1: \documentclass{elsart}
2: \textwidth 6.2in
3: \textheight 9in
4:
5: \usepackage{amssymb,epsfig,graphicx,amsmath}
6:
7:
8: \journal{Phys. Lett. A}
9: \begin{document}
10: \begin{frontmatter}
11:
12: \title{Resonances in the one-dimensional
13: Dirac equation in the presence of a point interaction and a constant
14: electric field}
15: \author[IVIC]{Luis Gonz\'alez-D\'{\i}az},
16: \ead{lugonzal@pion.ivic.ve}
17: \author[IVIC]{V\'{\i}ctor M. Villalba\corauthref{cor}}
18: \corauth[cor]{Corresponding author.} \ead{villalba@ivic.ve}
19:
20: \address[IVIC]{Centro de F\'{\i}sica
21: Instituto Venezolano de Investigaciones Cient\'{\i}ficas, IVIC \\
22: Apdo 21827, Caracas 1020-A, Venezuela}
23:
24:
25: \begin{abstract}
26: We show that the energy spectrum of the one-dimensional Dirac
27: equation in the presence of a spatial confining point interaction
28: exhibits a resonant behavior when one includes a weak electric
29: field. After solving the Dirac equation in terms of parabolic
30: cylinder functions and showing explicitly how the resonant behavior
31: depends on the sign and strength of the electric field, we derive an
32: approximate expression for the value of the resonance energy in
33: terms of the electric field and delta interaction strength.
34: \end{abstract}
35:
36: \begin{keyword}
37: Relativistic electron, Strong Fields, Dirac equation, Resonances
38: \PACS 03.65.Pm, 03.65.Nk, 03.65.Ge
39: \end{keyword}
40: \end{frontmatter}
41:
42:
43: \section{Introduction}
44:
45: Supercritical effects are perhaps one of the most interesting
46: phenomena associated with the charged vacuum in the presence of
47: strong electric fields \cite{Greiner,Rafelski} The study of
48: supercritical effects induced by strong vector potentials goes back
49: to the pioneering works of Pieper and Greiner \cite{Pieper},
50: Zeldovich and Popov \cite{Zeldovich}
51: among others. The idea behind
52: supercriticality is to have positron emission induced by the
53: presence of very strong attractive vector potentials. The phenomenon
54: can be described as follows: the energy level of an unoccupied bound
55: state dives into the negative energy continuum. i.e., an electron of
56: the Dirac sea is trapped by the potential, leaving a positron that
57: escapes to infinity. \ The
58: electric field responsible for supercriticality should be stronger than $%
59: 2m_{e}c^{2}$, which is the value of the gap between the negative and
60: positive energy continua. \ Such strong electric fields could be
61: produced in heavy-ion collisions \cite{Greiner,Greiner2}. A rigorous
62: mathematical study of the behavior of the Dirac energy levels near
63: the continuum spectrum and the problem of spontaneous pair creation
64: has been carried out by \u{S}{e}ba \cite{PS:86}, Klaus \cite{Klaus}
65: and Nenciu \cite{Nenciu} among others.
66:
67: In order to get a deeper understanding of the mechanism responsible
68: for supercriticality and for the resonant peaks appearing in the
69: energy spectrum when supercritical fields are present, we proceed to
70: work with a vector point interaction in the presence of a constant
71: electric field. Point interactions potentials may be used to
72: approximate, in a simple way, more complex short-ranged potentials.
73: Among the advantages of working with confining delta interactions we
74: should mention that, they only possess a single bound state and the
75: treatment of the interacting potential reduces to a boundary
76: condition. The study of bound states of the relativistic wave
77: equation in the presence of point interactions is a problem that has
78: been carefully discussed in the literature
79: \cite{Sutherland,McKellar,McKellar2,FD:89,Albeverio}. The
80: one-dimensional Dirac equation in the presence of a vector point
81: delta interaction has also been a subject of study in the search of
82: supercritical effects induced by attractive potentials
83: \cite{Greiner,Loewe}. Soon after the publication of the paper by
84: Loewe and Sanhueza \cite{Loewe}, Nogami et al \cite {Nogami} pointed
85: out that supercritical effects are also absent in a class of
86: non-local separable potentials in one dimension.
87:
88: Since we are interested in studying the mechanism of positron
89: production by supercritical fields, we proceed to analyze the
90: resonant behavior of the energy when a bound state dives into the
91: negative continuum. This resonant behavior is associated with the
92: appearance of simple poles of the resolvent on the second sheet at
93: a position very near the real axis \cite{Reed}.
94:
95:
96: The method of complex eigenvalues (Gamow vectors) was introduced in
97: quantum mechanics by Gamow \cite{GG:28} in connection with the
98: theory of Alpha decay. Titchmarsch \cite{ET:58} and Barut
99: \cite{WB:62} demonstrated an application, in the framework of
100: non-relativistic quantum mechanics, for the Gamow vector method to
101: the problem of an attractive delta interaction $\delta (r)$ of strength $%
102: -\cot \alpha $ with a weak electric field term associated with the
103: potential $V(r)=-\lambda r.$ They found that the Schr\"{o}dinger
104: equation with a weak electric field exhibits a continuous spectrum
105: from $-\infty $ to $+\infty ,$ and a resonance at $E^{'}$ in the
106: vicinity
107: of $E_{{0}}-\frac{1}{2}\lambda \tan \alpha ,$ where $%
108: E_{0}=-\cot ^{2}\alpha $ is the energy of the unperturbed state. The
109: eigenvalue $E^{'}$
110: is a complex number in the lower half plane. \ According to Barut, the Schr%
111: \"{o}dinger equation with the potential $-\lambda r$ describes a
112: system that tries to form a bound state that ``dissolves itself" in
113: the presence of the continuous spectrum. In this Letter, using \ the
114: idea developed by Titchmarsch \cite{ET:58}, we find the energy
115: spectrum of the one-dimensional Dirac equation with boundary
116: conditions associated with a vector Dirac delta interaction and a
117: constant electric field whose strength is weak, and therefore it
118: produces a perturbative effect on the delta energy spectrum. We find
119: that in this case the energy spectrum exhibits a resonance due to
120: supercriticality.
121:
122: The Letter is structured as follows: In Section 2, we solve the
123: one-dimensional Dirac equation in the presence of an attractive
124: $\delta$ potential and a constant electric field. In Section 3, we
125: compute the energy resonances and show how they depend on the
126: electric field strength. We also derive an approximate analytic
127: expression for the energy resonances. Finally, in Section 4 we
128: summarize our conclusions.
129:
130: \section{The one-dimensional Dirac equation}
131:
132: In this section we will consider the 1+1 Dirac equation in the presence of
133: the attractive vector point interaction potential represented by $%
134: eV(x)=-g_{v}\delta (x),$ and a constant electric field associated
135: with the potential $eV(x)=\lambda x$. \ The Dirac equation,
136: expressed in natural units\ ($\hslash =c=1$) can be written in the
137: form \cite{Greiner1}
138: \begin{equation}
139: \left( i\gamma ^{\mu }(\frac{\partial }{\partial x^{\mu }}-ieA_{\mu
140: })-m\right) \Psi =0, \label{1}
141: \end{equation}
142: where $A_{\mu}$ is the vector potential, $e$ is the charge and $m$
143: is the mass of the electron. The Dirac matrices $\gamma^{\mu}$
144: satisfy the commutation relation $\left\{ \gamma
145: ^{\mu },\gamma ^{\nu }\right\} =2\eta ^{\mu \nu }$ with $\eta ^{\mu \nu }=%
146: \mathrm{diag}(1,-1).$ Since we are working in 1+1 dimensions, we choose to
147: work in a two-dimensional representation of the \ Dirac matrices
148: \begin{equation}
149: \gamma ^{0}=\sigma _{3},\gamma ^{1}=-i\sigma _{2} \label{2}.
150: \end{equation}
151: Substituting the representation matrix representation (\ref{2}) into
152: Eq. (\ref{1}), and taking into account that the potential
153: interaction does not depend on time, we obtain
154:
155: \begin{equation}
156: \{-i\sigma _{{1}}\frac{d}{dx}+\left( \lambda x-E\right) +m\sigma _{{3}}\}%
157: \mathsf{X}(x)=0, \label{dirac}
158: \end{equation}
159: with $\Psi =\sigma _{3}\mathsf{X.}$, and
160: \begin{equation}
161: \mathsf{X}(x)=
162: \begin{pmatrix}
163: X_{{1}} \\
164: X_{{2}}
165: \end{pmatrix}
166: ,
167: \end{equation}
168: with the boundary conditions at $x=0$
169: \begin{equation}
170: \begin{split}
171: X_{{1}}(0^{{+}})& =X_{{1}}(0^{{-}})\cos {g_{{v}}}-iX_{{2}}(0^{{-}})\hspace{%
172: 0.1cm}\mathrm{\sin }{g_{{v}}}, \\
173: X_{{2}}(0^{{+}})& =-iX_{{1}}(0^{{-}})\hspace{0.1cm}\mathrm{\sin }{g_{{v}}}+X_{%
174: {2}}(0^{{-}})\cos {g_{{v}}}.
175: \end{split}
176: \label{dominio}
177: \end{equation}
178: The above conditions (\ref{dominio}) describe a the point vector
179: potential interaction of strength $g_{{v}}$ \cite{FD:89}.
180:
181: Equation (\ref{dirac}) is equivalent to the system of equations
182: \begin{equation}
183: \left( m+\lambda x-E\right) X_{{1}}-i\frac{dX_{2}}{dx}=0,
184: \label{pdirac}
185: \end{equation}
186: \begin{equation}
187: i\frac{dX_{1}}{dx}+\left( m-\lambda x+E\right) X_{{2}}=0,
188: \label{sdirac}
189: \end{equation}
190: Introducing the new functions $\Omega _{{1}}$ and $\Omega _{{2}}$
191: \begin{equation}
192: X_{{1}}=\Omega _{{1}}+i\Omega _{{2}},\quad X_{{2}}=\Omega _{{1}}-i\Omega _{{2%
193: }}, \label{rela}
194: \end{equation}
195: we obtain that the system of equations (\ref{pdirac}) -(\ref{sdirac})
196: reduces to the form
197: \begin{align}
198: \frac{d\Omega _{1}}{dx}+i\left( \lambda x-E\right) \Omega _{{1}}-m\Omega _{{2%
199: }}& =0, \label{podirac} \\
200: \frac{d\Omega _{{2}}}{dx}-i\left( \lambda x-E\right) \Omega _{{2}}-m\Omega _{%
201: {1}}& =0, \label{sodirac}
202: \end{align}
203: which is more tractable in the search of exact solutions. Substituting (\ref
204: {podirac}) into (\ref{sodirac}) we obtain the second-order differential
205: equation
206: \begin{equation}
207: \frac{d^{2}\Omega _{{1}}}{dx^{2}}\,+\{i\lambda +\left( \lambda
208: x-E\right) ^{2}-m^{2}\}\,\Omega _{{1}}=0. \label{ecua}
209: \end{equation}
210:
211: Looking at the asymptotic behavior of the parabolic cylinder
212: functions $D_{\nu}(z)$ \cite{HB:53} we obtain that the regular
213: solutions, for $\lambda>0$, of Eq. (\ref{ecua}) belonging to
214: $\mathcal{L}^{2}(-\infty ,0)$ and $\mathcal{L}^{2}(0,\infty )$
215: ($Im\,E>0$), respectively are
216: \begin{equation}
217: \begin{split}
218: \Omega _{{1}}^{-}(x)& =AD_{{-\rho -1}}\left( \sqrt{\frac{2}{\lambda }}e^{{-i%
219: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) \\
220: \Omega _{1}^{+}(x)& =BD_{\rho }\left( \sqrt{\frac{2}{\lambda }}e^{{i\frac{%
221: \pi }{4}}}\left( \lambda \,x-E\right) \right) ,
222: \end{split}
223: \label{sol1}
224: \end{equation}
225: where $D_{{\rho }}$ y $D_{{-\rho -1}}$ are parabolic cylinder
226: functions \cite{HB:53}, $\rho =\frac{im^{{2}}}{2\lambda }$, \ and
227: $A$ and $B$ are constants.
228:
229: Inserting (\ref{sol1}) into (\ref{podirac}) and using the recurrence
230: relations for the parabolic cylinder functions \cite{HB:53}, we
231: obtain
232: \begin{equation}
233: \begin{split}
234: \Omega _{2}^{-}(x)& =i\frac{\sqrt{2\lambda }}{m}e^{{i\frac{\pi }{4}}}AD_{{%
235: -\rho }}\left( \sqrt{\frac{2}{\lambda }}e^{{-i\frac{\pi }{4}}}\left( \lambda
236: \,x-E\right) \right), \\
237: \Omega _{2}^{+}(x)& =i\frac{m}{\sqrt{2\lambda }}e^{{i\frac{\pi }{4}}}BD_{{%
238: \rho -1}}\left( \sqrt{\frac{2}{\lambda }}e^{{i\frac{\pi }{4}}}\left(
239: \lambda \,x-E\right) \right).
240: \end{split}
241: \label{sol2}
242: \end{equation}
243: From (\ref{sol1}), (\ref{sol2}) and (\ref{rela}), we have that the
244: components of the spinor solution for $x<0$ are
245:
246: \begin{equation}
247: \begin{split}
248: X_{1}^{-}(x)& =A\left[ D_{{-\rho -1}}\left( \sqrt{\frac{2}{\lambda }}e^{{-i%
249: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) -\frac{\sqrt{2\lambda }}{%
250: m}e^{{i\frac{\pi }{4}}}D_{{-\rho }}\left( \sqrt{\frac{2}{\lambda }}e^{{-i%
251: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) \right], \\
252: X_{{2}}^{-}(x)& =A\left[ D_{{-\rho -1}}\left( \sqrt{\frac{2}{\lambda }}e^{{-i%
253: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) +\frac{\sqrt{2\lambda }}{%
254: m}e^{{i\frac{\pi }{4}}}D_{{-\rho }}\left( \sqrt{\frac{2}{\lambda }}e^{{-i%
255: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) \right],
256: \end{split}
257: \label{izq}
258: \end{equation}
259: analogously, we have that, for $x>0$, the spinor $\mathsf{X}$ has
260: the components:
261:
262: \begin{equation}
263: \begin{split}
264: X_{{1}}^{+}(x)& =B\left[ D_{{\rho }}\left( \sqrt{\frac{2}{\lambda }}e^{{i%
265: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) -\frac{m}{\sqrt{2\lambda
266: }}e^{{i\frac{\pi }{4}}}D_{{\rho -1}}\left( \sqrt{\frac{2}{\lambda }}e^{{i%
267: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) \right], \\
268: X_{{2}}^{+}(x)& =B\left[ D_{{\rho }}\left( \sqrt{\frac{2}{\lambda }}e^{{i%
269: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) +\frac{m}{\sqrt{2\lambda
270: }}e^{{i\frac{\pi }{4}}}D_{{\rho -1}}\left( \sqrt{\frac{2}{\lambda }}e^{{i%
271: \frac{\pi }{4}}}\left( \lambda \,x-E\right) \right) \right].
272: \end{split}
273: \label{der}
274: \end{equation}
275:
276:
277:
278: Inserting (\ref{izq}) and (\ref{der}) in (\ref{dominio}), we obtain that the
279: equation for the energy eigenvalues $E$ has the form:
280: \begin{equation}
281: \begin{split}
282: 2\lambda D_{{\rho }}\left( \sqrt{\frac{2}{\lambda }}e^{{i\frac{\pi }{4}}%
283: }E\right) & D_{{-\rho }}\left( {-\sqrt{\frac{2}{\lambda }}}e^{{-i\frac{\pi }{%
284: 4}}}E\right) \\
285: & -m^{{2}}D_{{\rho -1}}\left( \sqrt{\frac{2}{\lambda }}e^{{i\frac{\pi }{4}}%
286: }E\right) D_{{-\rho -1}}\left( {-\sqrt{\frac{2}{\lambda }}}e^{{-i\frac{\pi }{%
287: 4}}}E\right) e^{{-2ig_{{v}}}}=0.
288: \end{split}
289: \label{urav}
290: \end{equation}
291:
292: The regular solutions of Eq. (\ref{ecua}), for $\lambda<0$, can be
293: obtained after interchanging the roles of $\Omega_{1}^{-}(x)$ and
294: $\Omega_{1}^{+}(x)$ and of $\Omega_{2}^{-}(x)$ and
295: $\Omega_{2}^{+}(x)$ in Eq. (\ref{sol1}) and Eq. (\ref{sol2})
296: respectively; therefore Eq. (\ref{urav}) permits us to compute the
297: energy resonances for positive as well as negative values of the
298: electric field strength $\lambda$
299:
300: \section{Energy resonances}
301:
302: It is not an easy task to solve Eq. (\ref{urav}), nevertheless it is
303: possible to derive an analytic approximation for the energy
304: eigenvalues. In order to do this approximation we consider the
305: effect of the electric field in the vicinity of zero, i.e., where
306: the delta interaction is located. We insert the spinor solutions
307: obtained using this approach into (\ref {dominio}), obtaining in
308: this way an eigenvalue equation, whose roots are a
309: good approximation to those of (\ref{urav}). The approximate equation for $%
310: \Omega _{{1}}$ has the form
311: \begin{equation}
312: \frac{d^{2}\Omega _{{1}}}{dx^{2}}\,+\{i\lambda
313: +E^{2}-m^{2}\}\,\Omega _{{1}}=0. \label{a}
314: \end{equation}
315: The solution of Eq. (\ref{a}) that, for $x>0$, and $\lambda>0$
316: exhibit a regular asymptotic behavior is
317: \begin{equation}
318: \Omega _{1}^{+}=A\exp (-i\sqrt{i\lambda +E^{2}-m^{2}}x). \label{b}
319: \end{equation}
320: Substituting the solution (\ref{b}) into Eq. (\ref{podirac}) we obtain
321: \begin{equation}
322: \Omega _{2}^{+}=-iA\frac{\sqrt{i\lambda +E^{2}-m^{2}}+E}{m}\exp
323: (-i\sqrt{i\lambda +E^{2}-m^{2}}x), \label{d}
324: \end{equation}
325: and, analogously, we have that for $x<0,$ the corresponding
326: solutions are
327: \begin{equation}
328: \Omega _{1}^{-}=B\exp (i\sqrt{i\lambda +E^{2}-m^{2}}x), \label{c}
329: \end{equation}
330: \begin{equation}
331: \Omega _{2}^{-}=-iB\frac{-\sqrt{i\lambda +E^{2}-m^{2}}+E}{m}\exp
332: (i\sqrt{i\lambda +E^{2}-m^{2}}x). \label{e}
333: \end{equation}
334: Substituting the right $\Omega _{1,2}^{+}$ and left $\Omega
335: _{1,2}^{-}$\ spinor components into Eq. (\ref{dominio}), we find
336: that, for $g_{v}=-\pi $, the energy state sinks into the negative
337: energy continuum, exhibiting a resonant behavior that depends on the
338: electric field strength as:
339:
340: \begin{equation}
341: E\left( \lambda \right) \approx -\left( m+\frac{\lambda ^{{2}}}{8m^{{3}}}%
342: \right) +i\frac{\lambda }{2m}, \label{spectrum}
343: \end{equation}
344: where we have assumed that $\lambda$ is positive and small compared
345: to $m$. It should be noticed that, when we turn off the electric
346: perturbation, we recover the
347: energy eigenvalue $E(0)=-m$ for a point vector interaction with strength $%
348: g_{{v}}=-\pi $ \cite{FD:89}. It is worth mentioning the vanishing
349: of the first order perturbative correction to the supercritical
350: energy $E=-m$ as a result of the presence of the electric field
351: interaction $\lambda$. This behavior can also be observed looking at
352: the real part of the resonant energy given by Eq. (\ref{spectrum}).
353: An approximate expression for the energy resonances for negative
354: values of $\lambda$ can be obtained after substituting $\lambda$ by
355: $-\lambda$ in Eq. (\ref{spectrum}).
356:
357: We can try to obtain an approximate solution to the full Eq.
358: (\ref{ecua}) after making a series expansion of the function $\
359: i\lambda +\left( \lambda x-E\right) ^{2}-m^{2}$ around zero. In this
360: way, we have that the approximate function $\Omega _{{1}}$
361: satisfies the second order differential equation
362: \begin{equation}
363: \frac{d^{2}\Omega _{{1}}}{dx^{2}}\,+\{i\lambda +E^{2}-2\lambda
364: Ex-m^{2}\}\,\Omega _{{1}}=0. \label{f}
365: \end{equation}
366: The regular solutions $\Omega _{{1}}^{-}$ and for $x<0,$ and $\Omega _{{1}%
367: }^{+}$ for $x>0$ can be expressed, for $\lambda>0$, in terms of Hankel functions $%
368: H_{1/3}^{1}(z)$ and $H_{1/3}^{2}(z)$ \cite{Abramowitz} as follows
369: \begin{equation}
370: \Omega _{{1}}^{-}=Az^{1/3}H_{1/3}^{1}(z), \label{g}
371: \end{equation}
372: \begin{equation}
373: \Omega _{{1}}^{+}=Bz^{1/3}H_{1/3}^{2}(z), \label{h}
374: \end{equation}
375: where $A$ and $B$ are constants and $z$ is given by the expression
376: \begin{equation}
377: z=-\frac{\left( i\lambda +E^{2}-m^{2}-2E\lambda x\right) ^{3/2}}{3\lambda E}
378: \label{i}
379: \end{equation}
380: Substituting the solutions (\ref{g}) and (\ref{h}) for $\Omega
381: _{{1}}^{-}$ and $\Omega _{{1}}^{+}$ into Eq. (\ref{podirac}) we
382: obtain the corresponding expressions for $\Omega _{{2}}^{-}$ and
383: $\Omega _{{2}}^{+}.$ \ Taking into account the relation between
384: $X_{1}$, $X_{2}$ and $\Omega _{1},$ $\Omega _{2} $ (\ref{rela}), and
385: inserting the lower and upper components of $\mathsf{X}$ into the
386: boundary condition (\ref{dominio}), we find that an approximate
387: equation for the energy spectrum is:
388:
389: \begin{equation}
390: \left( X_{{2}}(0^{{-}})X_{{1}}(0^{{+}})-X_{{2}}(0^{{+}})X_{{1}}(0^{{-}%
391: })\right) \cos g_{v}+i\left( X_{{2}}(0^{{+}})X_{{2}}(0^{{+}})-X_{{1}}(0^{{+}%
392: })X_{{1}}(0^{{-}})\right) \sin g_{v}=0 \label{energ}
393: \end{equation}
394: It is worth noting that the energy spectrum obtained from Eq. (\ref{energ}%
395: ) with $\Omega _{{1}}^{-}$ and $\Omega _{{1}}^{+}$ given by Eqs.
396: (\ref{g}) and (\ref{h}) gives a result comparable to the one
397: obtained using the relation Eq. (\ref{spectrum})
398: \begin{figure}[tbp]
399: \begin{center}
400: \includegraphics[width=10cm]{espectro.eps}
401: \end{center}
402: \caption{Dependence of the real and imaginary parts of the energy on
403: the electric field strength for $0<\lambda<0.1m^2$.}
404: \end{figure}
405: Fig. 1 shows how the real and imaginary part of the energy depend on
406: the parameter $\lambda $ when a point vector interaction of
407: strength $g_{{v}}=-\pi $ is applied. The energy exhibits a resonant
408: behavior produced by the electric field.
409:
410: The inverse of the imaginary part of Eq. (\ref{spectrum}) permits
411: one to determine the time \cite{Newton,Goldberger} that the state
412: stays in negative continuum. The stronger the electric field
413: perturbation, the shorter the mean life of the resonance
414: \cite{Newton}. An interesting result that can be observed from Eq.
415: (\ref{spectrum}) and it is also depicted in Fig. 1 is that the
416: imaginary part of the resonant energy behaves as a square root with
417: respect to the real part of the energy.
418:
419: \section{Concluding remarks}
420:
421:
422: In this article we have shown that the presence of a weak electric
423: field plays a crucial role in the appearance of a resonant energy
424: state in the one-dimensional Dirac equation in the field of an
425: attractive vector delta interaction. This phenomenon is analogous to
426: the Stark effect \cite{Galindo}, where a perturbative constant
427: electric field produces a shifting and broadening of the energy
428: levels of the hydrogen atom. It also parallels the autoionization
429: process, where the presence electron-electron interaction term
430: induces a self-ionization process and levels decay like in the Auger
431: effect \cite{Reed,Galindo}. Such analogies seem to indicate that,
432: for resonant effects, the shape of the perturbative potential is
433: sometimes more important than the strength of the potential field.
434:
435: We have also shown that, the one-dimensional Dirac equation in the
436: field of an attractive vector delta interaction exhibits a resonant
437: supercritical behavior when one introduces a potential of the form
438: $\lambda x$. Since this effect is not present when one consider the
439: sole vector interaction $-g_{v}\delta (x)$ \cite{FD:89}, we conclude
440: that it is necessary the presence of the electric field in order to
441: sink the delta bound state into the negative continuum.
442:
443: \section{Acknowledgments}
444: We thank Dr. Ernesto Medina for reading and improving the
445: manuscript. We also wish to express our gratitude to the anonymous
446: referees for their critical remarks. This work was supported by
447: FONACIT under project G-2001000712.
448:
449: \begin{thebibliography}{99}
450: \bibitem{Greiner} W. Greiner, B. M\"{u}ller, and J. Rafelski, \textit{%
451: Quantum Electrodynamics of Strong fields} (Springer Verlag, Berlin,
452: 1985).
453:
454: \bibitem{Rafelski} J. Rafelski, P. Fulcher and A. Klein Phys. Rep. \textbf{%
455: 38} (1978) 227.
456:
457: \bibitem{Pieper} W. Pieper, and W. Greiner, Z. Phys. \textbf{218,}
458: (1969) 327.
459:
460: \bibitem{Zeldovich} Y. B. Zeldovich, and V. S. Popov, Sov. Phys. Usp.
461: \textbf{14} (1972) 673.
462:
463: \bibitem{Greiner2} W. Greiner and J. Reinhardt, Physica Scripta \textbf{T56},
464: (1995) 203.
465:
466: \bibitem{PS:86} P. \u{S}{e}ba, The Dirac eigenvalues near upper and
467: lower continuum, Preprint JINR E2-86-808, Dubna 1986.
468:
469: \bibitem{Klaus} M. Klaus, J. f\"{u}r Mathematik, \textbf{362}, 197 (1985).
470:
471: \bibitem{Nenciu} G. Nenciu, Commun. Math. Phys. \textbf{109} 303 (1987).
472:
473:
474: \bibitem{Sutherland} B. Sutherland and D. C. Mattis, Phys. Rev. A \textbf{24%
475: }, (1981) 1194.
476:
477: \bibitem{McKellar} B. H. J. McKellar and G. J. Stephenson, Phys. Rev. A.
478: \textbf{36} (1987) 2566.
479:
480: \bibitem{McKellar2} B. H. J. McKellar and G. J. Stephenson, Phys. Rev. C.
481: \textbf{35} (1987) 2262.
482:
483: \bibitem{FD:89} F. Dom\'{i}nguez-Adame and E. Maci\'{a}, J. Phys. A: Math.
484: Gen. \textbf{22} (1989) L419.
485:
486: \bibitem{Albeverio} S. Albeverio, F. Gesztesy, R. H{\o}egh-Krohn and
487: H. Holden, \textit{Solvable Models in Quantum Mechanics}(AMS Chelsea
488: Publishing, Providence, Rhode Island 2005)
489:
490: \bibitem{Loewe} M. Loewe and M. Sanhueza, J. Phys. A: Math.
491: Gen. \textbf{23}, (1990) 553.
492:
493: \bibitem{Nogami} Y. Nogami, N. P.Parent and F. M. Tokoyama, J. Phys. A:
494: Math. Gen \textbf{23}, (1990) 5667.
495:
496: \bibitem{Reed} M. Reed and B. Simon, \textit{Methods of Modern Mathematical
497: Physics: IV Analysis of Operators} (Academic Press, New York 1978).
498:
499:
500: \bibitem{GG:28} G. Gamow, Z. Phys. \textbf{51}, (1928) 204. \newline
501: E.U. Condon, and R. W. Gurney, Phys. Rev. \textbf{33}, (1929) 127.
502:
503: \bibitem{ET:58} E. C. Titchmarsh, \textit{Eigenfunction Expansions
504: Associated with Second Order Differential Equations}, Part II.
505: (Oxford University Press, Oxford 1958).
506:
507: \bibitem{WB:62} W. E. Brittin, \textit{Lectures in Theoretical Physics},
508: Vol. IV. Interscience Publishers, 1962, Pag. 460.
509:
510: \bibitem{Greiner1} W. Greiner, Relativistic Quantum Mechanics, Wave
511: equations, (Springer, New York 1990).
512:
513: \bibitem{HB:53} \textit{Higher Transcendental Functions}, Vol. II. McGraw -
514: Hill Book Company, Inc., 1953.
515:
516:
517: \bibitem{Abramowitz} M. Abramowitz and I. Stegun, \textit{Handbook of
518: Integrals, Series and Products} (Dover, New York 1965).
519:
520: \bibitem{Newton} R. G. Newton, \textit{Scattering Theory of Waves and
521: Particles}, (Dover, New York, 2002)
522:
523: \bibitem{Goldberger} M. Goldberger and K. Watson, \textit{Collision Theory}
524: (Dover, New York, 2004)
525:
526: \bibitem{Galindo} A. Galindo, and P. Pascual, \textit{Quantum Mechanics
527: II} (Springer-Verlag, Heidelberg, 1991).
528:
529: \end{thebibliography}
530:
531: \end{document}
532: