(2n+2)x]
428: \,\cos[(2n+2)y]}{(2n+2)^3}\right]\nonumber\\
429: &&-\frac{2}{\pi}\sum_{n=0}^\infty\left[\frac{1}{2}\,\frac{\sin[(2n+1)x]
430: \,\sin[(2n+1)y]}{(2n+1)^4}\right]\nonumber\\
431: &&-\frac{2}{\pi}\sum_{n=0}^\infty\left[\frac{1}{2}\,\frac{\sin[(2n+2)x]
432: \,\sin[(2n+2)y]}{(2n+2)^4}\right]\nonumber\\
433: &&-\frac{2}{\pi}\sum_{n=0}^\infty\left[\frac{xy}{4}\,\frac{(-1)^n\cos[(n+1)x]\,
434: \cos[(n+1)y]}{(n+1)^2}\right].
435: \label{e28}
436: \end{eqnarray}
437: We have been able to evaluate each of these Fourier series exactly and to
438: express the result as multiple integrals over delta functions:
439: \begin{eqnarray}
440: \cC^{(2)}(x,y)&=&-{\textstyle\frac{1}{8}}x^2\int_x^{\pi/2}dt\int_t^{\pi/2}ds\,
441: \delta(s+y-\pi)\nonumber\\
442: &&-{\textstyle\frac{1}{4}}x\int_x^{\pi/2}dt\int_t^{\pi/2}ds\int_s^{\pi/2}dr\,
443: \delta(r+y-\pi)\nonumber\\
444: &&-{\textstyle\frac{1}{4}}\int_x^{\pi/2}dt\int_t^{\pi/2}ds\int_s^{\pi/2}dr
445: \int_r^{\pi/2}dp\,\delta(p-y)\nonumber\\
446: &&-{\textstyle\frac{1}{8}}xy\int_y^{\pi/2}dt\int_x^{\pi/2}ds\,\delta(s+t-\pi)+
447: {\textstyle\frac{1}{8}}x^2y\nonumber\\
448: &&-{\textstyle\frac{1}{16}}xy\pi+{\textstyle\frac{1}{24}}x^3+(x\leftrightarrow
449: y).
450: \label{e29}
451: \end{eqnarray}
452: Evaluating these integrals gives $\cC^{(2)}(x,y)$ for the region $x<\frac{\pi}
453: {2}$, $y<\frac{\pi}{2}$:
454: \begin{equation}
455: \cC^{(2)}(x,y)=-\textstyle{\frac{1}{24}}|x-y|^3+\textstyle{\frac{1}{8}}x^2y+
456: \textstyle{\frac{1}{24}}x^3-\textstyle{\frac{1}{16}}x\pi y+(x\leftrightarrow y).
457: \label{e30}
458: \end{equation}
459:
460: The calculation of $\cC^{(2)}(x, y)$ for the remaining three regions follows a
461: similar procedure. Combining the contributions from the four regions and
462: transforming to the symmetric domain $-\frac{\pi}{2}<(x,y)<\frac{\pi}{2}$,
463: we obtain the single expression:
464: \begin{eqnarray}
465: \hspace{-0.6cm}\cC^{(2)}(x,y)&=&{\textstyle\frac{1}{96}}\pi^3+{\textstyle\frac{
466: 1}{8}}xy\pi-{\textstyle\frac{1}{16}}\pi^2(x+y)\,\varepsilon(x+y)+{\textstyle
467: \frac{1}{8}}\pi(x|x|+y|y|)\,\varepsilon(x+y)\nonumber\\
468: &&-{\textstyle\frac{1}{24}}(x^3+y^3)\,\varepsilon(x+y)-{\textstyle\frac{1}{24}}
469: (y^3-x^3)\,\varepsilon(y-x)\nonumber\\
470: &&-{\textstyle\frac{1}{4}}xy\{|x|[\theta(x-y)\,\theta(-x-y)+\theta(y-x)\,\theta
471: (x+y)]\nonumber\\
472: &&+|y|[\theta(y-x)\,\theta(-x-y)+\theta(x-y)\,\theta(x+y)]\}.
473: \label{e31}
474: \end{eqnarray}
475: We have plotted the function $\cC^{(2)}(x,y)$ on the symmetric domain $-\frac{
476: \pi}{2}<(x,y)<\frac{\pi}{2}$ in Fig.~\ref{f2}.
477:
478: \begin{figure}[th]\vspace{3.9in}
479: \special{psfile=Fig2.ps angle=0 hoffset=80 voffset=-33 hscale=80 vscale=80}
480: \caption{Three-dimensional plot of $\cC^{(2)}(x,y)$ in (\ref{e31}) on the
481: symmetric square domain $-\frac{\pi}{2}<(x,y)<\frac{\pi}{2}$. The function
482: $\cC^{(2)}(x,y)$ vanishes on the boundary of this square domain because the
483: eigenfunctions $\phi_n(x)$ from which it was constructed vanish at the
484: boundaries of the square well.}
485: \label{f2}
486: \end{figure}
487:
488: In summary, our final result for the $\cC$ operator to order $\epsilon^2$ on
489: the symmetric domain $-\frac{\pi}{2}<(x,y)<\frac{\pi}{2}$ is given by
490: \begin{equation}
491: \cC(x,y)=\delta(x+y)+\epsilon\cC^{(1)}(x,y)+\epsilon^2\cC^{(2)}(x,y)+
492: \mathcal{O}(\epsilon^3),
493: \label{e32}
494: \end{equation}
495: where $\cC^{(1)}(x,y)$ is given in (\ref{e27}) and $\cC^{(1)}(x,y)$ is given
496: in (\ref{e31}). We have verified by explicit calculation that to order
497: $\epsilon^2$ this $\cC$ operator obeys the algebraic equations (\ref{e8}).
498: For example, in coordinate space the third of these equations, $\cC^2=1$, reads
499: \begin{equation}
500: \int_{-\pi/2}^{\pi/2}dy\,\cC(x,y)\,\cC(y,z)=\delta(x-z)+\mathcal{O}(\epsilon^3).
501: \label{e33}
502: \end{equation}
503:
504: \subsection{Calculation of the $Q$ operator}
505: The last step in the calculation is to determine the operator $Q$ from the
506: result in (\ref{e32}) by using (\ref{e9}). This is a long and difficult
507: calculation: We first multiply $\cC(x,z)$ on the right by $\delta(z+y)$, the
508: parity operator $\cP$ in coordinate space, and then integrate with respect to
509: $z$. This gives the coordinate space representation of $\cC\cP=e^Q$. Next, we
510: take the logarithm of the resulting expression and expand it as a series in
511: powers of $\epsilon$ to obtain $Q$. We find that the coefficient of $\epsilon^2$
512: in this expansion is zero, and thus we obtain the simple result in (\ref{e11}),
513: which is the principal result in this paper.
514:
515: \section{Conclusions}
516: \label{s3}
517:
518: In this paper we have used perturbative methods to calculate the $\cC$ operator
519: to second order in powers of $\epsilon$ for the complex $\cP\cT$-symmetric
520: square-well potential (\ref{e2}). Expressing $\cC$ in the form $e^Q\cP$, we have
521: found that the operator $Q$ for this model has an expansion in odd powers of
522: $\epsilon$, just as in the case of the cubic $\cP\cT$-symmetric oscillator whose
523: Hamiltonian is given in (\ref{e7}). Our result (\ref{e11}) for $Q$ is an
524: elementary function. We have verified our calculation of the $\cC$ operator
525: by showing that it satisfies the algebraic conditions (\ref{e8}).
526:
527: The most noteworthy property of the $\cC$ operator is that the associated
528: operator $Q$ is a nonpolynomial function, and this kind of structure had not
529: been seen in previous studies of $\cC$. At the beginning of this calculation we
530: expected that for such a simple $\cP\cT$-symmetric Hamiltonian it would be
531: possible to calculate the $\cC$ operator exactly and in closed form. We find it
532: surprising that even for this elementary model the $\cC$ operator is so
533: nontrivial.
534:
535: \vspace{0.5cm}
536: \begin{footnotesize}
537: \noindent
538: CMB is supported by the US Department of Energy.
539: \end{footnotesize}
540:
541: \vspace{0.5cm}
542:
543: \begin{thebibliography}{999}
544: \bibitem{r1} C.~M.~Bender and S.~Boettcher, Phys.~Rev.~Lett.~{\bf 80}, 5243
545: (1998).
546:
547: \bibitem{r2} C.~M.~Bender, S.~Boettcher, and P.~N.~Meisinger,
548: J.~Math.~Phys.~{\bf 40}, 2201 (1999).
549: % ``$\mathcal{PT}$-Symmetric Quantum Mechanics''
550:
551: \bibitem{r3} P.~Dorey, C.~Dunning, and R.~Tateo, J.~Phys.~A:
552: Math.~Gen.~{\bf 34}, L391 (2001) and {\bf 34}, 5679 (2001).
553:
554: \bibitem{r4} M.~Znojil, Phys.~Lett.~A {\bf 285}, 7 (2001).
555:
556: \bibitem{r5} M.~Znojil and G.~L\'evai, Mod.~Phys.~Lett.~A {\bf 16}, 2273 (2001).
557:
558: \bibitem{r6}
559: B.~Bagchi, S.~Mallik, and C.~Quesne, Mod.~Phys.~Lett.~A {\bf17}, 1651 (2002).
560:
561: \bibitem{r7}
562: A.~Mostafazadeh and A.~Batal, J.~Phys.~A: Math.~Gen.~{\bf 37}, 11645 (2004).
563:
564: \bibitem{r8} M.~Znojil, J.~Math.~Phys.~{\bf 46}, 062109 (2005) and arXiv:
565: quant-ph/0511085.
566:
567: \bibitem{r9} C.~M.~Bender, D.~C.~Brody, and H.~F.~Jones, Phys.~Rev.~Lett.~{\bf
568: 89}, 270401 (2002).
569:
570: \bibitem{r10} C.~M.~Bender, D.~C.~Brody, and H.~F.~Jones, Phys.~Rev.~Lett.~{\bf
571: 93}, 251601 (2004) and Phys.~Rev.~D {\bf 70}, 025001 (2004).
572:
573: \bibitem{r11} C.~M.~Bender, P.~N.~Meisinger, and Q.~Wang, J.~Phys.~A:
574: Math.~Gen.~{\bf 36}, 1973 (2003).
575:
576: \bibitem{r12} C.~M.~Bender, J.~Brod, A.~Refig, and M.~E.~Reuter,
577: J.~Phys.~A: Math.~Gen.~{\bf 37}, 10139 (2004).
578:
579: \bibitem{r13} A.~Mostafazadeh, J.~Math.~Phys.~{\bf 33}, 205 (2002) and
580: J.~Phys.~A: Math.~Gen.~{\bf 36}, 7081 (2003).
581:
582: \end{thebibliography}
583: \end{document}
584:
585: