1: \documentclass[12pt]{iopart}
2:
3: %\usepackage{graphicx}
4: %\usepackage{verbatim}
5: \usepackage{iopams}
6: \usepackage{amsthm}
7: \usepackage{epsf}
8:
9:
10: \eqnobysec
11:
12: %%%%%%%%%%%%%%%%%%% DEFINITIONS %%%%%%%%%%%%%%%%%%%%%%%%%%%%
13:
14: \newcommand{\Sw}{{\cal S}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
15: \newcommand{\Swt}{{\cal S}^{\times}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
16: \newcommand{\Swp}{{\cal S}^{\prime}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
17: \newcommand{\rhsSwt}{\Sw \subset L^2 \subset \Swt}
18: \newcommand{\rhsSwp}{\Sw \subset L^2 \subset \Swp}
19: \newcommand{\Czi}{C_0^{\infty}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
20: \newcommand{\Swhpm}{\widehat{{\cal S}}_{\pm}(\mathbb R \frac{\ }{\ }\{ a,b \})}
21: \newcommand{\Swhpmt}{\widehat{{\cal S}}_{\pm}^{\times}(\mathbb R \frac{\ }{\ }\{ a,b \})}
22: \newcommand{\Swhpml}{\widehat{{\cal S}}_{\pm;l}(\mathbb R \frac{\ }{\ }\{ a,b \})}
23: \newcommand{\Swhpmr}{\widehat{{\cal S}}_{\pm;r}(\mathbb R \frac{\ }{\ }\{ a,b \})}
24: \newcommand{\Swhpmlt}{\widehat{{\cal S}}_{\pm;l}^{\times}(\mathbb R \frac{\ }{\ }\{ a,b \})}
25: \newcommand{\Swhpmrt}{\widehat{{\cal S}}_{\pm;r}^{\times}(\mathbb R \frac{\ }{\ }\{ a,b \})}
26: \newcommand{\Swhpmp}{\widehat{{\cal S}}_{\pm}^{\prime}(\mathbb R \frac{\ }{\ }\{ a,b \})}
27: \newcommand{\Swhpmlp}{\widehat{{\cal S}}_{\pm;l}^{\prime}(\mathbb R \frac{\ }{\ }\{ a,b \})}
28: \newcommand{\Swhpmrp}{\widehat{{\cal S}}_{\pm;r}^{\prime}(\mathbb R \frac{\ }{\ }\{ a,b \})}
29: \newcommand{\Swh}{ \widehat{{\cal S}}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
30: \newcommand{\Swht}{ \widehat{{\cal S}}^{\times}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
31: \newcommand{\Swhp}{\widehat{{\cal S}}^{\prime}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
32: \newcommand{\Swhhpm}{\widehat{\widehat{{\cal S}}\,}\! _{\pm}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
33: \newcommand{\Swhhpmt}{\widehat{\widehat{{\cal S}}\,}\! _{\pm} \, \hskip-.32cm^{\times}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
34: \newcommand{\Swhhpmp}{\widehat{\widehat{{\cal S}}\,}\! _{\pm} \,\hskip-.22cm ^{\prime}(\mathbb R \frac{\ }{\ } \{ a,b \} )}
35:
36:
37: %%%%%%%%%%%%%%%% END OF DEFINITIONS %%%%%%%%%%%%%%%%
38:
39:
40:
41:
42:
43:
44:
45:
46: \begin{document}
47: %\bibliographystyle{revtex}
48:
49: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ARROWS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50: \def\llra{\relbar\joinrel\longrightarrow} %THIS IS LONG
51: \def\mapright#1{\smash{\mathop{\llra}\limits_{#1}}} %ARROW ON LINE
52: \def\mapup#1{\smash{\mathop{\llra}\limits^{#1}}} %CAN PUT SOMETHING OVER IT
53: \def\mapupdown#1#2{\smash{\mathop{\llra}\limits^{#1}_{#2}}} %over&under it%
54: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% END ARROWS %%%%%%%%%%%%%%%%%%%%%%%%%%%%
55:
56:
57:
58: %%%%%%%%%%%%%% These are the AMS constructs for multiline limits %%%%%%%%%%%%%
59: %
60: \catcode`\@=11
61:
62: \def\BF#1{{\bf {#1}}}
63: \def\NEG#1{{\rlap/#1}}
64:
65: \def\Let@{\relax\iffalse{\fi\let\\=\cr\iffalse}\fi}
66: \def\vspace@{\def\vspace##1{\crcr\noalign{\vskip##1\relax}}}
67: \def\multilimits@{\bgroup\vspace@\Let@
68: \baselineskip\fontdimen10 \scriptfont\tw@
69: \advance\baselineskip\fontdimen12 \scriptfont\tw@
70: \lineskip\thr@@\fontdimen8 \scriptfont\thr@@
71: \lineskiplimit\lineskip
72: \vbox\bgroup\ialign\bgroup\hfil$\m@th\scriptstyle{##}$\hfil\crcr}
73: \def\Sb{_\multilimits@}
74: \def\endSb{\crcr\egroup\egroup\egroup}
75: \def\Sp{^\multilimits@}
76: \let\endSp\endSb
77: %
78: %%%%%%%%%%%%%%%%%%%%END of explanations for multiline limits %%%%%%%%%%%%%%%%%
79:
80: \title[The RHS in Quantum Mechanics]{The role of the rigged Hilbert space in
81: Quantum Mechanics}
82:
83: \author{Rafael de la Madrid}
84: \address{Departamento de F\'\i sica Te\'orica, Facultad de Ciencias,
85: Universidad del Pa\'\i s Vasco, 48080 Bilbao, Spain \\
86: E-mail: {\texttt{wtbdemor@lg.ehu.es}}}
87:
88:
89: \date{\small{January 4, 2005}}
90:
91:
92: \begin{abstract}
93: There is compelling evidence that, when continuous spectrum is
94: present, the natural mathematical setting for Quantum Mechanics is the rigged
95: Hilbert space rather than just the Hilbert space. In particular, Dirac's
96: bra-ket formalism is fully implemented by the rigged Hilbert space rather than
97: just by the Hilbert space. In this paper, we provide a pedestrian introduction
98: to the role the rigged Hilbert space plays in Quantum Mechanics, by way of a
99: simple, exactly solvable example. The procedure will be constructive and based
100: on a recent publication. We also provide a thorough discussion
101: on the physical significance of the rigged Hilbert space.
102: \end{abstract}
103:
104: \pacs{03.65.-w, 02.30.Hq}
105:
106: \maketitle
107:
108:
109:
110:
111:
112:
113:
114: \section{Introduction}
115: \setcounter{equation}{0}
116: \label{sec:introduction}
117:
118:
119: It has been known for several decades that Dirac's bra-ket formalism is
120: mathematically justified not by the Hilbert space alone, but by the
121: rigged Hilbert space (RHS). This is the reason why there is an increasing
122: number of Quantum Mechanics textbooks that already include the
123: rigged Hilbert space as part of their contents (see, for example,
124: Refs.~\cite{ATKINSON}-\cite{KUKULIN}). Despite the importance of the RHS,
125: there is still a lack of
126: simple examples for which the corresponding RHS is constructed in a didactical
127: manner. Even worse, there is no pedagogical
128: discussion on the physical significance of the RHS. In this paper, we
129: use the one-dimensional (1D) rectangular barrier potential to introduce the RHS
130: at the graduate student level. As well, we discuss the physical significance
131: of each of the ingredients that form the RHS. The construction of the RHS of
132: such a simple model will unambiguously show that the RHS is needed at the
133: most basic level of Quantum Mechanics.
134:
135: The present paper is complemented by a previous publication,
136: Ref.~\cite{FOCO}, to which we shall refer the reader interested in a
137: detailed mathematical account on the construction of the RHS of the 1D
138: rectangular barrier. For a general background on the Hilbert and the rigged
139: Hilbert space methods, the reader may consult Ref.~\cite{DIS} and
140: references therein.
141:
142: Dirac's bra-ket formalism was introduced by Dirac in his classic
143: monograph~\cite{DIRAC}. Since its inception, Dirac's abstract algebraic model
144: of {\it bras} and {\it kets} (from the bracket notation for the inner
145: product) proved to be of great calculational value, although there were
146: serious difficulties in finding a mathematical justification for the actual
147: calculations within the Hilbert space, as Dirac~\cite{DIRAC} and von
148: Neumann~\cite{VON} themselves state in their books~\cite{QUOTEVONDIRAC}. As
149: part of his bra-ket formalism, Dirac introduced the so-called
150: Dirac delta function, a formal entity without a counterpart in the classical
151: theory of functions. It was L.~Schwartz who gave a precise meaning to the
152: Dirac delta function as a functional over a space of test
153: functions~\cite{SCHWARTZ}. This led to the development of a new branch of
154: functional analysis, the theory of distributions. By combining von Neumann's
155: Hilbert space with the theory of distributions, I.~Gelfand and collaborators
156: introduced the RHS~\cite{GELFAND,MAURIN}. It was already clear to the creators
157: of the RHS that their formulation was the mathematical support of Dirac's
158: bra-ket formalism~\cite{CITEMAURIN}. The RHS made its first appearance in the
159: Physics literature in the 1960s~\cite{ROBERTS,ANTOINE,B60}, when some
160: physicists also realized that the RHS provides a rigorous mathematical
161: rephrasing of all of the aspects of Dirac's bra-ket formalism. Nowadays,
162: there is a growing consensus that the RHS, rather than the Hilbert space
163: alone, is the natural mathematical setting of Quantum
164: Mechanics~\cite{QUOTEBALLENTINE}.
165:
166: A note on semantics. The word ``rigged'' in rigged Hilbert space has a
167: nautical connotation, such as the phrase ``fully rigged ship;'' it
168: has nothing to do with any unsavory practice such as ``fixing'' or
169: predetermining a result. The phrase ``rigged Hilbert space'' is a direct
170: translation of the phrase ``osnashchyonnoe Hilbertovo prostranstvo'' from
171: the original Russian. A more faithful translation would be
172: ``equipped Hilbert space.'' Indeed, the rigged Hilbert space is just the
173: Hilbert space equipped with distribution theory---in Quantum Mechanics, to
174: rig a Hilbert space means simply to equip that Hilbert space with distribution
175: theory. Thus, the RHS is not a replacement but an enlargement of the Hilbert
176: space.
177:
178: The RHS is {\it neither} an extension {\it nor} an
179: interpretation of the physical principles of Quantum Mechanics, but rather
180: the most natural, concise and logic language to formulate Quantum
181: Mechanics. The RHS is simply a mathematical tool to extract
182: and process the information contained in observables
183: that have continuous spectrum. Observables with discrete spectrum and a
184: finite number of eigenvectors (e.g., spin) do not need the RHS. For such
185: observables, the Hilbert space is sufficient. Actually, as we shall explain,
186: in general only unbounded observables with continuous spectrum need the RHS.
187:
188: The usefulness of the RHS is not simply restricted to accounting for Dirac's
189: bra-ket formalism. The RHS has also proved to be a very useful
190: research tool in the quantum theory of scattering and decay
191: (see Ref.~\cite{DIS} and references therein), and in the construction of
192: generalized spectral decompositions of chaotic maps~\cite{AT93,SUCHANECKI}. In
193: fact, it seems that the RHS is the natural language to deal with problems
194: that involve continuous and resonance spectra.
195:
196: Loosely speaking, a rigged Hilbert space (also called a Gelfand triplet) is
197: a triad of spaces
198: \begin{equation}
199: {\mathbf \Phi} \subset {\cal H} \subset {\mathbf \Phi}^{\times}
200: \label{RHStIntro}
201: \end{equation}
202: such that $\cal H$ is a Hilbert space, $\mathbf \Phi$ is a dense
203: subspace of $\cal H$~\cite{DENSE}, and $\mathbf \Phi ^{\times}$ is the space of
204: antilinear functionals over $\mathbf \Phi$~\cite{FUNCTIONAL}. Mathematically,
205: $\mathbf \Phi$ is the space of test functions, and $\mathbf \Phi ^{\times}$
206: is the space of distributions. The space $\mathbf \Phi ^{\times}$ is called
207: the antidual space of $\mathbf \Phi$. Associated with the
208: RHS~(\ref{RHStIntro}), there is always another RHS,
209: \begin{equation}
210: {\mathbf \Phi} \subset {\cal H} \subset {\mathbf \Phi}^{\prime} \, ,
211: \label{RHSpIntro}
212: \end{equation}
213: where ${\mathbf \Phi}^{\prime}$ is called the dual space of ${\mathbf \Phi}$
214: and contains the linear functionals over $\mathbf \Phi$~\cite{FUNCTIONAL}.
215:
216: The basic reason why we need the spaces ${\mathbf \Phi}^{\prime}$ and
217: ${\mathbf \Phi}^{\times}$ is that the bras and kets associated with the
218: elements in the continuous spectrum of an observable belong, respectively, to
219: ${\mathbf \Phi}^{\prime}$ and ${\mathbf \Phi}^{\times}$ rather than to
220: ${\cal H}$. The basic reason reason why we need the space $\mathbf \Phi$ is
221: that unbounded operators are not defined on the whole of ${\cal H}$ but only
222: on dense subdomains of ${\cal H}$ that are not invariant under the
223: action of the observables. Such non-invariance makes expectation values,
224: uncertainties and commutation relations not well defined on the whole
225: of $\cal H$. The space $\mathbf \Phi$ is the largest subspace of the Hilbert
226: space on which such expectation values, uncertainties and commutation
227: relations are well defined.
228:
229: The original formulation of the RHS~\cite{GELFAND,MAURIN} does not provide a
230: systematic procedure to construct the RHS generated by the Hamiltonian of the
231: Schr\"odinger equation, since the space $\mathbf \Phi$ is assumed to be
232: given beforehand. Such systematic procedure is important because,
233: after all, claiming that the RHS is the natural setting for
234: Quantum Mechanics is about the same as claiming that, when the Hamiltonian
235: has continuous spectrum, the natural setting for the solutions of the
236: Schr\"odinger equation is the RHS rather than just the Hilbert space. The
237: task of developing a systematic procedure to construct the RHS generated
238: by the Schr\"odinger equation was undertaken in Ref.~\cite{DIS}. The
239: method proposed in Ref.~\cite{DIS}, which was partly based on
240: Refs.~\cite{ROBERTS,ANTOINE,B60}, has been applied to two simple
241: three-dimensional potentials, see Refs.~\cite{JPA02,FP02}, to the
242: three-dimensional free Hamiltonian, see Ref.~\cite{IJTP03}, and to
243: the 1D rectangular barrier potential, see Ref.~\cite{FOCO}. In this paper, we
244: present the method of Ref.~\cite{DIS} in a didactical manner.
245:
246: The organization of the paper is as follows. In Sec.~\ref{sec:why}, we outline
247: the major reasons why the RHS provides the mathematical setting for Quantum
248: Mechanics. In Sec.~\ref{sec:e1dsbp}, we recall the basics of the
249: 1D rectangular potential model. Section~\ref{sec:crhs} provides the RHS of
250: this model. In Sec.~\ref{sec:phymean}, we discuss the physical meaning of
251: each of the ingredients that form the RHS. In Sec.~\ref{sec:gener}, we
252: discuss the relation of the Hilbert space spectral measures with the bras
253: and kets, as well as the limitations of our method to construct
254: RHSs. Finally, Sec.~\ref{sec:conclusions} contains the conclusions to the
255: paper.
256:
257:
258:
259:
260:
261: \section{Motivating the rigged Hilbert space}
262: %\setcounter{equation}{0}
263: \label{sec:why}
264:
265: The {\it linear superposition principle} and the
266: {\it probabilistic interpretation} of Quantum Mechanics are two
267: major guiding principles in our
268: understanding of the microscopic world. These two principles suggest that the
269: space of states be a linear space (which accounts for the superposition
270: principle) endowed with a scalar product (which is used to calculate
271: probability amplitudes). A linear space endowed with a scalar product is
272: called a Hilbert space and is usually denoted by $\cal H$~\cite{HSDEF}.
273:
274: In Quantum Mechanics, observable quantities are represented
275: by linear, self-adjoint operators acting on $\cal H$. The eigenvalues
276: of an operator represent the possible values of the
277: measurement of the corresponding observable. These eigenvalues, which
278: mathematically correspond to the spectrum of the operator, can be discrete
279: (as the energies of a particle in a box), continuous (as the energies
280: of a free, unconstrained particle), or a combination of
281: discrete and continuous (as the energies of the Hydrogen atom).
282:
283: When the spectrum of an observable $A$ is discrete and $A$ is
284: bounded~\cite{UNB}, then $A$ is defined on the whole of $\cal H$ and
285: the eigenvectors of $A$ belong to $\cal H$. In this case, $A$ can be
286: essentially seen as a matrix. This means that, as
287: far as discrete spectrum is concerned, there is no need to extend
288: $\cal H$. However, quantum mechanical
289: observables are in general unbounded~\cite{UNB} and their spectrum has
290: in general a continuous part. In order to deal with continuous
291: spectrum, textbooks usually follow Dirac's bra-ket
292: formalism, which is a heuristic generalization of the linear algebra of
293: Hermitian matrices used for discrete spectrum. As we shall see, the
294: mathematical methods of the Hilbert space are not sufficient to make sense of
295: the prescriptions of Dirac's formalism, the reason for which we shall
296: extend the Hilbert space to the rigged Hilbert space.
297:
298: For pedagogical reasons, we recall the essentials of the linear algebra of
299: Hermitian matrices before proceeding with Dirac's formalism.
300:
301:
302:
303:
304:
305: \subsection{Hermitian matrices}
306:
307:
308: If the measurement of an observable $A$ (e.g., spin) yields a discrete,
309: finite number $N$ of results $a_n$, $n=1, 2, \ldots , N$, then $A$ is realized
310: by a Hermitian matrix on a Hilbert space $\cal H$ of dimension $N$. Since
311: $\cal H$ is an $N$-dimensional linear
312: space, there are $N$ linearly independent vectors $\{ e_n \} _{n=1}^N$ that
313: form an orthonormal basis system for $\cal H$. We denote these basis vectors
314: $e_n$ also by $|e_n\rangle$. The scalar products of the elements of the basis
315: system are written in one of the following ways:
316: \begin{equation}
317: e_n \cdot e_m \equiv (e_n,e_m)\equiv \langle e_n|e_m\rangle =
318: \delta_{nm} \, , \qquad n,m=1,2,\ldots , N \, ,
319: \label{dcnII}
320: \end{equation}
321: where $\delta_{nm}$ is the Kronecker delta. As the basis system for the space
322: $\cal H$, it is always possible to choose the eigenvectors of
323: $A$. Therefore, one can choose basis vectors $e_n\in {\cal H}$ which also
324: fulfill
325: \begin{equation}
326: Ae_n=a_n e_n \, .
327: \end{equation}
328: Since $A$ is Hermitian, the eigenvalues $a_n$ are real. The eigenvectors $e_n$
329: are often labeled by their eigenvalues $a_n$ and denoted by
330: \begin{equation}
331: e_n\equiv |a_n\rangle \, ,
332: \end{equation}
333: and they are represented by column vectors. For each column eigenvector
334: $e_n\equiv |a_n\rangle$, there also exists a row eigenvector
335: $\tilde{e}_n\equiv \langle a_n|$ that is a left eigenvector of $A$,
336: \begin{equation}
337: \tilde{e}_n A = a_n \tilde{e}_n \, .
338: \end{equation}
339: Thus, when $A$ is a Hermitian matrix acting on an $N$-dimensional
340: Hilbert space $\mathcal H$, for each eigenvalue $a_n$ of $A$ there exist
341: a right (i.e., column) eigenvector of $A$
342: \begin{equation}
343: A|a_n\rangle =a_n|a_n \rangle \, , \quad n=1,2,\ldots,N \, ,
344: \label{fddeII}
345: \end{equation}
346: and also a left (i.e., row) eigenvector of $A$
347: \begin{equation}
348: \langle a_n|A =a_n \langle a_n| \, , \quad n=1,2,\ldots,N \, ,
349: \label{fddeIIl}
350: \end{equation}
351: such that these row and column eigenvectors are orthonormal,
352: \begin{equation}
353: \langle a_n|a_m \rangle =\delta _{nm} \, , \quad n,m=1,2,\ldots ,N \, ,
354: \label{orthonomr}
355: \end{equation}
356: and such that every vector $\varphi \in {\cal H}$ can be written as
357: \begin{equation}
358: \varphi=\sum^N_{n=1}|a_n\rangle \langle a_n|\varphi \rangle \, .
359: \label{fddsdII}
360: \end{equation}
361: Equation~(\ref{fddsdII}) is called the eigenvector expansion
362: of $\varphi$ with respect to the eigenvectors of $A$. The complex numbers
363: $\langle a_n|\varphi \rangle$ are the components of the vector $\varphi$
364: with respect to the basis of eigenvectors of $A$. Physically,
365: $\langle a_n|\varphi \rangle$ represents the probability amplitude of
366: obtaining the value $a_n$ in the measurement of the observable $A$ on the
367: state $\varphi$. By acting on both sides of Eq.~(\ref{fddsdII}) with $A$,
368: and recalling Eq.~(\ref{fddeII}), we obtain that
369: \begin{equation}
370: A \varphi=\sum^N_{n=1}a_n |a_n\rangle \langle a_n|\varphi \rangle \, .
371: \label{fddsdIIA}
372: \end{equation}
373:
374:
375:
376:
377:
378:
379: \subsection{Dirac's bra-ket formalism}
380:
381:
382: Dirac's formalism is an elegant, heuristic generalization of the algebra of
383: finite dimensional matrices to the continuous-spectrum, infinite-dimensional
384: case. Four of the most important features of Dirac's formalism are:
385:
386: \begin{enumerate}
387: \item To each element of the spectrum of an observable $A$,
388: there correspond a left and a right eigenvector (for the moment, we assume
389: that the spectrum is non-degenerate). If discrete eigenvalues are
390: denoted by $a_n$ and continuous eigenvalues by $a$, then the
391: corresponding right eigenvectors, which are denoted by the kets
392: $|a_n\rangle$ and $|a \rangle$, satisfy
393: \numparts
394: \begin{equation}
395: A|a_n \rangle =a_n|a_n\rangle \, ,
396: \label{dketqueintro}
397: \end{equation}
398: \begin{equation}
399: A|a \rangle =a |a \rangle \, ,
400: \label{cketequeintro}
401: \end{equation}
402: \endnumparts
403: and the corresponding left eigenvectors, which are denoted
404: by the bras $\langle a_n|$ and $\langle a|$, satisfy
405: \numparts
406: \begin{equation}
407: \langle a_n|A=a_n \langle a_n| \, ,
408: \end{equation}
409: \begin{equation}
410: \langle a|A=a \langle a| \, .
411: \label{braeqeneintro}
412: \end{equation}
413: \endnumparts
414: The bras $\langle a|$ generalize the notion of row eigenvectors, whereas the
415: kets $|a \rangle$ generalize the notion of column eigenvectors.
416:
417: \item In analogy to Eq.~(\ref{fddsdII}), the eigenbras and eigenkets of
418: an observable form a complete basis, that is, any wave function $\varphi$ can
419: be expanded in the so-called Dirac basis expansion:
420: \begin{equation}
421: \varphi = \sum_n |a_n\rangle \langle a_n|\varphi \rangle +
422: \int \rmd a \, |a \rangle \langle a |\varphi \rangle \, .
423: \label{introDirbaexp}
424: \end{equation}
425: In addition, the bras and kets furnish a resolution of the identity,
426: \begin{equation}
427: I = \sum_n |a_n\rangle \langle a_n| +
428: \int \rmd a \, |a \rangle \langle a | \, ,
429: \label{introresiden}
430: \end{equation}
431: and, in a generalization of Eq.~(\ref{fddsdIIA}), the action of $A$ can be
432: written as
433: \begin{equation}
434: A = \sum_n a_n |a_n\rangle \langle a_n| +
435: \int \rmd a \, a |a \rangle \langle a | \, .
436: \label{introactionA}
437: \end{equation}
438:
439: \item The bras and kets are normalized according to the following rule:
440: \numparts
441: \begin{equation}
442: \langle a_n|a_m\rangle =\delta _{nm} \, ,
443: \end{equation}
444: \begin{equation}
445: \langle a |a ^{\prime} \rangle = \delta (a -a ^{\prime}) \, ,
446: \label{deltanorintro}
447: \end{equation}
448: \endnumparts
449: where $\delta _{nm}$ is the Kronecker delta and $\delta (a-a ^{\prime})$ is
450: the Dirac delta. The Dirac delta normalization generalizes the
451: orthonormality~(\ref{orthonomr}) of the eigenvectors of a Hermitian matrix.
452: \item Like in the case of two finite-dimensional matrices,
453: all algebraic operations such as the commutator of two observables $A$ and $B$,
454: \begin{equation}
455: [A,B]=AB-BA \, ,
456: \label{comuts}
457: \end{equation}
458: are always well defined.
459: \end{enumerate}
460:
461:
462:
463:
464:
465:
466:
467: \subsection{The need of the rigged Hilbert space}
468:
469:
470: In Quantum Mechanics, observables are usually given by differential
471: operators. In the Hilbert space framework, the formal prescription of an
472: observable leads to the definition of a linear operator as follows: One has
473: to find
474: first the Hilbert space $\cal H$, then one sees on what elements of $\cal H$
475: the action of the observable makes sense, and finally one checks whether the
476: action of the observable remains in $\cal H$. For example, the position
477: observable $Q$ of a 1D particle is given by
478: \begin{equation}
479: Qf(x)=xf(x) \, .
480: \label{fdopx}
481: \end{equation}
482: The Hilbert space of a 1D particle is given by the collection of square
483: integrable functions,
484: \begin{equation}
485: L^2 =\{ f(x) \, | \ \int_{-\infty}^{\infty}\rmd x \,
486: |f(x)|^2 < \infty \} \, ,
487: \label{l2space}
488: \end{equation}
489: and the action of $Q$, although in principle well defined on every
490: element of $L^2$, remains in $L^2$
491: only for the elements of the following subspace:
492: \begin{equation}
493: {\cal D}(Q)= \{ f(x) \in L^2 \, | \
494: \int_{-\infty}^{\infty}\rmd x \,
495: |xf(x)|^2 < \infty \} \, .
496: \label{domainQ}
497: \end{equation}
498: The space ${\cal D}(Q)$ is the domain of the position operator. Domain
499: (\ref{domainQ}) is not the whole of $L^2$, since
500: the function $g(x)=1/(x+\rmi)$ belongs to $L^2$ but not to
501: ${\cal D}(Q)$; as well, $Q$ is an unbounded operator, because
502: $\| Qg \| = \infty$; as well, $Q{\cal D}(Q)$ is not included in
503: ${\cal D}(Q)$, since $h(x)=1/(x^2+1)$ belongs to ${\cal D}(Q)$ but $Qh$ does
504: not belong to ${\cal D}(Q)$. The denseness and the non-invariance of the
505: domains of unbounded operators create much trouble in the Hilbert space
506: framework, because one has always to be careful whether formal operations
507: are valid. For example, $Q^2=QQ$ is not
508: defined on the whole of $L^2$, not even on the whole
509: of ${\cal D}(Q)$, but only on those square integrable functions such that
510: $x^2f \in L^2$. Also, the expectation value of the measurement of $Q$ in
511: the state $\varphi$,
512: \begin{equation}
513: (\varphi ,Q\varphi ) \, ,
514: \label{exintrodispQ}
515: \end{equation}
516: is not finite for every $\varphi \in L^2$, but only when
517: $\varphi \in {\cal D}(Q)$. Similarly, the uncertainty of the measurement
518: of $Q$ in $\varphi$,
519: \begin{equation}
520: \Delta _{\varphi}Q=
521: \sqrt{ (\varphi ,Q^2\varphi )-(\varphi ,Q\varphi )^2} \, ,
522: \label{introdispQ}
523: \end{equation}
524: is not defined on the whole of $L^2$.
525:
526: On the other hand, if we denote the momentum observable by
527: \begin{equation}
528: Pf(x)=-\rmi \hbar \frac{\rmd}{\rmd x}f(x) \, ,
529: \label{fdopp}
530: \end{equation}
531: then the product of $P$ and $Q$, $PQ$, is not defined everywhere in the
532: Hilbert space, but only on those square integrable functions for which the
533: quantity
534: \begin{equation}
535: PQf(x) = -\rmi \hbar \frac{\rmd}{\rmd x}xf(x)=
536: -\rmi \hbar \left( f(x)+ xf'(x) \right)
537: \label{pqf}
538: \end{equation}
539: makes sense and is square integrable. Obviously, $PQf$ makes
540: sense only when $f$ is differentiable, and $PQf$ remains in
541: $L^2$ only when $f$, $f'$ and $xf'$ are also in
542: $L^2$; thus, $PQ$ is not defined everywhere in
543: $L^2$ but only on those square integrable functions that
544: satisfy the aforementioned conditions. Similar domain concerns arise
545: in calculating the commutator of $P$ with $Q$.
546:
547: As in the case of the position operator, the domain ${\cal D}(A)$ of an
548: unbounded operator $A$ does not coincide with the whole of
549: $\cal H$~\cite{RS84}, but is just a dense subspace of
550: $\cal H$~\cite{DENSE}; also, in general ${\cal D}(A)$ does not remain
551: invariant under the action of $A$, that is, $A{\cal D}(A)$ is not included
552: in ${\cal D}(A)$. Such non-invariance makes expectation values,
553: \begin{equation}
554: (\varphi ,A\varphi ) \, ,
555: \label{exintrodispP}
556: \end{equation}
557: uncertainties,
558: \begin{equation}
559: \Delta _{\varphi}A=
560: \sqrt{ (\varphi ,A^2\varphi )-(\varphi ,A\varphi )^2} \, ,
561: \label{introdispA}
562: \end{equation}
563: and algebraic operations such as commutation relations not well defined on
564: the whole of the Hilbert space $\cal H$~\cite{INFENER}. Thus, when the
565: position, momentum and energy operators $Q$, $P$, $H$ are unbounded, it is
566: natural to seek a subspace $\mathbf \Phi$ of $\cal H$ on which all of these
567: physical quantities can be calculated and yield meaningful, finite
568: values. Because the reason why these quantities may not be well defined is
569: that the domains of $Q$, $P$ and $H$
570: are not invariant under the action of these operators, the subspace
571: $\mathbf \Phi$ must be such that
572: it remains invariant under the actions of $Q$, $P$ and $H$. This is why we
573: take as $\mathbf \Phi$ the intersection of the domains of all the powers
574: of $Q$, $P$ and $H$~\cite{ROBERTS}:
575: \begin{equation}
576: {\mathbf \Phi} =\bigcap _{\Sb n,m=0 \\ A,B=Q,P,H \endSb}^{\infty}
577: {\cal D}(A^nB^m) \, .
578: \label{maximalinvas}
579: \end{equation}
580: This space is known as the maximal invariant subspace of the algebra generated
581: by $Q$, $P$ and $H$, because it is the largest subdomain of the Hilbert space
582: that remains invariant under the action of any power of $Q$, $P$ or $H$,
583: \begin{equation}
584: A {\mathbf \Phi} \subset {\mathbf \Phi} \, , \qquad A=Q,P,H \, .
585: \end{equation}
586: On $\mathbf \Phi$, all physical quantities such as
587: expectation values and uncertainties can be associated well-defined, finite
588: values, and algebraic operations such as the commutation relation
589: (\ref{comuts}) are well defined. In addition, the elements of
590: $\mathbf \Phi$ are represented by smooth, continuous functions that
591: have a definitive value at each point, in contrast
592: to the elements of $\cal H$, which are represented by classes of functions
593: which can vary arbitrarily on sets of zero Lebesgue measure.
594:
595: Not only there are compelling reasons to shrink the Hilbert space $\cal H$ to
596: $\mathbf \Phi$, but, as we are going to explain now, there are also reasons
597: to enlarge $\cal H$ to the spaces ${\mathbf \Phi}^{\times}$ and
598: ${\mathbf \Phi}^{\prime}$ of Eqs.~(\ref{RHStIntro}) and
599: (\ref{RHSpIntro}). When the spectrum of $A$ has a continuous
600: part, prescriptions~(\ref{braeqeneintro}) and (\ref{cketequeintro}) associate a
601: bra $\langle a|$ and a ket $|a\rangle$ to each element $a$ of the
602: continuous spectrum of $A$. Obviously, the bras $\langle a|$ and kets
603: $|a\rangle$ are not in the Hilbert space~\cite{SNHS}, and therefore we
604: need two linear spaces larger than the Hilbert space to accommodate
605: them. It turns
606: out that the bras and kets acquire mathematical meaning as distributions. More
607: specifically, the bras $\langle a|$ are {\it linear} functionals over
608: the space $\mathbf \Phi$, and the kets $|a\rangle$ are
609: {\it antilinear} functionals over the space
610: $\mathbf \Phi$. That is, $\langle a| \in {\mathbf \Phi}^{\prime}$ and
611: $|a \rangle \in {\mathbf \Phi}^{\times}$.
612:
613: In this way, the Gelfand triplets of Eqs.~(\ref{RHStIntro}) and
614: (\ref{RHSpIntro}) arise in a natural way. The Hilbert space $\mathcal H$
615: arises from the requirement that the wave functions be square
616: normalizable. Aside from providing mathematical concepts such as
617: self-adjointness or unitarity, the Hilbert space plays a very important
618: physical role, namely $\cal H$ selects the scalar product that is used to
619: calculate probability amplitudes. The subspace $\mathbf \Phi$ contains those
620: square integrable functions that should be considered as physical, because
621: any expectation value, any uncertainty and any algebraic operation can be
622: calculated for its elements, whereas this is not possible for the rest of the
623: elements of the Hilbert space. The dual space ${\mathbf \Phi}^{\prime}$ and
624: the antidual space ${\mathbf \Phi}^{\times}$ contain respectively the bras
625: and the kets associated with the continuous spectrum of the observables. These
626: bras and kets can be used to expand any $\varphi \in {\mathbf \Phi}$ as in
627: Eq.~(\ref{introDirbaexp}). Thus, the rigged Hilbert space, rather than
628: the Hilbert space alone, can accommodate prescriptions
629: (\ref{dketqueintro})-(\ref{comuts}) of Dirac's formalism.
630:
631: It should be clear that the rigged Hilbert space is just a combination of
632: the Hilbert space with distribution theory. This combination enables us to
633: deal with singular objects such as bras, kets, or Dirac's delta function,
634: something that is impossible if we only use the Hilbert space.
635:
636: Even though it is apparent that the rigged Hilbert space should be an
637: essential part of the mathematical methods for Quantum Mechanics, one may
638: still wonder if the rigged Hilbert space is a helpful tool in teaching
639: Quantum Mechanics, or rather is a technical nuance. Because basic quantum
640: mechanical operators such as $P$ and $Q$ are in general unbounded operators
641: with continuous spectrum~\cite{RS274}, and because this kind of operators
642: necessitates the rigged Hilbert space, it seems pertinent to introduce the
643: rigged Hilbert space in graduate courses on Quantum Mechanics.
644:
645: From a pedagogical standpoint, however, this section's introduction to the
646: rigged Hilbert space is not sufficient. In the classroom, new concepts are
647: better introduced by way of a simple, exactly solvable example. This is why
648: we shall construct the RHS of the 1D rectangular barrier system. We note that
649: this system does not have bound states, and therefore in what follows
650: we shall not deal with discrete spectrum.
651:
652:
653:
654:
655:
656:
657:
658:
659:
660: \subsection{Representations}
661:
662:
663:
664: In working out specific examples, the prescriptions of Dirac's
665: formalism have to be written in a particular representation. Thus, before
666: constructing the RHS of the 1D rectangular barrier, it is convenient to
667: recall some of the basics of representations.
668:
669: In Quantum Mechanics, the most common of all representations is the position
670: representation, sometimes called the $x$-representation. In the
671: $x$-representation, the position operator $Q$ acts as multiplication by
672: $x$. Since the spectrum of $Q$ is $(-\infty ,\infty )$, the $x$-representation
673: of the Hilbert space $\cal H$ is given by the space
674: $L^2$. In this paper, we shall mainly work in the
675: position representation.
676:
677: In general, given an observable $B$, the $b$-representation is that in which
678: the operator $B$ acts as multiplication by $b$, where the $b$'s denote
679: the eigenvalues of $B$. If we denote the spectrum of $B$ by ${\rm Sp}(B)$, then
680: the $b$-representation of the Hilbert space $\cal H$ is given by the
681: space $L^2( {\rm Sp}(B),\rmd b)$, which is the space of square integrable
682: functions $f(b)$ with $b$ running over ${\rm Sp}(B)$. In the
683: $b$-representation, the restrictions to purely continuous spectrum of
684: prescriptions~(\ref{dketqueintro})-(\ref{introresiden}) become
685: \numparts
686: \begin{equation}
687: \langle b|A|a \rangle =a \langle b|a \rangle \, ,
688: \label{bqueintro}
689: \end{equation}
690: \begin{equation}
691: \langle a|A|b\rangle =a \langle a|b\rangle \, ,
692: \label{bbraeqeneintro}
693: \end{equation}
694: \begin{equation}
695: \langle b| \varphi \rangle =
696: \int \rmd a \, \langle b|a \rangle \langle a |\varphi \rangle \, ,
697: \label{bintroDirbaexp}
698: \end{equation}
699: \begin{equation}
700: \delta (b-b')=\langle b|b'\rangle =
701: \int \rmd a \, \langle b|a \rangle \langle a |b'\rangle \, .
702: \label{bintroresiden}
703: \end{equation}
704: \endnumparts
705: The ``scalar product'' $\langle b|a\rangle$ is obtained from
706: Eq.~(\ref{bqueintro}) as the solution of a differential eigenequation in
707: the $b$-representation. The $\langle b|a\rangle$ can also be seen as
708: transition elements from the
709: $a$- to the $b$-representation. Mathematically, the $\langle b|a\rangle$
710: are to be treated as distributions, and therefore they often appear
711: as kernels of integrals. In this paper, we shall encounter a few of these
712: ``scalar products'' such as $\langle x| p\rangle$, $\langle x| x'\rangle$
713: and $\langle x| E^{\pm} \rangle _{\rm l,r}$.
714:
715:
716:
717:
718:
719:
720:
721: \section{Example: The one-dimensional rectangular barrier potential}
722: %\setcounter{equation}{0}
723: \label{sec:e1dsbp}
724:
725:
726: The example we consider in this paper is supposed to represent a spinless
727: particle moving in one dimension and impinging on a rectangular barrier. The
728: observables relevant to this system are the position $Q$, the momentum $P$, and
729: the Hamiltonian $H$. In the position representation, $Q$ and $P$ are
730: respectively realized by the differential operators (\ref{fdopx}) and
731: (\ref{fdopp}), whereas $H$ is realized by
732: \begin{equation}
733: Hf(x)= \left( -\frac{\hbar ^2}{2m}\frac{\rmd ^2}{\rmd x^2}+V(x) \right)
734: f(x)
735: \, ,
736: \label{fdoph}
737: \end{equation}
738: where
739: \begin{equation}
740: V(x)=\left\{ \begin{array}{ll}
741: 0 &-\infty <x<a \\
742: V_0 &a<x<b \\
743: 0 &b<x<\infty
744: \end{array}
745: \right.
746: \label{sbpotential}
747: \end{equation}
748: is the 1D rectangular barrier potential. Formally, these observables satisfy
749: the following commutation relations:
750: \numparts
751: \begin{equation}
752: \left[ Q,P \right] =\rmi \hbar I \, , \label{cr1}
753: \end{equation}
754: \begin{equation}
755: \left[ H,Q \right] =- \frac{\rmi \hbar}{m} P \, ,
756: \end{equation}
757: \begin{equation}
758: \left[ H,P \right] = \rmi \hbar \frac{\partial V}{\partial x} \, .
759: \label{cr3}
760: \end{equation}
761: \endnumparts
762:
763: Since our particle can move in the full real line, the Hilbert space
764: on which the differential operators~(\ref{fdopx}), (\ref{fdopp}) and
765: (\ref{fdoph}) should act is
766: $L^2$ of Eq.~(\ref{l2space}). The corresponding scalar
767: product is
768: \begin{equation}
769: (f,g)=\int_{-\infty}^{\infty}\rmd x \, \overline{f(x)}g(x) \, , \qquad
770: f,g \in L^2 \, ,
771: \label{scapro}
772: \end{equation}
773: where $\overline{f(x)}$ denotes the complex conjugate of $f(x)$.
774:
775: The differential operators~(\ref{fdopx}), (\ref{fdopp}) and (\ref{fdoph})
776: induce three linear
777: operators on the Hilbert space $L^2$. These operators are
778: unbounded~\cite{FOCO}, and therefore they cannot be defined on the whole of
779: $L^2$, but only on the following subdomains of
780: $L^2$~\cite{FOCO}:
781: \numparts
782: \begin{equation}
783: {\cal D}(Q)=\left\{ f\in L^2 \, | \
784: xf \in L^2 \right\} \, ,
785: \end{equation}
786: \begin{equation}
787: {\cal D}(P)=\left\{ f\in L^2 \, | \
788: f \in AC, \
789: Pf \in L^2 \right\} \, ,
790: \end{equation}
791: \begin{equation}
792: {\cal D}(H)=\left\{ f\in L^2 \, | \
793: f \in AC^2, \
794: Hf \in L^2 \right\} \, ,
795: \label{domainH}
796: \end{equation}
797: \endnumparts
798: where, essentially, $AC$ is the space of functions whose
799: derivative exists, and $AC^2$ is the space of functions whose
800: second derivative exists (see Ref.~\cite{FOCO} for more details). On
801: these domains, the operators $Q$, $P$ and $H$ are
802: self-adjoint~\cite{FOCO}.
803:
804: In our example, the eigenvalues (i.e., the spectrum) and the eigenfunctions
805: of the observables are provided by the Sturm-Liouville theory. Mathematically,
806: the eigenvalues and eigenfunctions of operators extend the notions of
807: eigenvalues and eigenvectors of a matrix to the infinite-dimensional case. The
808: Sturm-Liouville theory tells us that these operators have the
809: following spectra~\cite{FOCO}:
810: \numparts
811: \begin{equation}
812: {\rm Sp}(Q)=(-\infty, \infty) \, ,
813: \end{equation}
814: \begin{equation}
815: {\rm Sp}(P)=(-\infty, \infty) \, ,
816: \end{equation}
817: \begin{equation}
818: {\rm Sp}(H)=[0, \infty) \, .
819: \end{equation}
820: \endnumparts
821: These spectra coincide with those we would expect on physical grounds. We
822: expect the possible measurements of $Q$ to be the full real line, because
823: the particle can in principle reach any point of the real line. We also
824: expect the possible measurements of $P$ to be the full real line, since the
825: momentum of the particle is not restricted in magnitude or direction. The
826: possible measurements of $H$ have the same range as that of the kinetic
827: energy, because the potential does not have any wells of negative energy,
828: and therefore we expect the spectrum of $H$ to be the positive real line.
829:
830: To obtain the eigenfunction corresponding to each eigenvalue, we have to
831: solve the eigenvalue equation~(\ref{cketequeintro}) for each observable. Since
832: we are working in the position representation, we have to write
833: Eq.~(\ref{cketequeintro}) in the position representation for each observable:
834: \numparts
835: \begin{equation}
836: \langle x|Q|x'\rangle = x' \langle x|x'\rangle \, ,
837: \label{xqee}
838: \end{equation}
839: \begin{equation}
840: \langle x|P|p\rangle = p \langle x|p\rangle \, ,
841: \label{xpee}
842: \end{equation}
843: \begin{equation}
844: \langle x|H|E\rangle = E \langle x|E\rangle \, .
845: \label{xhee}
846: \end{equation}
847: \endnumparts
848: By recalling Eqs.~(\ref{fdopx}), (\ref{fdopp}) and (\ref{fdoph}), we can
849: write Eqs.~(\ref{xqee})-(\ref{xhee}) as
850: \numparts
851: \begin{equation}
852: x \langle x|x'\rangle = x' \langle x|x'\rangle \, ,
853: \label{eeq}
854: \end{equation}
855: \begin{equation}
856: -\rmi \hbar \frac{\rmd}{\rmd x} \langle x|p\rangle =
857: p \langle x|p\rangle \, ,
858: \label{eep}
859: \end{equation}
860: \begin{equation}
861: \left( -\frac{\hbar ^2}{2m}\frac{\rmd ^2}{\rmd x^2}
862: +V(x) \right) \langle x|E\rangle = E \langle x|E\rangle \, .
863: \label{eeh}
864: \end{equation}
865: \endnumparts
866: For each position $x'$, Eq.~(\ref{eeq}) yields the corresponding eigenfunction
867: of $Q$ as a delta function,
868: \begin{equation}
869: \langle x| x'\rangle = \delta (x-x') \, .
870: \label{deltax}
871: \end{equation}
872: For each momentum $p$, Eq.~(\ref{eep}) yields the corresponding eigenfunction
873: of $P$ as a plane wave,
874: \begin{equation}
875: \langle x| p\rangle = \frac{\rme ^{\rmi px/\hbar}}{\sqrt{2\pi \hbar}}
876: \, .
877: \label{expp}
878: \end{equation}
879: For each energy $E$, Eq.~(\ref{eeh}) yields the following two linearly
880: independent eigenfunctions~\cite{FOCO}:
881: \numparts
882: \begin{equation}
883: \langle x|E^+\rangle _{\rm r} =
884: \left( \frac{m}{2\pi k \hbar ^2} \right)^{1/2} \times
885: \left\{ \begin{array}{lc}
886: T (k)\rme ^{-\rmi kx} &-\infty <x<a \\
887: A_{\rm r}(k)\rme ^{\rmi \kappa x}+
888: B_{\rm r}(k) \rme^{-\rmi \kappa x} &a<x<b \\
889: R_{\rm r}(k)\rme ^{\rmi kx} + \rme ^{-\rmi kx} &b<x<\infty \, ,
890: \end{array}
891: \right.
892: \label{chir+}
893: \end{equation}
894: \begin{equation}
895: \langle x|E^+\rangle _{\rm l}=
896: \left( \frac{m}{2\pi k \hbar ^2} \right)^{1/2}
897: \times \left\{ \begin{array}{lc}
898: \rme ^{\rmi kx}+R_{\rm l}(k)\rme ^{-\rmi kx} &-\infty <x<a \\
899: A_{\rm l}(k)\rme ^{\rmi \kappa x}+B_{\rm l}(k)\rme ^{-\rmi \kappa x}
900: &a<x<b \\
901: T (k)\rme ^{\rmi kx} &b<x<\infty \, ,
902: \end{array}
903: \right.
904: \label{chil+}
905: \end{equation}
906: \endnumparts
907: where
908: \begin{equation}
909: k=\sqrt{\frac{2m}{\hbar ^2}E} \, , \quad
910: \kappa =\sqrt{\frac{2m}{\hbar ^2}(E-V_0)} \, ,
911: \end{equation}
912: and where the coefficients that appear in Eqs.~(\ref{chir+})-(\ref{chil+}) can
913: be easily found by the standard matching conditions at the discontinuities of
914: the potential~\cite{FOCO}. Thus, in contrast to the spectra of $Q$ and $P$,
915: the spectrum of $H$ is doubly degenerate.
916:
917: Physically,
918: the eigenfunction $\langle x|E^+\rangle _{\rm r}$ represents a particle of
919: energy $E$ that impinges on the barrier from the right (hence the subscript r)
920: and gets reflected to the right with probability amplitude $R_{\rm r}(k)$ and
921: transmitted to the left with probability amplitude $T (k)$, see
922: Fig.~\ref{fig:plus}a. The eigenfunction
923: $\langle x|E^+\rangle _{\rm l}$ represents a particle of energy $E$ that
924: impinges on the barrier from the left (hence the subscript l) and gets
925: reflected to the left with probability amplitude $R_{\rm l}(k)$ and
926: transmitted to the right with probability amplitude $T(k)$, see
927: Fig.~\ref{fig:plus}b.
928:
929: Note that, instead of~(\ref{chir+})-(\ref{chil+}), we could choose another
930: pair of linearly independent solutions of Eq.~(\ref{eeh}) as
931: follows~\cite{FOCO}:
932: \numparts
933: \begin{equation}
934: \langle x|E^-\rangle _{\rm r} =
935: \left( \frac{m}{2\pi k \hbar ^2} \right)^{1/2}
936: \times \left\{ \begin{array}{lc}
937: T^*(k)\rme ^{\rmi kx} &-\infty <x<a \\
938: A_{\rm r}^*(k)\rme ^{-\rmi \kappa x}+
939: B_{\rm r}^*(k) \rme ^{\rmi \kappa x}&a<x<b \\
940: R_{\rm r}^*(k)\rme ^{-\rmi kx} + \rme ^{\rmi kx} &b<x<\infty \, ,
941: \end{array}
942: \right.
943: \label{chir-}
944: \end{equation}
945: \begin{equation}
946: \langle x|E^-\rangle _{\rm l} =
947: \left( \frac{m}{2\pi k \hbar ^2} \right)^{1/2}
948: \times \left\{ \begin{array}{lc}
949: \rme ^{-\rmi kx}+R_{\rm l}^*(k)\rme ^{\rmi kx} &-\infty <x<a \\
950: A_{\rm l}^*(k)\rme ^{-\rmi \kappa x}+
951: B_{\rm l}^*(k)\rme ^{\rmi \kappa x}&a<x<b \\
952: T^*(k)\rme ^{-\rmi kx} &b<x<\infty \, ,
953: \end{array}
954: \right.
955: \label{chil-}
956: \end{equation}
957: \endnumparts
958: where the coefficients of these eigenfunctions can also be calculating
959: by means of the standard matching conditions at $x=a,b$~\cite{FOCO}. The
960: eigenfunction $\langle x|E^-\rangle _{\rm r}$
961: represents two plane waves---one impinging on the barrier from the left with
962: probability amplitude $T^*(k)$ and another impinging on the barrier from the
963: right with probability amplitude $R_{\rm r}^*(k)$---that combine in such a way
964: as to produce an outgoing plane wave to the right, see
965: Fig.~\ref{fig:minus}a. The eigenfunction
966: $\langle x|E^-\rangle _{\rm l}$ represents two other planes waves---one
967: impinging on the barrier from left with probability amplitude $R_{\rm l}^*(k)$
968: and another impinging on the
969: barrier from the right with probability amplitude $T^*(k)$---that combine in
970: such a way as to produce an outgoing wave to the left, see
971: Fig.~\ref{fig:minus}b. The eigensolutions
972: $\langle x|E^-\rangle _{\rm r,l}$ correspond to the {\it final} condition
973: of an outgoing plane wave propagating away from the barrier respectively to
974: the right and to the left, as opposed to $\langle x|E^+\rangle _{\rm r,l}$,
975: which correspond to the {\it initial} condition of a plane wave that
976: propagates towards the barrier respectively from the right and from the left.
977:
978: The eigenfunctions (\ref{deltax}), (\ref{expp}), (\ref{chir+})-(\ref{chil+})
979: and (\ref{chir-})-(\ref{chil-})
980: are not square integrable, that is, they do not belong to
981: $L^2$. Mathematically speaking, this is the reason why
982: they are
983: to be dealt with as distributions (note that all of them except for the delta
984: function are also proper functions). Physically speaking, they
985: are to be interpreted in analogy to electromagnetic plane waves, as we shall
986: see in Section~\ref{sec:phymean}.
987:
988:
989:
990:
991:
992:
993:
994:
995:
996:
997: \section{Construction of the rigged Hilbert space}
998: %\setcounter{equation}{0}
999: \label{sec:crhs}
1000:
1001:
1002:
1003:
1004: In the previous section, we saw that the observables of our system are
1005: implemented by unbounded operators with continuous spectrum. We also saw
1006: that the eigenfunctions of the observables do not belong to
1007: $L^2$. Thus, as we explained in Sec.~\ref{sec:why}, we need
1008: to construct the rigged Hilbert spaces of Eqs.~(\ref{RHStIntro})
1009: and (\ref{RHSpIntro}) [see Eqs.~(\ref{RHSCONTpr}) and (\ref{RHSCONTprb})
1010: below]. We start by constructing $\mathbf \Phi$.
1011:
1012:
1013:
1014:
1015: \subsection{Construction of $\mathbf \Phi \equiv \Sw$}
1016:
1017:
1018: The subspace $\mathbf \Phi$ is given by Eq.~(\ref{maximalinvas}). In view of
1019: expressions (\ref{fdopx}), (\ref{fdopp}) and (\ref{fdoph}), the elements of
1020: $\mathbf \Phi$ must fulfill the following conditions:
1021: \begin{itemize}
1022: \item[$\bullet$] they are infinitely differentiable, so the differentiation
1023: operation can be applied as many times as wished,
1024: \item[$\bullet$] they vanish at $x=a$ and $x=b$, so differentiation is
1025: meaningful at the discontinuities of the potential~\cite{ZERODERIV},
1026: \item[$\bullet$] the action of all powers of $Q$, $P$ and $H$ remains square
1027: integrable.
1028: \end{itemize}
1029: Hence,
1030: \begin{eqnarray}
1031: \hskip-1cm {\mathbf \Phi} =\{ \varphi \in L^2 \, | \
1032: \varphi \in C^{\infty}(\mathbb R), \ \varphi ^{(n)}(a)=\varphi ^{(n)}(b)=0
1033: \, , \ n=0,1,\ldots \, , \nonumber \\
1034: \hskip3.6cm
1035: P^nQ^mH^l\varphi (x) \in L^2
1036: \, , \ n,m,l=0,1, \ldots \} \, ,
1037: \label{ddomain}
1038: \end{eqnarray}
1039: where $C^{\infty}(\mathbb R)$ is the collection of infinite differentiable
1040: functions, and $\varphi ^{(n)}$ denotes the $n$th derivative of
1041: $\varphi$. From the last condition in Eq.~(\ref{ddomain}), we deduce that the
1042: elements of $\mathbf \Phi$ satisfy the following estimates:
1043: \begin{equation}
1044: \| \varphi \| _{n,m,l} \equiv
1045: \sqrt{\int_{-\infty}^{\infty}\rmd x
1046: \, \left| P^nQ^mH^l\varphi (x)\right| ^2 \,}
1047: < \infty \, , \quad n,m,l=0,1,\ldots \, .
1048: \label{nmnorms}
1049: \end{equation}
1050: These estimates mean that the action of any combination of any power of the
1051: observables remains square integrable. For this to happen, the functions
1052: $\varphi (x)$ must be infinitely differentiable and must fall off at infinity
1053: faster than any polynomial. The estimates~(\ref{nmnorms}) induce a topology
1054: on $\mathbf \Phi$, that is, they induce a meaning of convergence of
1055: sequences, in the following way. A sequence $\{ \varphi _{\alpha} \}$
1056: $\mathbf \Phi$-converges to $\varphi$ when $\{ \varphi _{\alpha} \}$ converges
1057: to $\varphi$ with respect to all the estimates~(\ref{nmnorms}),
1058: \begin{equation}
1059: \varphi _{\alpha}\,
1060: \mapupdown{\tau_{\mathbf \Phi}}{\alpha \to \infty}
1061: \, \varphi \quad {\rm if} \quad
1062: \| \varphi _{\alpha }-\varphi \| _{n,m,l}
1063: \, \mapupdown{}{\alpha \to \infty}\, 0 \, , \quad n,m,l=0,1, \ldots \, .
1064: \end{equation}
1065: Intuitively, a sequence $\varphi _{\alpha}$ converges to $\varphi$ if
1066: whenever we follow the terms of the sequence, we get closer and closer to
1067: the limit point $\varphi$ with respect to a certain sense of closeness. In our
1068: system, the notion of closeness is determined by the estimates
1069: $\| \ \|_{n,m,l}$, which originate from the physical
1070: requirements that led us to construct $\mathbf \Phi$.
1071:
1072: From Eqs.~(\ref{ddomain}) and (\ref{nmnorms}), we can see that $\mathbf \Phi$
1073: is very similar to the Schwartz space ${\cal S}(\mathbb R)$, the
1074: major differences being that the derivatives of the elements of $\mathbf \Phi$
1075: vanish at $x=a,b$ and that $\mathbf \Phi$ is not only invariant under
1076: $P$ and $Q$ but also under $H$. This is why we shall write
1077: \begin{equation}
1078: \mathbf \Phi \equiv \Sw \, .
1079: \end{equation}
1080:
1081: It is always a good, though lengthy exercise to check that $\Sw$ is indeed
1082: invariant under the action of the observables,
1083: \begin{equation}
1084: A \, \Sw \subset \Sw \, , \qquad A=P,Q,H \, .
1085: \end{equation}
1086: This invariance guarantees that the expectation values
1087: \begin{equation}
1088: (\varphi , A^n\varphi ) \, , \quad \varphi \in \Sw \, ,
1089: \ A=P, Q, H , \ n=0,1,\ldots
1090: \end{equation}
1091: are finite, and that the commutation relations~(\ref{cr1})-(\ref{cr3}) are
1092: well
1093: defined~\cite{CRVANISHES}. It can also be checked that $P$, $Q$ and $H$, which
1094: are not continuous
1095: with respect the topology of the Hilbert space $L^2$, are
1096: now continuous with respect to the topology $\tau _{\mathbf \Phi}$ of
1097: $\Sw$~\cite{FOCO,DIS}.
1098:
1099:
1100:
1101:
1102:
1103:
1104: \subsection{Construction of $\mathbf \Phi ^{\times}\equiv \Swt$. The Dirac
1105: kets}
1106:
1107:
1108:
1109: The space $\mathbf \Phi ^{\times}$ is simply the collection of
1110: $\tau _{\mathbf \Phi}$-continuous {\it antilinear} functionals over
1111: $\mathbf \Phi$~\cite{FUNCTIONAL}. By combining the spaces $\mathbf \Phi$,
1112: $\cal H$ and $\mathbf \Phi ^{\times}$, we obtain the RHS of our system,
1113: \begin{equation}
1114: \mathbf \Phi \subset {\cal H}\subset \mathbf \Phi ^{\times}
1115: \, ,
1116: \label{RHSCONT}
1117: \end{equation}
1118: which we denote in the position representation by
1119: \begin{equation}
1120: \rhsSwt \, .
1121: \label{RHSCONTpr}
1122: \end{equation}
1123: The space $\Swt$ is meant to accommodate the eigenkets $|p \rangle$,
1124: $|x\rangle$ and $|E^{\pm}\rangle _{\rm l,r}$ of $P$, $Q$ and $H$. In the
1125: remainder of this subsection, we construct these eigenkets explicitly and see
1126: that they indeed belong to $\Swt$. We shall also see
1127: that $|p \rangle$, $|x\rangle$ and $|E^{\pm}\rangle _{\rm l,r}$ are indeed
1128: eigenvectors of the observables.
1129:
1130: The definition of a ket is borrowed from the theory of
1131: distributions as follows~\cite{GELFAND}. Given a function $f(x)$ and a space
1132: of test functions $\mathbf \Phi$, the antilinear functional $F$ that
1133: corresponds to the function $f(x)$ is an integral operator whose kernel is
1134: precisely $f(x)$:
1135: \numparts
1136: \begin{equation}
1137: F(\varphi)=\int \rmd x \, \overline{\varphi (x)} f(x) \, ,
1138: \label{afFtff1}
1139: \end{equation}
1140: which in Dirac's notation becomes
1141: \begin{equation}
1142: \langle \varphi|F\rangle =\int \rmd x \, \langle \varphi|x\rangle
1143: \langle x|f\rangle \, .
1144: \label{afFtff2}
1145: \end{equation}
1146: \endnumparts
1147: It is important to keep in mind that, though related, the function $f(x)$ and
1148: the functional $F$ are two different things, the relation between them being
1149: that $f(x)$ is the kernel of $F$ when we write $F$ as an integral operator. In
1150: the physics literature, the term {\it distribution} is usually reserved
1151: for $f(x)$.
1152:
1153: Definition~(\ref{afFtff1}) provides the link between the quantum mechanical
1154: formalism and the theory of distributions. In practical applications, what
1155: one obtains from the quantum mechanical formalism is the distribution
1156: $f(x)$ (in this paper, the plane waves
1157: $\frac{1}{\sqrt{2\pi \hbar}} \rme ^{\rmi px/\hbar}$, the delta function
1158: $\delta (x-x')$ and the eigenfunctions
1159: $\langle x|E^{\pm}\rangle _{\rm l,r}$). Once $f(x)$ is given, one
1160: can use definition~(\ref{afFtff1}) to generate the functional
1161: $|F\rangle$. Then,
1162: the theory of distributions can be used to obtain the properties of the
1163: functional $|F\rangle$, which in turn yield the properties of the distribution
1164: $f(x)$.
1165:
1166: By using prescription~(\ref{afFtff1}), we can define for
1167: each eigenvalue $p$ the eigenket $|p\rangle$ associated with the
1168: eigenfunction (\ref{expp}):
1169: \numparts
1170: \begin{equation}
1171: \langle \varphi |p\rangle \equiv
1172: \int_{-\infty}^{\infty}\rmd x \, \overline{\varphi (x)}
1173: \frac{1}{\sqrt{2\pi \hbar}} \rme ^{\rmi px/\hbar} \, ,
1174: \label{definitionketp}
1175: \end{equation}
1176: which, using Dirac's notation for the integrand, becomes
1177: \begin{equation}
1178: \langle \varphi |p\rangle \equiv
1179: \int_{-\infty}^{\infty}\rmd x \, \langle \varphi |x\rangle
1180: \langle x|p\rangle \, .
1181: \label{definitionketpDirac}
1182: \end{equation}
1183: \endnumparts
1184: Similarly, for each $x$, we can define the ket $|x\rangle$ associated with
1185: the eigenfunction (\ref{deltax}) of the position operator as
1186: \numparts
1187: \begin{equation}
1188: \langle \varphi |x\rangle \equiv
1189: \int_{-\infty}^{\infty}\rmd x' \, \overline{\varphi (x')}
1190: \delta (x-x') \, ,
1191: \label{definitionketx}
1192: \end{equation}
1193: which, using Dirac's notation for the integrand, becomes
1194: \begin{equation}
1195: \langle \varphi |x\rangle \equiv
1196: \int_{-\infty}^{\infty}\rmd x' \, \langle \varphi | x'\rangle
1197: \langle x'|x\rangle \, .
1198: \label{definitionketxDirac}
1199: \end{equation}
1200: \endnumparts
1201: The definition of the kets $|E^{\pm}\rangle _{\rm l,r}$ that correspond to the
1202: Hamiltonian's eigenfunctions~(\ref{chir+})-(\ref{chil+}) and
1203: (\ref{chir-})-(\ref{chil-})
1204: follows the same prescription:
1205: \numparts
1206: \begin{equation}
1207: \langle \varphi |E^{\pm}\rangle _{\rm l,r} \equiv
1208: \int_{-\infty}^{\infty}\rmd x \, \overline{\varphi (x)}
1209: \langle x|E^{\pm}\rangle _{\rm l,r} \, ,
1210: \label{definitionketE}
1211: \end{equation}
1212: that is,
1213: \begin{equation}
1214: \langle \varphi |E^{\pm}\rangle _{\rm l,r} \equiv
1215: \int_{-\infty}^{\infty}\rmd x \, \langle \varphi |x\rangle
1216: \langle x|E^{\pm}\rangle _{\rm l,r} \, .
1217: \label{definitionketEDirac}
1218: \end{equation}
1219: \endnumparts
1220: (Note that this equation defines four different
1221: kets.) One can now show that the definition of the kets $|p\rangle$,
1222: $|x\rangle$ and $|E^{\pm}\rangle _{\rm l,r}$ makes sense, and that
1223: these kets indeed belong to the space of distributions $\Swt$~\cite{FOCO}.
1224:
1225: As in the general case of Eqs.~(\ref{afFtff1})-(\ref{afFtff2}),
1226: it is important to keep in mind the difference between eigenfunctions and
1227: kets. For instance, $\langle x|p\rangle$ is an eigenfunction
1228: of a differential equation, Eq.~(\ref{eep}), whereas $|p\rangle$ is a
1229: functional, the relation between them being given by
1230: Eq.~(\ref{definitionketpDirac}). A similar relation holds between
1231: $\langle x'|x\rangle$ and $|x\rangle$, and between
1232: $\langle x|E^{\pm}\rangle _{\rm l,r}$ and $|E^{\pm}\rangle _{\rm l,r}$. It
1233: is also important to keep in mind that ``scalar products'' like
1234: $\langle x|p\rangle$, $\langle x'|x\rangle$ or
1235: $\langle x|E^{\pm}\rangle _{\rm l,r}$ do not represent an actual scalar
1236: product of two functionals; these ``scalar products'' are simply solutions
1237: to differential equations.
1238:
1239: We now turn to the question of whether the kets $|p\rangle$, $|x\rangle$ and
1240: $|E^{\pm}\rangle_{\rm l,r}$ are eigenvectors of the corresponding
1241: observable [see Eqs.~(\ref{kpeP})-(\ref{kEeH}) below]. Since the
1242: observables act in principle only on their Hilbert space domains, and since
1243: the kets lie outside the Hilbert space, we need to extend the definition
1244: of the observables from $\mathbf \Phi$ into $\mathbf \Phi ^{\times}$, in order
1245: to specify how the observables act on the kets. The theory of distributions
1246: provides us with a precise prescription of how an observable acts on
1247: $\mathbf \Phi ^{\times}$, and therefore of how it acts on the kets, as
1248: follows~\cite{GELFAND}. The action of a self-adjoint operator $A$ on a
1249: functional $|F\rangle \in {\mathbf \Phi}^{\times}$ is defined as
1250: \begin{equation}
1251: \langle \varphi |A|F\rangle \equiv \langle A\varphi |F\rangle
1252: \, , \quad \mbox{for all} \ \varphi \ \mbox{in} \ \mathbf \Phi \, .
1253: \label{adualext}
1254: \end{equation}
1255: Note that this definition extends the Hilbert space definition of a
1256: self-adjoint operator,
1257: \begin{equation}
1258: (f,Ag)=(Af,g) \, ,
1259: \label{saosr}
1260: \end{equation}
1261: which is valid only when $f$ and $g$ belong to the domain of
1262: $A$. In turn, Eq.~(\ref{adualext}) can be used to define the notion
1263: of eigenket of an observable: A functional $|a\rangle$ in
1264: $\mathbf \Phi ^{\times}$ is an eigenket of $A$ with eigenvalue $a$ if
1265: \begin{equation}
1266: \langle \varphi |A|a\rangle = \langle A \varphi |a\rangle =
1267: a \langle \varphi |a\rangle \, , \quad
1268: \mbox{for all} \ \varphi \ \mbox{in} \ \mathbf \Phi \, .
1269: \label{eeRHS}
1270: \end{equation}
1271: When the ``left sandwiching'' of this equation with the elements of
1272: $\mathbf \Phi$ is understood and therefore omitted, we shall simply write
1273: \begin{equation}
1274: A|a\rangle = a |a\rangle \, ,
1275: \label{eeRHSws}
1276: \end{equation}
1277: which is just Dirac's eigenket equation (\ref{cketequeintro}). Thus,
1278: Dirac's eigenket equation acquires a precise meaning through
1279: Eq.~(\ref{eeRHS}), in the sense that it has to be understood as
1280: ``left sandwiched'' with the wave functions $\varphi$ of $\mathbf \Phi$.
1281:
1282: By using definition~(\ref{eeRHS}), one can show that $|p\rangle$, $|x\rangle$
1283: and $|E^{\pm}\rangle_{\rm l,r}$ are indeed eigenvectors of $P$, $Q$ and $H$,
1284: respectively~\cite{FOCO}:
1285: \begin{equation}
1286: P|p\rangle=p|p\rangle \, , \quad p\in \mathbb R \, ,
1287: \label{kpeP}
1288: \end{equation}
1289: \begin{equation}
1290: Q|x\rangle=x|x\rangle \, , \quad x \in \mathbb R \, ,
1291: \end{equation}
1292: \begin{equation}
1293: H|E^{\pm}\rangle_{\rm l,r} =E|E^{\pm}\rangle _{\rm l,r}
1294: \, , \quad E\in [0,\infty ) \, .
1295: \label{kEeH}
1296: \end{equation}
1297:
1298:
1299:
1300:
1301:
1302:
1303:
1304:
1305:
1306: \subsection{Construction of $\mathbf \Phi ^{\prime}\equiv \Swp$. The Dirac
1307: bras}
1308:
1309:
1310:
1311: In complete analogy with the construction of the Dirac kets, we construct in
1312: this subsection the Dirac bras $\langle p|$, $\langle x|$ and
1313: $_{\rm l,r}\langle ^{\pm}E|$
1314: of $P$, $Q$ and $H$. Mathematically, the Dirac bras are distributions that
1315: belong to the space $\mathbf \Phi ^{\prime}$, which is the space of
1316: {\it linear} functionals over $\mathbf \Phi$~\cite{FUNCTIONAL}. The
1317: corresponding RHS is
1318: \begin{equation}
1319: {\mathbf \Phi} \subset {\cal H} \subset {\mathbf \Phi}^{\prime} \, ,
1320: \end{equation}
1321: which we denote in the position representation by
1322: \begin{equation}
1323: \Sw \subset L^2 \subset \Swp \, .
1324: \label{RHSCONTprb}
1325: \end{equation}
1326:
1327: Likewise the definition of a ket, the definition of a bra is borrowed from
1328: the theory of distributions~\cite{GELFAND}. Given a function $f(x)$ and
1329: a space of test functions $\mathbf \Phi$, the linear functional $\tilde{F}$
1330: generated by the function $f(x)$ is an integral operator whose kernel
1331: is the complex conjugate of $f(x)$:
1332: \numparts
1333: \begin{equation}
1334: \tilde{F}(\varphi)=\int \rmd x \, \varphi (x) \overline{f(x)} \, ,
1335: \label{afGtff1}
1336: \end{equation}
1337: which in Dirac's notation becomes
1338: \begin{equation}
1339: \langle F|\varphi \rangle =\int \rmd x \, \langle f|x\rangle
1340: \langle x|\varphi \rangle \, .
1341: \end{equation}
1342: \endnumparts
1343: Note that this definition is very similar to that of a linear functional,
1344: Eq.~(\ref{afFtff1}), except that the complex conjugation affects $f(x)$
1345: rather than $\varphi (x)$, which makes $\tilde{F}$ linear rather than
1346: antilinear. Likewise the
1347: antilinear case~(\ref{afFtff1}), it is important to keep in mind that, though
1348: related, the function $f(x)$ and the functional $\tilde{F}$ are two different
1349: objects, the relation between them being that $\overline{f(x)}$ is the kernel
1350: of $\tilde{F}$ when we write $\tilde{F}$ as an integral operator.
1351:
1352: By using prescription~(\ref{afGtff1}), we can now define for each eigenvalue
1353: $p$ the eigenbra $\langle p|$ associated with the eigenfunction~(\ref{expp}):
1354: \numparts
1355: \begin{equation}
1356: \langle p| \varphi \rangle \equiv
1357: \int_{-\infty}^{\infty}\rmd x \, \varphi (x)
1358: \frac{1}{\sqrt{2\pi \hbar}} \rme ^{-\rmi px/\hbar} \, ,
1359: \label{definitionbrap}
1360: \end{equation}
1361: which, using Dirac's notation for the integrand, becomes
1362: \begin{equation}
1363: \langle p| \varphi \rangle \equiv
1364: \int_{-\infty}^{\infty}\rmd x \, \langle p|x \rangle
1365: \langle x| \varphi \rangle \, .
1366: \label{definitionbrapDirac}
1367: \end{equation}
1368: \endnumparts
1369: Comparison with Eq.~(\ref{definitionketp}) shows that the action of
1370: $\langle p|$ is the complex conjugate of the action of $|p \rangle$,
1371: \begin{equation}
1372: \langle p| \varphi \rangle = \overline{\langle \varphi |p \rangle} \, ,
1373: \end{equation}
1374: and that
1375: \begin{equation}
1376: \langle p| x \rangle = \overline{\langle x |p \rangle} =
1377: \frac{1}{\sqrt{2\pi \hbar}} \rme ^{-\rmi px/\hbar} \, .
1378: \label{pxiscmplxp}
1379: \end{equation}
1380: The bra $\langle x|$ is defined as
1381: \numparts
1382: \begin{equation}
1383: \langle x| \varphi \rangle \equiv
1384: \int_{-\infty}^{\infty}\rmd x' \, \varphi (x')
1385: \delta (x-x') \, ,
1386: \label{definitionbrax}
1387: \end{equation}
1388: which, using Dirac's notation for the integrand, becomes
1389: \begin{equation}
1390: \langle x|\varphi \rangle \equiv
1391: \int_{-\infty}^{\infty}\rmd x' \, \langle x | x'\rangle
1392: \langle x'| \varphi \rangle \, .
1393: \label{definitionbraxDirac}
1394: \end{equation}
1395: \endnumparts
1396: Comparison with Eq.~(\ref{definitionketx}) shows that the action of
1397: $\langle x|$ is complex conjugated to the action of $|x \rangle$,
1398: \begin{equation}
1399: \langle x| \varphi \rangle = \overline{\langle \varphi |x \rangle} \, ,
1400: \end{equation}
1401: and that
1402: \begin{equation}
1403: \langle x| x' \rangle = \langle x' |x \rangle = \delta (x-x') \, .
1404: \end{equation}
1405: Analogously, the eigenbras of the Hamiltonian are defined as
1406: \numparts
1407: \begin{equation}
1408: _{\rm l,r}\langle ^{\pm}E|\varphi\rangle \equiv
1409: \int_{-\infty}^{\infty}\rmd x \ \varphi (x) \
1410: _{\rm l,r} \langle ^{\pm}E|x\rangle \, ,
1411: \label{definitionbraE}
1412: \end{equation}
1413: that is,
1414: \begin{equation}
1415: _{\rm l,r}\langle ^{\pm}E|\varphi\rangle \equiv
1416: \int_{-\infty}^{\infty}\rmd x \
1417: _{\rm l,r}\langle ^{\pm}E|x\rangle \langle x|\varphi \rangle \, ,
1418: \label{definitionbraEDirac}
1419: \end{equation}
1420: \endnumparts
1421: where
1422: \begin{equation}
1423: _{\rm l,r}\langle ^{\pm}E|x\rangle =
1424: \overline{\langle x|E^{\pm}\rangle}_{\rm l,r} \, .
1425: \end{equation}
1426: (Note that in Eq.~(\ref{definitionbraE}) we have defined four different
1427: bras.) Comparison of Eq.~(\ref{definitionbraE}) with
1428: Eq.~(\ref{definitionketE}) shows that
1429: the actions of the bras $_{\rm l,r}\langle ^{\pm}E|$ are the complex
1430: conjugates of the actions of the kets $|E ^{\pm} \rangle _{\rm l,r}$:
1431: \begin{equation}
1432: _{\rm l,r}\langle ^{\pm}E|\varphi\rangle =
1433: \overline{\langle \varphi |E ^{\pm} \rangle}_{\rm l,r} \, .
1434: \label{braketccE}
1435: \end{equation}
1436: Now, by using the RHS mathematics, one can show that the definitions of
1437: $\langle p|$, $\langle x|$ and $_{\rm l,r} \langle ^{\pm}E|$ make sense
1438: and that $\langle p|$, $\langle x|$ and $_{\rm l,r} \langle ^{\pm}E|$
1439: belong to $\Swp$~\cite{FOCO}.
1440:
1441: Our next task is to see that the bras we just defined are left eigenvectors of
1442: the corresponding observable [see Eqs.~(\ref{bpeP})-(\ref{bEeH}) below]. For
1443: this purpose, we need to specify how the observables act on the bras,
1444: that is, how they act on the dual space $\Swp$. We shall do so in analogy to
1445: the definition of their action on the kets, by means of the theory of
1446: distributions~\cite{GELFAND}. The action to the left of a self-adjoint
1447: operator $A$ on a linear functional $\langle F| \in {\mathbf \Phi}^{\prime}$
1448: is defined as
1449: \begin{equation}
1450: \langle F|A|\varphi \rangle \equiv \langle F| A\varphi \rangle
1451: \, , \quad \mbox{for all} \ \varphi \ \mbox{in} \ \mathbf \Phi \, .
1452: \label{dualext}
1453: \end{equation}
1454: Likewise definition~(\ref{adualext}), this definition generalizes
1455: Eq.~(\ref{saosr}). In turn, Eq.~(\ref{dualext}) can be used to define
1456: the notion of eigenbra of an observable: A functional $\langle a|$ in
1457: $\mathbf \Phi ^{\prime}$ is an eigenbra of $A$ with eigenvalue $a$ if
1458: \begin{equation}
1459: \langle a |A|\varphi \rangle = \langle a|A \varphi \rangle =
1460: a \langle a|\varphi \rangle \, , \quad
1461: \mbox{for all} \ \varphi \ \mbox{in} \ \mathbf \Phi \, .
1462: \label{eeRHSb}
1463: \end{equation}
1464: When the ``right sandwiching'' of this equation with the elements of
1465: $\mathbf \Phi$ is understood and therefore omitted, we shall simply write
1466: \begin{equation}
1467: \langle a |A = a \langle a| \, ,
1468: \label{eeRHSwsb}
1469: \end{equation}
1470: which is just Dirac's eigenbra equation (\ref{braeqeneintro}). Thus,
1471: Dirac's eigenbra equation acquires a precise meaning through
1472: Eq.~(\ref{eeRHSb}), in the sense that it has to be understood as
1473: ``right sandwiched'' with the wave functions $\varphi$ of $\mathbf \Phi$.
1474:
1475: By using definition~(\ref{eeRHSb}),
1476: one can show that $\langle p|$, $\langle x|$ and $_{\rm l,r} \langle ^{\pm}E|$
1477: are indeed left eigenvectors of $P$, $Q$ and $H$, respectively~\cite{FOCO}:
1478: \begin{equation}
1479: \langle p|P=p\langle p| \, , \quad p\in \mathbb R \, ,
1480: \label{bpeP}
1481: \end{equation}
1482: \begin{equation}
1483: \langle x|Q=x\langle x| \, , \quad x \in \mathbb R \, ,
1484: \end{equation}
1485: \begin{equation}
1486: _{\rm l,r} \langle ^{\pm}E|H=
1487: E \hskip0.12cm
1488: _{\rm l,r} \langle ^{\pm}E| \, , \quad E\in [0,\infty ) \, .
1489: \label{bEeH}
1490: \end{equation}
1491:
1492: It is worthwhile noting that, in accordance with
1493: Dirac's formalism, there is a one-to-one correspondence between bras and
1494: kets~\cite{ONEONEROBERTS}; that is, given an observable $A$, to each element
1495: $a$ in the
1496: spectrum of $A$ there correspond a bra $\langle a|$ that is a left eigenvector
1497: of $A$ and also a ket $|a\rangle$ that is a right eigenvector of $A$. The bra
1498: $\langle a|$ belongs to $\mathbf \Phi ^{\prime}$, whereas the ket $|a\rangle$
1499: belongs to $\mathbf \Phi ^{\times}$.
1500:
1501:
1502:
1503:
1504: \subsection{The Dirac basis expansions}
1505:
1506:
1507: A crucial ingredient of Dirac's formalism is that the bras and kets of an
1508: observable form a complete basis system, see Eqs.~(\ref{introDirbaexp}) and
1509: (\ref{introresiden}). When applied to $P$, $Q$ and $H$,
1510: Eq.~(\ref{introresiden}) yields
1511: \begin{equation}
1512: \int_{-\infty}^{\infty} \rmd p \, |p\rangle \langle p| = I \, ,
1513: \label{resonidentP}
1514: \end{equation}
1515: \begin{equation}
1516: \int_{-\infty}^{\infty} \rmd x' \, |x'\rangle \langle x'| = I \, ,
1517: \label{resonidentQ}
1518: \end{equation}
1519: \begin{equation}
1520: \int_{0}^{\infty} \rmd E \, |E^{\pm}\rangle _{\rm l}\,
1521: _{\rm l}\langle ^{\pm}E| +
1522: \int_{0}^{\infty}\rmd E \, |E^{\pm}\rangle _{\rm r}\,
1523: _{\rm r}\langle ^{\pm}E| = I \, ,
1524: \label{resonidentH}
1525: \end{equation}
1526: In the present subsection, we derive various Dirac basis expansions
1527: for the algebra of the 1D rectangular barrier potential. We will do so by
1528: formally sandwiching Eqs.~(\ref{resonidentP})-(\ref{resonidentH}) in between
1529: different vectors.
1530:
1531: If we sandwich Eqs.~(\ref{resonidentP})-(\ref{resonidentH}) in between
1532: $\langle x|$ and $\varphi$, we obtain
1533: \begin{equation}
1534: \langle x|\varphi \rangle= \int_{-\infty}^{\infty} \rmd p \,
1535: \langle x|p\rangle \langle p|\varphi \rangle \, ,
1536: \label{resonidentPxphi}
1537: \end{equation}
1538: \begin{equation}
1539: \langle x|\varphi \rangle = \int_{-\infty}^{\infty} \rmd x' \,
1540: \langle x|x'\rangle \langle x'|\varphi \rangle \, ,
1541: \label{resonidentQxphi}
1542: \end{equation}
1543: \begin{equation}
1544: \langle x|\varphi \rangle = \int_{0}^{\infty} \rmd E \,
1545: \langle x|E^{\pm}\rangle _{\rm l} \,
1546: _{\rm l}\langle ^{\pm}E|\varphi \rangle +
1547: \int_{0}^{\infty} \rmd E \,
1548: \langle x|E^{\pm}\rangle _{\rm r} \,
1549: _{\rm r}\langle ^{\pm}E|\varphi \rangle \, .
1550: \label{resonidentHxphi}
1551: \end{equation}
1552: Equations~(\ref{resonidentPxphi})-(\ref{resonidentHxphi}) can be rigorously
1553: proved by way of the RHS~\cite{FOCO}. In proving
1554: these equations, we give meaning to
1555: Eqs.~(\ref{resonidentP})-(\ref{resonidentH}), which are just formal
1556: equations: Equations~(\ref{resonidentP})-(\ref{resonidentH}) have always to be
1557: understood as part of a ``sandwich.'' Note
1558: that Eqs.~(\ref{resonidentPxphi})-(\ref{resonidentHxphi})
1559: are not valid for every element of the Hilbert space but only for those
1560: $\varphi$ that belong to $\Sw$, because the action
1561: of the bras and kets is well defined only on $\Sw$~\cite{EXTOHS}. Thus, the
1562: RHS, rather than just the Hilbert space, fully justifies the Dirac basis
1563: expansions. Physically, the Dirac basis expansions provide the means to
1564: visualize wave packet formation out of a continuous linear superposition
1565: of bras and kets.
1566:
1567: We can obtain similar expansions to
1568: Eqs.~(\ref{resonidentPxphi})-(\ref{resonidentHxphi}) by sandwiching
1569: Eqs.~(\ref{resonidentP})-(\ref{resonidentH}) in between other vectors. For
1570: example, sandwiching Eq.~(\ref{resonidentQ}) in between $\langle p|$ and
1571: $\varphi$ yields~\cite{FOCO}
1572: \begin{equation}
1573: \langle p|\varphi \rangle =
1574: \int_{-\infty}^{\infty}\rmd x \
1575: \langle p|x\rangle
1576: \langle x|\varphi \rangle \, ,
1577: \label{inveqDvaeipx}
1578: \end{equation}
1579: and sandwiching Eq.~(\ref{resonidentQ}) in between
1580: $_{\rm l,r}\langle ^{\pm}E|$ and $\varphi$ yields~\cite{FOCO}
1581: \begin{equation}
1582: _{\rm l,r}\langle ^{\pm}E|\varphi \rangle =
1583: \int_{-\infty}^{\infty}\rmd x \
1584: _{\rm l,r}\langle ^{\pm}E|x\rangle
1585: \langle x|\varphi \rangle \, .
1586: \label{inveqDvaeix}
1587: \end{equation}
1588: It is worthwhile noting the parallel between the Dirac basis
1589: expansions and the Fourier expansions (\ref{resonidentPxphi}) and
1590: (\ref{inveqDvaeipx})~\cite{FOCO}. This parallel will be used in
1591: Sec.~\ref{sec:phymean} to physically interpret the Dirac bras and kets.
1592:
1593: We can also sandwich Eqs.~(\ref{resonidentP})-(\ref{resonidentH}) in between
1594: two elements $\psi$ and $\varphi$ of $\Sw$, and obtain~\cite{FOCO}
1595: \begin{equation}
1596: (\varphi ,\psi ) = \int_{-\infty}^{\infty}\rmd p \,
1597: \langle \varphi |p\rangle \langle p|\psi \rangle \, ,
1598: \label{spinofpbk}
1599: \end{equation}
1600: \begin{equation}
1601: (\varphi ,\psi ) = \int_{-\infty}^{\infty}\rmd x \,
1602: \langle \varphi |x\rangle \langle x|\psi \rangle \, ,
1603: \label{spinofxbk}
1604: \end{equation}
1605: \begin{equation}
1606: (\varphi ,\psi ) = \int_0^{\infty}\rmd E\,
1607: \langle \varphi |E^{\pm}\rangle_{\rm l}\,
1608: _{\rm l}\langle ^{\pm}E|\psi \rangle +
1609: \int_0^{\infty}\rmd E\,
1610: \langle \varphi |E^{\pm}\rangle_{\rm r}\,
1611: _{\rm r}\langle ^{\pm}E|\psi \rangle \, .
1612: \label{spinofEbk}
1613: \end{equation}
1614: Equations~(\ref{spinofpbk})-(\ref{spinofEbk}) allow us to calculate the
1615: overlap of two wave functions $\varphi$ and $\psi$ by way of the action
1616: of the bras and kets on those wave functions.
1617:
1618: The last aspect of Dirac's formalism we need to implement is
1619: prescription~(\ref{introactionA}), which expresses the action of an observable
1620: $A$ in terms of the action of its bras and kets. When applied to $P$, $Q$ and
1621: $H$, prescription~(\ref{introactionA}) yields
1622: \begin{equation}
1623: P = \int_{-\infty}^{\infty}\rmd p \, p |p\rangle \langle p| \, ,
1624: \label{presAP}
1625: \end{equation}
1626: \begin{equation}
1627: Q = \int_{-\infty}^{\infty}\rmd x \, x |x\rangle \langle x| \, ,
1628: \label{presAQ}
1629: \end{equation}
1630: \begin{equation}
1631: H = \int_{0}^{\infty}\rmd E \, E |E^{\pm}\rangle_{\rm l}\,
1632: _{\rm l}\langle ^{\pm}E| +
1633: \int_{0}^{\infty}\rmd E \, E |E^{\pm}\rangle_{\rm r}\,
1634: _{\rm r}\langle ^{\pm}E| \, .
1635: \label{presAH}
1636: \end{equation}
1637: Needless to say, these equations are formal expressions that acquire meaning
1638: when properly sandwiched. For example, sandwiching them in between
1639: $\langle x|$ and $\varphi$ yields~\cite{FOCO}
1640: \begin{equation}
1641: \langle x|P \varphi \rangle =\int_{-\infty}^{\infty}\rmd p \, p
1642: \langle x |p\rangle \langle p|\varphi \rangle \, ,
1643: \label{GMT2xP}
1644: \end{equation}
1645: \begin{equation}
1646: \langle x|Q \varphi \rangle = \int_{-\infty}^{\infty}\rmd x' \, x'
1647: \langle x |x'\rangle \langle x'|\varphi \rangle \, ,
1648: \label{GMT2xQ}
1649: \end{equation}
1650: \begin{equation}
1651: \langle x|H \varphi \rangle = \int_0^{\infty} \rmd E \,
1652: E \langle x |E^{\pm}\rangle_{\rm l}\,
1653: _{\rm l}\langle ^{\pm}E|\varphi \rangle +
1654: \int_0^{\infty}\rmd E\, E
1655: \langle x |E^{\pm}\rangle_{\rm r}\,
1656: _{\rm r}\langle ^{\pm}E|\varphi \rangle \, ,
1657: \label{GMT2xH}
1658: \end{equation}
1659: and sandwiching them in between two
1660: elements $\varphi$ and $\psi$ of $\Sw$ yields~\cite{FOCO}
1661: \begin{equation}
1662: (\varphi ,P \psi )=\int_{-\infty}^{\infty}\rmd p \, p
1663: \langle \varphi |p\rangle \langle p|\psi \rangle \, ,
1664: \label{GMT2P}
1665: \end{equation}
1666: \begin{equation}
1667: (\varphi ,Q \psi )=\int_{-\infty}^{\infty}\rmd x \, x
1668: \langle \varphi |x\rangle \langle x|\psi \rangle \, ,
1669: \label{GMT2Q}
1670: \end{equation}
1671: \begin{equation}
1672: (\varphi ,H \psi )= \int_0^{\infty} \rmd E \,
1673: E \langle \varphi |E^{\pm}\rangle_{\rm l}\,
1674: _{\rm l}\langle ^{\pm}E|\psi \rangle +
1675: \int_0^{\infty}\rmd E\, E
1676: \langle \varphi |E^{\pm}\rangle_{\rm r}\,
1677: _{\rm r}\langle ^{\pm}E|\psi \rangle \, .
1678: \label{GMT2H}
1679: \end{equation}
1680: Note that, in particular, the operational definition of an
1681: observable---according to which an observable is simply an operator whose
1682: eigenvectors form a complete basis such that Eqs.~(\ref{introDirbaexp}),
1683: (\ref{introresiden}) and~(\ref{introactionA}) hold, see for example
1684: Ref.~\cite{COHEN}---acquires meaning within the~RHS.
1685:
1686: The sandwiches we have made so far always involved at least a wave function
1687: $\varphi$ of $\Sw$. When
1688: the sandwiches do not involve elements of $\Sw$ at all, we obtain expressions
1689: that are simply formal. These formal expressions are often useful though,
1690: because they help us understand the meaning of concepts such as the delta
1691: normalization or the ``matrix elements'' of an operator. Let us start with the
1692: meaning of the delta normalization. When we sandwich Eq.~(\ref{resonidentQ})
1693: in between $\langle p'|$ and $|p\rangle$, we get
1694: \begin{equation}
1695: \int_{-\infty}^{\infty} \rmd x \,
1696: \langle p'|x\rangle \langle x|p\rangle = \langle p'|p \rangle \, .
1697: \label{resonidentQsppp}
1698: \end{equation}
1699: This equation is a formal expression that is to be understood in a
1700: distributional sense, that is, both sides must appear smeared out by a smooth
1701: function $\varphi (p) =\langle p|\varphi \rangle$ in an integral over $p$:
1702: \begin{equation}
1703: \int_{-\infty}^{\infty} \rmd p \, \varphi (p)
1704: \int_{-\infty}^{\infty} \rmd x \,
1705: \langle p'|x\rangle \langle x|p\rangle =
1706: \int_{-\infty}^{\infty} \rmd p \, \varphi (p) \langle p'|p \rangle \, .
1707: \label{smeresonidentQsppp}
1708: \end{equation}
1709: The left-hand side of Eq.~(\ref{smeresonidentQsppp}) can be written as
1710: \begin{eqnarray}
1711: \int_{-\infty}^{\infty} \rmd x \, \langle p'|x\rangle
1712: \int_{-\infty}^{\infty} \rmd p \, \varphi (p)
1713: \langle x|p\rangle
1714: &=&
1715: \int_{-\infty}^{\infty} \rmd x \, \langle p'|x\rangle
1716: \int_{-\infty}^{\infty} \rmd p \,
1717: \langle x|p\rangle\langle p|\varphi \rangle \nonumber \\
1718: &=&
1719: \int_{-\infty}^{\infty} \rmd x \,
1720: \langle p'|x\rangle \langle x|\varphi \rangle \nonumber \\
1721: &=& \varphi (p')
1722: \label{formdelnp}
1723: \end{eqnarray}
1724: Plugging Eq.~(\ref{formdelnp}) into Eq.~(\ref{smeresonidentQsppp}) leads to
1725: \begin{equation}
1726: \int_{-\infty}^{\infty} \rmd p \, \varphi (p) \langle p'|p\rangle =
1727: \varphi (p') \, .
1728: \label{vpapaovap}
1729: \end{equation}
1730: By recalling the definition of the delta function, we see that
1731: Eq.~(\ref{vpapaovap}) leads to
1732: \begin{equation}
1733: \langle p'|p\rangle = \delta (p-p') \, ,
1734: \label{pppdeltappp}
1735: \end{equation}
1736: and to
1737: \begin{equation}
1738: \int_{-\infty}^{\infty}\rmd x \, \langle p'|x\rangle \langle x|p\rangle =
1739: \delta (p-p') \, .
1740: \label{pppdeltappp2}
1741: \end{equation}
1742: By using Eq.~(\ref{pxiscmplxp}), we can write Eq.~(\ref{pppdeltappp2}) in
1743: a well-known form:
1744: \begin{equation}
1745: \frac{1}{2\pi \hbar} \int_{-\infty}^{\infty}
1746: \rmd x \, \rme^{\rmi(p-p')x/\hbar} = \delta (p-p') \, .
1747: \label{pppdeltappp3}
1748: \end{equation}
1749: This formal equation is interpreted by saying that the bras and kets of the
1750: momentum operator are delta normalized. That the energy bras and kets are
1751: also delta normalized can be seen in a similar, though slightly more involved
1752: way~\cite{FP02}:
1753: \numparts
1754: \begin{equation}
1755: _{\alpha}\langle ^{\pm}E'|E^{\pm} \rangle _{\beta} =
1756: \delta (E-E') \, \delta _{\alpha \beta} \, ,
1757: \label{deltanornebk1}
1758: \end{equation}
1759: \begin{equation}
1760: \int_{-\infty}^{\infty} \rmd x \ _{\alpha}\langle ^{\pm}E'|x\rangle
1761: \langle x|E^{\pm} \rangle _{\beta} =
1762: \delta (E-E') \, \delta _{\alpha \beta} \, ,
1763: \label{deltanornebk2}
1764: \end{equation}
1765: \endnumparts
1766: where $\alpha, \beta$ stand for the labels ${\rm l,r}$ that respectively
1767: denote left and right incidence. The derivation of
1768: expressions involving the Dirac delta function such as
1769: Eqs.~(\ref{pppdeltappp}), (\ref{pppdeltappp3})
1770: or~(\ref{deltanornebk1})-(\ref{deltanornebk2}) shows
1771: that these formal expressions must be understood in a distributional sense,
1772: that is, as kernels of integrals that include the wave functions $\varphi$
1773: of $\Sw$, like in Eq.~(\ref{smeresonidentQsppp}).
1774:
1775: In a similar way, we can also understand the meaning of the
1776: ``matrix elements'' of the observables in a particular representation, e.g.:
1777: \begin{equation}
1778: \langle x|Q|x'\rangle = x' \, \delta (x-x') \, ,
1779: \label{meQxx}
1780: \end{equation}
1781: \begin{equation}
1782: \langle x|P|x'\rangle = -\rmi \hbar \frac{\rmd}{\rmd x} \
1783: \delta (x-x') \, ,
1784: \label{mePxx}
1785: \end{equation}
1786: \begin{equation}
1787: \langle x|H|x'\rangle =
1788: \left( -\frac{\hbar ^2}{2m}\frac{\rmd ^2}{\rmd x^2}+V(x) \right)
1789: \delta (x-x') \, .
1790: \label{meHxx}
1791: \end{equation}
1792: Equations~(\ref{meQxx})-(\ref{meHxx}) can be obtained by formally inserting
1793: Eq.~(\ref{resonidentQ}) into respectively Eq.~(\ref{fdopx}), (\ref{fdopp})
1794: and (\ref{fdoph}).
1795:
1796: It is illuminating to realize that the expressions~(\ref{meQxx})-(\ref{meHxx})
1797: generalize the matrix representation of an observable $A$ in a
1798: finite-dimensional Hilbert space. If $a_1, \ldots ,a_N$ are the eigenvalues
1799: of $A$, then, in the basis $\{ |a_1 \rangle , \ldots , |a_N \rangle \}$,
1800: $A$ is represented as
1801: \begin{equation}
1802: A\equiv \left( \begin{array}{cccc}
1803: a_1 & 0 & \cdots & 0 \\
1804: 0& a_2 & \cdots & 0 \\
1805: \cdots & \cdots & \cdots & \cdots \\
1806: 0& 0 & \cdots & a_N
1807: \end{array}
1808: \right) \, ,
1809: \end{equation}
1810: which in Dirac's notation reads as
1811: \begin{equation}
1812: \langle a_i|A|a_j\rangle = a_i \delta _{ij} \, .
1813: \label{matreleA}
1814: \end{equation}
1815: Clearly, expressions~(\ref{meQxx})-(\ref{meHxx}) are the infinite-dimensional
1816: extension of expression~(\ref{matreleA}).
1817:
1818:
1819:
1820:
1821:
1822:
1823:
1824:
1825:
1826: \section{Physical meaning of the Dirac bras and kets}
1827: \label{sec:phymean}
1828:
1829:
1830:
1831: The bras and kets associated with eigenvalues in the continuous spectrum are
1832: not normalizable. Hence, the standard probabilistic interpretation does not
1833: apply to them straightforwardly. In this section, we
1834: are going to generalize the probabilistic interpretation of normalizable
1835: states to the non-normalizable bras and kets. As well, in order to gain
1836: further insight into the physical meaning of bras and kets, we shall present
1837: the analogy between classical plane waves and the bras and kets.
1838:
1839: In Quantum Mechanics, the scalar product of the Hilbert space is employed to
1840: calculate probability amplitudes. In our example, the Hilbert space is
1841: $L^2$, and the corresponding scalar product is given
1842: by Eq.~(\ref{scapro}). That an eigenvalue of an observable $A$ lies in the
1843: discrete or in the continuous part of the spectrum is determined by this
1844: scalar product. An eigenvalue $a_n$ belongs to the discrete part of the spectrum when its
1845: corresponding eigenfunction $f_n(x)\equiv \langle x|a_n\rangle$ is square
1846: normalizable:
1847: \begin{equation}
1848: (f_n,f_n)=\int_{-\infty}^{\infty}\rmd x \, |f_n(x)|^2 <\infty \, .
1849: \label{pspn}
1850: \end{equation}
1851: An eigenvalue $a$ belongs to the continuous part of the spectrum when its
1852: corresponding eigenfunction $f_a(x)\equiv \langle x|a\rangle$ is {\it not}
1853: square normalizable:
1854: \begin{equation}
1855: (f_a,f_a)=\int_{-\infty}^{\infty}\rmd x \, |f_a(x)|^2 =\infty \, .
1856: \label{pspE}
1857: \end{equation}
1858: In the latter case, one has to use the
1859: theory of distributions to ``normalize'' these states, e.g., delta function
1860: normalization:
1861: \begin{equation}
1862: (f_a,f_{a'})=\int_{-\infty}^{\infty}\rmd x \,
1863: \overline{f_a(x)} f_{a'}(x) =\delta (a-a') \, .
1864: \label{pspEdn}
1865: \end{equation}
1866: This Dirac delta normalization generalizes the Kronecker delta normalization
1867: of ``discrete'' states:
1868: \begin{equation}
1869: (f_n,f_{n'})=\int_{-\infty}^{\infty}\rmd x \,
1870: \overline{f_n(x)} f_{n'}(x) =\delta _{nn'} \, .
1871: \end{equation}
1872: Because they are square integrable, the ``discrete'' eigenvectors
1873: $f_n(x)\equiv \langle x|a_n\rangle$ can be
1874: interpreted in the usual way as probability amplitudes. But because they are
1875: {\it not} square integrable, the ``continuous'' eigenvectors
1876: $f_a(x)\equiv \langle x|a\rangle$ must be interpreted as ``kernels''
1877: of probability amplitudes, in the sense that when we multiply
1878: $\langle x|a\rangle$ by $\langle \varphi |x \rangle$ and then integrate, we
1879: obtain the density of probability amplitude $\langle \varphi |a\rangle$:
1880: \begin{equation}
1881: \langle \varphi |a\rangle = \int_{-\infty}^{\infty}\rmd x \,
1882: \langle \varphi |x \rangle \langle x|a\rangle \, .
1883: \end{equation}
1884: Thus, in particular, $\langle x|p\rangle$, $\langle x|x'\rangle$ and
1885: $\langle x|E^{\pm}\rangle _{\rm l,r}$ represent ``kernels''
1886: of probability amplitudes.
1887:
1888: Another way to interpret the bras and kets
1889: is in analogy to the plane waves of classical optics and classical
1890: electromagnetism. Plane waves $\rme^{\rmi kx}$ represent monochromatic light
1891: pulses of wave number $k$ and frequency (in vacuum) $w=kc$. Monochromatic light
1892: pulses are impossible to prepare experimentally; all that can be prepared
1893: are light pulses $\varphi (k)$ that have some wave-number spread. The
1894: corresponding pulse in the position representation, $\varphi (x)$, can be
1895: ``Fourier decomposed'' in terms of the monochromatic plane waves as
1896: \numparts
1897: \begin{equation}
1898: \varphi (x) = \frac{1}{\sqrt{2\pi }} \int \rmd k \,
1899: \rme^{\rmi kx}\varphi (k) \, ,
1900: \label{fourinnet}
1901: \end{equation}
1902: which in Dirac's notation becomes
1903: \begin{equation}
1904: \langle x|\varphi \rangle = \int \rmd k \,
1905: \langle x|k \rangle \langle k|\varphi \rangle \, .
1906: \end{equation}
1907: \endnumparts
1908: Thus, physically preparable pulses can be expanded in a Fourier
1909: integral by the unpreparable plane waves, the weights of the
1910: expansion being $\varphi (k)$. When $\varphi (k)$ is highly peaked around
1911: a particular wave number $k_0$, then the pulse can in general be
1912: represented for all practical purposes by a monochromatic plane wave
1913: $\rme ^{\rmi k_0x}$. Also, in finding out how a light pulse behaves under
1914: given conditions (e.g., reflection and refraction at a plane interface between
1915: two different media), we only have to find out how plane waves behave
1916: and, after that, by means of the Fourier expansion~(\ref{fourinnet}), we
1917: know how the light pulse $\varphi (x)$ behaves. Because obtaining the
1918: behavior of plane waves is somewhat easy, it is advantageous to use them to
1919: obtain the behavior of the whole pulse~\cite{USEFULNESSPW}.
1920:
1921: The quantum mechanical bras and kets can be interpreted in analogy
1922: to the classical plane waves. The eigenfunction
1923: $\langle x|p\rangle = \rme ^{\rmi px/\hbar} / \sqrt{2\pi \hbar}$ represents a
1924: particle of sharp momentum $p$; the eigenfunction
1925: $\langle x|x'\rangle = \delta (x-x')$ represents a particle sharply localized
1926: at $x'$; the monoenergetic eigenfunction
1927: $\langle x|E^{\pm}\rangle _{\rm l,r}$ represents a particle with well-defined
1928: energy $E$ (and with additional boundary conditions determined by the
1929: labels $\pm$ and ${\rm l,r}$). In complete analogy to the Fourier
1930: expansion of a light pulse by classical plane waves,
1931: Eq.~(\ref{fourinnet}),
1932: the eigenfunctions $\langle x|p\rangle$, $\langle x|x'\rangle$ and
1933: $\langle x|E^{\pm}\rangle _{\rm l,r}$ expand a wave function $\varphi$,
1934: see Eqs.~(\ref{resonidentPxphi})-(\ref{resonidentHxphi}). When the wave
1935: packet $\varphi (p)$ is highly peaked around a particular
1936: momentum $p_0$, then in general the approximation
1937: $\varphi (x) \sim \rme ^{\rmi p_0x/\hbar} / \sqrt{2\pi \hbar}$ holds for all
1938: practical purposes; when the wave packet $\varphi (x)$ is highly peaked around
1939: a particular position $x_0$, then in general the approximation
1940: $\varphi (x) \sim \delta (x-x_0)$ holds for all practical purposes; and when
1941: $\varphi (E)$ is highly peaked around a particular energy $E_0$,
1942: then in general the approximation
1943: $\varphi (x) \sim \langle x|E_0^{\pm}\rangle _{\rm l,r}$ holds for all
1944: practical purposes (up to the boundary conditions determined by the
1945: labels $\pm$ and ${\rm l,r}$). Thus, although in principle
1946: $\langle x|p\rangle$, $\langle x|x'\rangle$ and
1947: $\langle x|E^{\pm}\rangle _{\rm l,r}$ are impossible to prepare, in many
1948: practical situations they can give good approximations when the wave packet
1949: is well peaked around some particular values $p_0$, $x_0$, $E_0$ of the
1950: momentum, position and energy. Also, in finding out how a wave function
1951: behaves under given conditions (e.g., reflection and transmission off a
1952: potential barrier), all we have to find out is how the bras and kets behave
1953: and, after that, by means of the Dirac basis expansions, we know how the
1954: wave function $\varphi (x)$ behaves. Because obtaining the behavior of
1955: the bras and kets is somewhat easy, it is advantageous to use them to obtain
1956: the behavior of the whole wave function~\cite{USEFULNESSBK}.
1957:
1958: From the above discussion, it should be clear that
1959: there is a close analogy between classical Fourier methods and
1960: Dirac's formalism. In fact, one can say that Dirac's formalism is the extension
1961: of Fourier methods to Quantum Mechanics: Classical monochromatic plane waves
1962: correspond to the Dirac bras and kets; the light pulses correspond to the
1963: wave functions $\varphi$; the classical Fourier expansion corresponds to
1964: the Dirac basis expansions; the classical Fourier expansion provides the
1965: means to form light pulses out of a continuous linear superposition of
1966: monochromatic plane waves, and the Dirac basis expansions provide the
1967: means to form wave functions out of a continuous linear superposition of
1968: bras and kets; the classical uncertainty principle of Fourier
1969: Optics corresponds to the quantum uncertainty generated by the
1970: non-commutativity of two observables~\cite{DEBROGLIE}. However, although this
1971: analogy is very close from a formal point of view, there is a crucial
1972: difference from a conceptual point of view. To wit, whereas in the classical
1973: domain the solutions of the wave equations represent a physical wave, in
1974: Quantum Mechanics the solutions of the equations do {\it not} represent a
1975: physical object, but rather a probability amplitude---In Quantum Mechanics what
1976: is ``waving'' is probability.
1977:
1978:
1979:
1980:
1981:
1982:
1983:
1984: \section{Further considerations}
1985: \label{sec:gener}
1986:
1987:
1988: In Quantum Mechanics, the main objective is to obtain the probability of
1989: measuring an observable $A$ in a state $\varphi$. Within the Hilbert space
1990: setting, such probability can be obtained by means of the spectral
1991: measures ${\sf E}_a$ of $A$ (see, for example, Ref.~\cite{GALINDO}). These
1992: spectral measures satisfy
1993: \begin{equation}
1994: I = \int_{{\rm Sp}(A)} \rmd {\sf E}_a \,
1995: \end{equation}
1996: and
1997: \begin{equation}
1998: A = \int_{{\rm Sp}(A)} a \, \rmd {\sf E}_a \,
1999: \end{equation}
2000: Comparison of these equations with Eqs.~(\ref{introresiden}) and
2001: (\ref{introactionA}) yields
2002: \begin{equation}
2003: \rmd {\sf E}_a = |a\rangle \langle a| \, \rmd a \, .
2004: \label{factorizationofE}
2005: \end{equation}
2006: Thus, the RHS is able to ``factor out'' the Hilbert space spectral measures in
2007: terms of the bras and kets~\cite{FSHORT}. For the position, momentum and
2008: energy observables, Eq.~(\ref{factorizationofE}) reads as
2009: \begin{equation}
2010: \rmd {\sf E}_x = |x\rangle \langle x| \, \rmd x \, ,
2011: \end{equation}
2012: \begin{equation}
2013: \rmd {\sf E}_p = |p\rangle \langle p| \, \rmd p \, ,
2014: \end{equation}
2015: \begin{equation}
2016: \rmd {\sf E}_E = |E^{\pm} \rangle _{\rm l} \,
2017: _{\rm l}\langle ^{\pm} E| \, \rmd E +
2018: |E^{\pm} \rangle_{\rm r} \, _{\rm r} \langle ^{\pm} E| \, \rmd E
2019: \, .
2020: \label{spmesH}
2021: \end{equation}
2022: Although the spectral measures $\rmd {\sf E}_a$ associated with a given
2023: self-adjoint operator $A$ are unique, the factorization
2024: in terms of bras and kets is not. For example, as we can see from
2025: Eq.~(\ref{spmesH}), the spectral measures of our Hamiltonian can be written in
2026: terms of the basis $\{ |E^{+} \rangle _{\rm l,r} \}$ or the basis
2027: $\{ |E^{-} \rangle _{\rm l,r} \}$. From a physical point of view, those
2028: two basis are very different. As we saw in Sec.~\ref{sec:e1dsbp}, the basis
2029: $\{ |E^{+} \rangle _{\rm l,r} \}$
2030: represents the initial condition of an incoming particle, whereas the basis
2031: $\{ |E^{-} \rangle _{\rm l,r} \}$ represents the final condition of an
2032: outgoing particle. However, the spectral measures of the Hilbert space are
2033: insensitive to such difference, in contrast to the RHS, which can differentiate
2034: both cases. Therefore, when computing probability amplitudes, the RHS gives
2035: more precise information on how those probabilities are physically produced
2036: than the Hilbert space.
2037:
2038: In this paper, we have restricted our discussion to the simple,
2039: straightforward algebra
2040: of the 1D rectangular barrier. But, what about more complicated potentials? In
2041: general, the situation is not as easy. First, the theory of rigged Hilbert
2042: spaces as constructed by Gelfand and collaborators is based on the assumption
2043: that the space $\mathbf \Phi$ has a property
2044: called {\it nuclearity}~\cite{GELFAND,MAURIN}. However, it is not clear
2045: that one can always find a nuclear space $\mathbf \Phi$ that remains invariant
2046: under the action of the observables. Nevertheless, Roberts has shown that
2047: such $\mathbf \Phi$ exists when the potential is infinitely often
2048: differentiable except for a closed set of zero Lebesgue
2049: measure~\cite{ROBERTS}. Second, the problem of constructing the
2050: RHS becomes more involved when the observable $A$ is not
2051: cyclic~\cite{GELFAND}. And third, solving the eigenvalue
2052: equation of an arbitrary self-adjoint operator is rarely as easy as in our
2053: example.
2054:
2055:
2056:
2057:
2058: \section{Summary and conclusions}
2059: \label{sec:conclusions}
2060:
2061:
2062: We have used the 1D rectangular barrier model to see that, when the spectra of
2063: the observables have a continuous part, the natural setting for Quantum
2064: Mechanics is the rigged Hilbert space rather than just the Hilbert space. In
2065: particular, Dirac's bra-ket formalism is fully implemented by the rigged
2066: Hilbert space rather than just by the Hilbert space.
2067:
2068: We have explained the physical and mathematical meanings of each of
2069: the ingredients that form the rigged Hilbert space. Physically, the space
2070: $\mathbf \Phi \equiv \Sw$ is interpreted as the space of wave
2071: functions, since its elements can be
2072: associated well-defined, finite physical quantities, and algebraic operations
2073: such as commutation relations are well defined on
2074: $\mathbf \Phi$. Mathematically, $\mathbf \Phi$ is the space of
2075: test functions. The spaces ${\mathbf \Phi}^{\prime} \equiv \Swp$ and
2076: ${\mathbf \Phi}^{\times} \equiv \Swt$ contain respectively the bras and kets
2077: associated with the eigenvalues that lie in the continuous
2078: spectrum. Physically, the bras and kets are interpreted as ``kernels'' of
2079: probability amplitudes. Mathematically, the bras and kets are
2080: distributions. The following table summarizes the meanings of each space:
2081: \begin{center}
2082: \begin{tabular}{|c|l|l|}
2083: \hline
2084: {\sc Space} & \, {\sc Physical Meaning} & \, {\sc Mathematical Meaning}
2085: \, \\
2086: \hline
2087: \hline
2088: ${\mathbf \Phi}$ & \, Space of wave functions $\varphi$ \, &
2089: \, Space of test functions $\varphi$ \, \\
2090: ${\cal H}$ & \, Probability amplitudes & \, Hilbert space \\
2091: $\, \, {\mathbf \Phi}^{\times}$ & \, Space of kets $|a\rangle$ &
2092: \, Antidual space \\
2093: $\, {\mathbf \Phi}^{\prime}$ & \, Space of bras $\langle a|$ &
2094: \, Dual space \\
2095: \hline
2096: \end{tabular}
2097: \end{center}
2098:
2099: We have seen that, from a physical point of view, the rigged Hilbert
2100: space does not entail an extension of Quantum Mechanics, whereas, from a
2101: mathematical point of view, the rigged Hilbert space is an extension of
2102: the Hilbert space. Mathematically, the rigged Hilbert space
2103: arises when we equip the Hilbert space with distribution theory. Such
2104: equipment enables us to cope with singular objects such as bras and kets.
2105:
2106: We have also seen that formal expressions involving bras and kets
2107: must be understood as ``sandwiched'' by wave functions $\varphi$. Such
2108: ``sandwiching'' by $\varphi$'s is what controls the singular behavior of
2109: bras and kets. This is why mathematically the sandwiching by $\varphi$'s is
2110: so important and must always be implicitly assumed. In practice, we can
2111: freely apply the formal manipulations of Dirac's formalism with confidence,
2112: since such formal manipulations are justified by the rigged Hilbert space.
2113:
2114: We hope that this paper can serve as a pedagogical, enticing introduction to
2115: the rigged Hilbert space.
2116:
2117:
2118:
2119: \ack
2120:
2121: Research supported by the Basque Government through reintegration
2122: fellowship No.~BCI03.96, and by the University of the Basque Country
2123: through research project No.~9/UPV00039.310-15968/2004.
2124:
2125:
2126:
2127: \section*{References}
2128:
2129:
2130: \begin{thebibliography}{99}
2131:
2132:
2133:
2134:
2135: \bibitem{ATKINSON} D.~Atkinson, P.~W.~Johnson, \emph{Quantum Field
2136: Theory -- a Self-Contained Introduction},
2137: Rinton Press, Princeton (2002).
2138:
2139: \bibitem{BOGOLIOBOV} N.~N.~Bogolubov, A.~A.~Logunov, I.~T.~Todorov,
2140: \emph{Introduction to Axiomatic Quantum Field Theory}, Benjamin, Reading,
2141: Massachusetts (1975).
2142:
2143: \bibitem{BALLENTINE} L.~E.~Ballentine, \emph{Quantum Mechanics}, Prentice-Hall
2144: International, Inc., Englewood Cliffs, New Jersey (1990).
2145:
2146: \bibitem{BOHM} A.~Bohm, \emph{Quantum Mechanics: Foundations
2147: and Applications}, Springer-Verlag, New York (1994).
2148:
2149: \bibitem{BG} A.~Bohm and M.~Gadella, {\it Dirac kets, Gamow
2150: Vectors, and Gelfand Triplets}, Springer Lectures Notes in Physics Vol.~348,
2151: Springer, Berlin (1989).
2152:
2153: \bibitem{CAPRI} A.~Z.~Capri, \emph{Nonrelativistic Quantum
2154: Mechanics}, Benjamin, Menlo Park, California (1985).
2155:
2156: \bibitem{DUBIN} D.~A.~Dubin, M.~A.~Hennings, \emph{Quantum Mechanics, Algebras
2157: and Distributions}, Longman, Harlow (1990).
2158:
2159: \bibitem{GALINDO} A.~Galindo, P.~Pascual, \emph{Quantum Mechanics I},
2160: Springer-Verlag, Berlin (1990).
2161:
2162: \bibitem{KUKULIN} V.~I.~Kukulin, V.~M.~Krasnopol'sky, and J.~Horacek,
2163: \emph{Theory of resonances}, Kluwer Academic Publishers, Dordrecht (1989).
2164:
2165: \bibitem{FOCO} R.~de la Madrid, J.~Phys.~A: Math.~Gen.~{\bf 37}, 8129-8157 (2004); {\sf quant-ph/0407195}.
2166:
2167: \bibitem{DIS} R.~de la Madrid, ``Quantum mechanics in rigged
2168: Hilbert space language,'' Ph.D.~thesis, Universidad de Valladolid
2169: (2001). Available at \texttt{http://www.ehu.es/$\sim$wtbdemor/}.
2170:
2171: \bibitem{DIRAC} P.~A.~M.~Dirac, \emph{The principles of Quantum Mechanics},
2172: 3rd ed., Clarendon Press, Oxford (1947).
2173:
2174: \bibitem{VON} J.~von Neumann, \emph{Mathematische Grundlagen der
2175: Quantentheorie}, Springer, Berlin (1931); English translation by R.~T.~Beyer,
2176: {\it Mathematical Foundations of Quantum Mechanics}, Princeton University
2177: Press, Princeton (1955).
2178:
2179: \bibitem{QUOTEVONDIRAC} In Ref.~\cite{DIRAC}, page~40, Dirac states
2180: that ``{\it the bra and ket vectors that we now use form
2181: a more general space than a Hilbert space}.''
2182:
2183: In Ref.~\cite{VON}, page~viii, von Neumann states that ``{\it Dirac
2184: has given a representation of quantum mechanics which is scarcely to be
2185: surpassed in brevity and elegance,} [...].'' On pages~viii-ix, von Neumann
2186: says that ``{\it The method of Dirac, mentioned above, (and this is overlooked
2187: today in a great part of quantum mechanical literature, because of the clarity
2188: and elegance of the theory) in no way satisfies the
2189: requirements of mathematical rigor -- not even if these are reduced in a
2190: natural and proper fashion to the extent common elsewhere in theoretical
2191: physics.}'' On page~ix, von Neumann says that ``[...],{\it this requires
2192: the introduction of `improper' functions with self-contradictory
2193: properties. The insertion of such mathematical `fiction' is frequently
2194: necessary in Dirac's approach,}[...].'' Thus, essentially, although von
2195: Neumann recognizes the clarity and beauty of Dirac's formalism, he states
2196: very clearly that such formalism cannot be implemented within the framework
2197: of the Hilbert space.
2198:
2199: \bibitem{SCHWARTZ} L.~Schwartz, \emph{Th\'eory de Distributions}, Hermann,
2200: Paris (1950).
2201:
2202: \bibitem{GELFAND} I.~M.~Gelfand, N.~Y.~Vilenkin,
2203: \emph{Generalized Functions}, Vol.~IV, Academic Press, New York
2204: (1964).
2205:
2206: \bibitem{MAURIN} K.~Maurin, \emph{Generalized Eigenfunction Expansions and
2207: Unitary Representations of Topological Groups}, Polish Scientific
2208: Publishers, Warsaw (1968).
2209:
2210: \bibitem{CITEMAURIN} In Ref.~\cite{MAURIN}, page~7, Maurin states that
2211: ``{\it It seems to us that this is the formulation
2212: which was anticipated by Dirac in his classic
2213: monograph.''}
2214:
2215: \bibitem{ROBERTS} J.~E.~Roberts, J.~Math.~Phys.~{\bf 7}, 1097--1104 (1966);
2216: J.~E.~Roberts, Commun.~Math.~Phys.~{\bf 3}, 98--119
2217: (1966).
2218:
2219: \bibitem{ANTOINE}J.-P.~Antoine, J.~Math.~Phys.~{\bf 10}, 53--69 (1969);
2220: J.-P.~Antoine, J.~Math.~Phys.~{\bf 10}, 2276--2290 (1969).
2221:
2222: \bibitem{B60} A.~Bohm, ``The Rigged Hilbert Space in Quantum
2223: Mechanics,'' \emph{Boulder Lectures in Theoretical Physics, 1966}, Vol.~9A
2224: (Gordon and Breach, New York, 1967).
2225:
2226: \bibitem{QUOTEBALLENTINE} The following quotation, extracted from
2227: Ref.~\cite{BALLENTINE}, page~19, gives a clear idea of the status
2228: the RHS is achieving: {\it ``...rigged Hilbert space seems to
2229: be a more natural mathematical setting for quantum mechanics than Hilbert
2230: space.''}
2231:
2232: \bibitem{AT93} I.~Antoniou, S.~Tasaki, Int.~J.~Quant.~Chem.~{\bf 44},
2233: 425--474 (1993).
2234:
2235: \bibitem{SUCHANECKI} Z.~Suchanecki, I.~Antoniou, S.~Tasaki,
2236: O.~F.~Brandtlow, J.~Math.~Phys.~{\bf 37}, 5837--5847 (1996).
2237:
2238: \bibitem{DENSE} A subspace $S$ of $\cal H$ is dense in $\cal H$
2239: if we can approximate any element of $\cal H$ by an element of $S$
2240: as well as we wish. Thus, for any $f$ of $\cal H$ and for any small
2241: $\epsilon >0$, we can find a $\varphi$ in $S$ such that
2242: $\| f-\varphi \| < \epsilon$. In physical terms, this inequality means that
2243: we can replace $f$ by $\varphi$ within an accuracy $\epsilon$.
2244:
2245: \bibitem{FUNCTIONAL} A function $F: {\mathbf \Phi} \to {\mathbb C}$ is called
2246: a linear [respectively antilinear] functional over $\mathbf \Phi$ if for any
2247: complex numbers $\alpha , \beta$
2248: and for any $\varphi , \psi \in \mathbf \Phi$, it holds that
2249: $F(\alpha \varphi +\beta \psi)=\alpha F(\varphi) +\beta F(\psi )$
2250: [respectively
2251: $F(\alpha \varphi +\beta \psi)=\alpha ^* F(\varphi) +\beta ^*F(\psi )$].
2252:
2253: \bibitem{JPA02} R.~de la Madrid, J.~Phys.~A: Math.~Gen.~{\bf 35}, 319--342
2254: (2002); {\sf quant-ph/0110165}.
2255:
2256: \bibitem{FP02} R.~de la Madrid, A.~Bohm, and M.~Gadella, Fortsch.~Phys.~{\bf 50}, 185--216 (2002); {\sf quant-ph/0109154}.
2257:
2258: \bibitem{IJTP03} R.~de la Madrid, Int.~J.~Theor.~Phys.~{\bf 42}, 2441--2460
2259: (2003); {\sf quant-ph/0210167}.
2260:
2261: \bibitem{HSDEF} Strictly speaking, a Hilbert space possesses additional
2262: properties (e.g., it must be complete with respect to the topology induced by
2263: the scalar product). For a more technical definition of the Hilbert space,
2264: see for example Ref.~\cite{DIS}.
2265:
2266: \bibitem{UNB} An operator $A$ is bounded if there is some finite $K$ such
2267: that $\| Af \| <K \|f \|$ for all $f\in \cal H$, where $\| \ \|$ denotes
2268: the Hilbert space
2269: norm. When such $K$ does not exist, $A$ is said to be unbounded. For a
2270: detailed account of the properties of bounded and unbounded operators, see for
2271: example Ref.~\cite{DIS}.
2272:
2273: \bibitem{RS84} The mathematical reason why quantum mechanical unbounded
2274: operators cannot be defined on all the vectors of the Hilbert space can be
2275: found, for example, in Ref.~\cite{RS}, page~84.
2276:
2277: \bibitem{RS} M.~Reed, B.~Simon, ``Methods of modern mathematical physics,''
2278: vol.~I, Academic Press, Inc., New York (1972).
2279:
2280: \bibitem{INFENER} If we nevertheless insisted in for example calculating the
2281: expectation value~(\ref{exintrodispP}) for elements of $\cal H$ that are not
2282: in ${\cal D}(A)$, we would obtain an unphysical infinity value. For instance,
2283: if $A$ represents an unbounded Hamiltonian $H$, then the expectation
2284: value~(\ref{exintrodispP}) would be infinite for those $\varphi$ of $\cal H$
2285: that lie outside of ${\cal D}(H)$. Because they have infinite
2286: energy, those states do not represent physically preparable wave packets.
2287:
2288: \bibitem{SNHS} If they were in the Hilbert space, $|a\rangle$ and $\langle a|$
2289: would be square integrable, and $a$ would belong to the discrete spectrum.
2290:
2291: \bibitem{RS274} It is well known that Heisenberg's commutation relation
2292: necessarily implies that either $P$ or $Q$ is unbounded. See, for example,
2293: Ref.~\cite{RS}, page~274.
2294:
2295: \bibitem{ZERODERIV} The reason why the derivatives of $\varphi (x)$ must
2296: vanish at $x=a,b$ is that we want to be able to apply the Hamiltonian $H$
2297: as many times as we wish. Since repeated applications of $H$ to $\varphi (x)$
2298: involve the derivatives of $V(x)\varphi (x)$, and since $V(x)$ is
2299: discontinuous at $x=a,b$, the function $V(x)\varphi (x)$ is infinitely
2300: differentiable at $x=a,b$ only when the derivatives of $\varphi (x)$
2301: vanish at $x=a,b$. For more details, see Ref.~\cite{ROBERTS}. The vanishing
2302: of the derivatives of $\varphi (x)$ at $x=a,b$ must be viewed as a
2303: mathematical consequence of the unphysical sharpness of the discontinuities of
2304: the potential, rather than as a physical consequence of Quantum
2305: Mechanics. Note
2306: also that in standard numerical simulations, for example, Gaussian wave
2307: packets impinging on a rectangular barrier, one never sees that the wave
2308: packet vanishes at $x=a,b$. This is due to the fact that on a Gaussian wave
2309: packet, the Hamiltonian~(\ref{fdoph}) can only be applied once.
2310:
2311: \bibitem{CRVANISHES} We note that, when acting on elements $\varphi$ of
2312: $\Sw$, the commutator $[H,P] =\rmi \hbar \frac{\partial V}{\partial x}$
2313: reduces to $[H,P]=0$, due to the vanishing of the derivatives of $\varphi$
2314: at $x=a,b$.
2315:
2316: \bibitem{ONEONEROBERTS} We recall that some authors have erroneously claimed
2317: that ``there are more kets than bras''~\cite{ROBERTS}, and that therefore such
2318: one-to-one correspondence between bras and kets does not hold.
2319:
2320: \bibitem{EXTOHS} We can nevertheless extend
2321: Eqs.~(\ref{resonidentPxphi}) and (\ref{resonidentHxphi}) to the
2322: whole Hilbert space $L^2$ by a limiting procedure,
2323: although the
2324: resulting expansions do not involve the Dirac bras and kets any more, but
2325: simply the eigenfunctions of the differential operators.
2326:
2327: \bibitem{COHEN} C.~Cohen-Tannoudji, B.~Diu, and F.~Lalo\"e,
2328: {\it Quantum Mechanics}, Wiley, New York (1977).
2329:
2330: \bibitem{USEFULNESSPW} This is one of the major reasons why plane waves
2331: are so useful in practical calculations.
2332:
2333: \bibitem{USEFULNESSBK} This is one of the major reasons why bras and kets
2334: are so useful in practical calculations.
2335:
2336: \bibitem{DEBROGLIE} There are many other links between the classical and the
2337: quantum worlds, such as for example the de Broglie relation $p=\hbar k$, which
2338: entails a formal identity between the classical $\rme ^{\rmi kx}$ and
2339: the quantum $\rme ^{\rmi px/\hbar}$ plane waves.
2340:
2341: \bibitem{FSHORT} We recall that the direct integral decomposition of the
2342: Hilbert space falls short of such factorization, see Ref.~\cite{ANTOINE}.
2343:
2344:
2345:
2346: \end{thebibliography}
2347:
2348:
2349:
2350:
2351: \newpage
2352:
2353:
2354: \begin{figure}
2355: \begin{center}
2356: \epsfbox{pwp.eps}
2357: \end{center}
2358: \caption{Schematic representation of the eigenfunctions
2359: $\langle x|E^+\rangle _{\rm r}$, Fig.~1a, and
2360: $\langle x|E^+\rangle _{\rm l}$, Fig.~1b.}
2361: \label{fig:plus}
2362: \end{figure}
2363:
2364:
2365: \newpage
2366:
2367:
2368:
2369: \begin{figure}
2370: \begin{center}
2371: \epsfbox{pwm.eps}
2372: \end{center}
2373: \caption{Schematic representation of the eigenfunctions
2374: $\langle x|E^-\rangle _{\rm r}$, Fig.~2a, and
2375: $\langle x|E^-\rangle _{\rm l}$, Fig.~2b.}
2376: \label{fig:minus}
2377: \end{figure}
2378:
2379:
2380:
2381: \end{document}
2382:
2383:
2384:
2385: