hep-th0304010/EBM.tex
1: \documentclass{elsart}
2: \usepackage{epsf,amssymb}
3: \journal{Physics Letters A}
4: \begin{document}
5: \newcommand{\bra}{\langle}
6: \newcommand{\ket}{\rangle}
7: \newcommand{\eq}[2]{\begin{equation}\label{#1} #2 \end{equation}}
8: \newcommand{\tbf}[1]{{\bf #1}}
9: \newcommand{\tit}[1]{{\it #1}}
10: \newcommand{\intRR}{\int\limits_{-\infty}^{\infty}}
11: \newcommand{\intR}{\int\limits_{0}^{\infty}}
12: \newcommand{\intRt}{\int_{0}^{\infty}}
13: \newcommand{\LRD}[1]{\frac{{{\displaystyle\leftrightarrow}
14: \atop {\displaystyle\partial}}}{\partial #1}}
15: \newcommand{\lrd}[1]{\stackrel{\displaystyle \leftrightarrow}
16: {\displaystyle\partial_{#1}}}
17: \newcommand{\lrx}[1]{\stackrel{\displaystyle \leftrightarrow}
18: {\displaystyle #1}}
19: \newcommand{\lrX}[1]{\stackrel{\displaystyle \longleftrightarrow}
20: {\displaystyle #1}}
21: \newcommand{\diff}[1]{\partial /{\partial #1}}
22: \newcommand{\diffm}[2]{\partial#1/\partial #2}
23: \newcommand{\Diff}[1]{\frac{\partial}{\partial #1}}
24: \newcommand{\hc}[1]{#1^{\dagger}}
25: \newcommand{\DiffM}[2]{\frac{\partial #1}{\partial #2}}
26: \newcommand{\SDiffM}[2]{\frac{\partial^2 #1}{\partial {#2}^2}}
27: \newcommand{\ds}{\displaystyle}
28: \newcommand{\sign}{\,{\rm sign}\,}
29: %\textwidth=21cm
30: \textheight=23cm
31: \begin{frontmatter}
32: \title{Pair creation by homogeneous electric field from the
33: point of view of an accelerated observer}
34: \author{N.B. Narozhny}\ead{narozhny@theor.mephi.ru},
35: \author{V.D. Mur},
36: \author{A.M. Fedotov\corauthref{cor}},
37: \corauth[cor]{Corresponding author} \ead{fedotov@cea.ru}
38: \address{Moscow Engineering Physics Institute (state university), 115409 Moscow, Russia}
39: \date{}
40: 
41: \begin{abstract}
42: With a special choice of gauge the operator of the
43: Klein-Fock-Gordon equation in homogeneous electric field respects
44: boost symmetry. Using this symmetry we obtain solutions for the
45: scalar massive field equation in such a background (boost modes in
46: the electric field). We calculate the spectrum of particles
47: created by the electric field, as seen by an accelerated observer
48: at spatial infinity of the right wedge of Minkowski spacetime. It
49: is shown that the spectrum and the total number of created pairs
50: measured by a remote uniformly accelerated observer in Minkowski
51: spacetime are precisely the same as for inertial observers.
52: 
53: \end{abstract}
54: 
55: \begin{keyword}
56: Homogeneous electric field\sep pair creation\sep boost symmetry\sep
57: uniformly accelerated detector
58: \PACS 03.70.+{\bf k}\sep
59: 04.70.Dy
60: \end{keyword}
61: \end{frontmatter}
62: 
63: %\parindent=0.5cm
64: 
65: 
66: \textbf{1.} The Schwinger process of pair creation by a constant
67: electric field from vacuum \cite{Schwinger} have been studied in
68: details (using different methods and different gauges of the
69: external field) in the early 70-s \cite{N,NN,VSP,GMF,TMF,PLI}. In
70: this literature the quantum states of created particles were
71: labelled either by values of generalized momentum (non-stationary
72: gauge) or energy (stationary gauge). However the constant electric
73: field permits another symmetry which is based on invariance of the
74: field with respect to Lorentz transformations along its direction.
75: If the field is directed along the $z$-axis, this symmetry is
76: generated by the boost operator
77: $\mathcal{B}=i(t\diff{z}+z\diff{t})$. The presence of boost
78: symmetry in the problem makes it very similar to the problem of
79: particle creation by an eternal black hole \cite{Unruh,BD}. Thus
80: treatment of the Schwinger process in terms of boost modes is of
81: great importance for the quantum field theory in curved spacetime.
82: 
83: Besides providing a possible algorithm for investigation of the
84: process of particle creation in the background of eternal black
85: hole, the boost modes based approach to the Schwinger problem is
86: of independent interest. It is very natural to use Milne--Rindler
87: coordinate map (see, e.g., \cite{BD}) in problems with boost
88: symmetry. Physically transition to these coordinates means
89: transition to a non-inertial reference frame, namely, to a
90: uniformly accelerated reference frame in the right and left wedges
91: of Minkowski spacetime (MS). We know the average number of
92: particles created by a constant electric field as it is seen in
93: inertial reference frames. Therefore calculation of this quantity
94: measured by an accelerated observer will clear up the question of
95: how inertial forces influence the spectrum and the total number of
96: created particles.
97: 
98: \textbf{2.} We consider the Schwinger problem for massive scalar
99: particles where for the electric field $E$ we chose a special
100: gauge \eq{gauge}{A_{\mu}=-(E/2)\epsilon_{\mu\nu}x^{\nu}, \quad
101: x^{\mu}=\{t,z\}, \,\; \epsilon_{\mu\nu}=-\epsilon_{\nu\mu}, \;
102: \epsilon_{tz}=1,} which provides invariance of the
103: Klein-Fock-Gordon (KFG) equation under boost
104: transformations\footnote{In the current paper we use natural units
105: $\hbar=c=1$ and metric with the signature $(+,-)$. Note that the
106: proposed gauge is very similar to the well-known gauge
107: $\mathbf{A}=(1/2)\mathbf{H}\times\mathbf{r}$, which provides
108: explicit rotational symmetry of the Schr\"odinger equation in the
109: presence of a uniform magnetic field.}. The latter equation in the
110: gauge (\ref{gauge}) takes the form \eq{KFG}{
111: \left(\square-eE\mathcal{B}+\frac14
112: e^2E^2x_+x_-+m^2\right)\phi(x)=0,} where $e$ and $m$ are the
113: electric charge and mass of the particle, $x_{\pm}=t \pm
114: z,\;\square = 4\partial^2/\partial x_+
115: \partial x_-$ and the boost generator
116: $\mathcal{B}=\mathcal{D}_{(+)} -\mathcal{D}_{(-)}, \;
117: \mathcal{D}_{(\pm)} =ix_{\pm}
118: \partial/\partial x_{\pm}.$ One can easily check that the
119: KFG operator at the l.h.s. of Eq.(\ref{KFG}) commutes with
120: $\mathcal{B}$. Therefore solutions for equation (\ref{KFG}) can be
121: labelled by eigenvalues $\kappa$ of the boost generator
122: $\mathcal{B}$. We will call the solutions to Eq.(\ref{KFG})
123: $\phi_{\kappa}(x)$ satisfying the equation
124: $\mathcal{B}\phi_{\kappa}=\kappa\phi_{\kappa}$ "the boost modes in
125: the electric field".
126: 
127: The commuting boost $\mathcal{B}$ and dilatation
128: $\mathcal{D}=\mathcal{D}_{(+)} +\mathcal{D}_{(-)}$ vector fields
129: generate local pseudo\-eucli\-dean coordinates which admit
130: separation of variables in Eq. (\ref{KFG}). In the future ($F$)
131: and past ($P$) wedges of MS, see Fig.\ref{scheme}, these are Milne
132: coordinates \eq{Mln}{\tau=\pm m(x_+x_-)^{1/2}, \;
133: \sigma=1/2\ln(x_+/x_-),} while in the Rindler right ($R$) and left
134: ($L$) wedges, Rindler coordinates
135: \eq{Rnd}{\eta=1/2\ln(-x_+/x_-),\;\rho=\pm m(-x_+x_-)^{1/2},}
136: (compare to \cite{BD}). The four regions $R$, $L$, $P$ and $F$ are
137: bounded by horizons which belong to the light cone $x^2\equiv
138: t^2-z^2=0$, and (together with the horizons and the origin $x=0$)
139: cover the whole MS. Since the coordinate lines $\rho = const$ in
140: the Rindler wedges $R$ and $L$ coincide with the world lines of
141: uniformly accelerated observers in MS the resulting reference
142: frame can be considered as a uniformly accelerated one.
143: 
144: \textbf{3.} It was shown in Refs. \cite{NN,FIAN} that classical
145: solutions for the relevant field equation contain complete
146: information on the pair creation process. Here we will follow
147: these papers in our analysis instead of constructing a consistent
148: second quantized theory. We will assume that the initial state of
149: the field in $P$-wedge at far past ($\tau \rightarrow -\infty,$ or
150: $x_{\pm}\rightarrow -\infty$) is a vacuum state.
151: 
152: It follows from \cite{NN,FIAN} that there exist two non-equivalent
153: complete sets of modes in the Schwinger problem. It was shown in
154: \cite{FIAN} that complete information on the pair creation process
155: can be derived from any solution belonging to any of the sets. It
156: is convenient for us to use the solution which in the $P$-wedge
157: takes the form \eq{bm}{
158: \begin{array}{c}\ds
159: \phi_{\kappa}(x)=\frac{1}{\sqrt{2}}(-mx_-)^{i\kappa}
160: \exp\left\{-\frac{i}{4}\,\mathcal{E}m^2x_+
161: x_--\frac{\pi}{4\,\mathcal{E}} \right\}\times
162: \\ \\ \ds
163: \times
164: \Psi\left(\frac12+\frac{i}{2\,\mathcal{E}}\,
165: ,\,1+i\kappa\,,\frac{i}{2}\, \mathcal{E}m^2x_+x_-\right),
166: \end{array}}
167: where $\mathcal{E}=E/E_{cr}$ is the electric field strength in the
168: units of critical QED field $E_{cr}=m^2c^3/e\hbar$, and
169: $\Psi(a,c,\zeta)$ is the Tricomi function \cite{BE1}. The explicit
170: form for the solution (\ref{bm}) in other wedges may be obtained
171: by analytic continuation in such a way that for transitions from
172: one wedge to another one should use substitutions $-x_{\pm}
173: \rightarrow x_{\pm}e^{-i\pi},$ (compare to Ref. \cite{PRD}). At
174: the limit $E \rightarrow 0$ the modes (\ref{bm}) coincide up to a
175: phase factor with the boost modes $\;\Psi^*_{-\kappa}(x)$ for
176: empty MS, see Ref. \cite{PRD}.
177: 
178: Let us first analyze the particle creation process in the
179: $P$-wedge. Using the asymptotic expression for the Tricomi
180: function \cite{BE1} we can see that at far past
181: ($x_{\pm}\to-\infty$) the solution (\ref{bm}) acquires
182: semiclassical form \eq{past}{
183: \phi_{\kappa}(x)=(2\,|\mathcal{D}S_{\kappa}|)^{-1/2}
184: \exp(iS_{\kappa}),} where
185: $$
186: \ds S_{\kappa}(x_+,x_-)=-\frac{\mathcal{E}}{4}\,m^2x_+x_-
187: -\frac{1}{2\,\mathcal{E}}\ln\left(m^2x_+x_-\right)+\kappa\ln(-mx_-)
188: +\ldots
189: $$
190: is the classical action of a particle with a charge $e$ moving in
191: the electric field (\ref{gauge}). This solution is normalized by
192: the condition $\mathcal{J}^{\,\tau}=-e$ which means that the
193: charge per unit interval of $\sigma$ in the local reference frame
194: is equal to $-e$. Hence, in the in-region the solution (\ref{bm})
195: describes an incoming antiparticle accelerated by the electric
196: field $E$. Here
197: $$
198: \mathcal{J}^{\mu}=e\sqrt{-g}g^{\mu\nu}
199: \left(i\phi_{\kappa}^*\LRD{x^{\nu}}\varphi_{\kappa}-
200: 2eA_{\nu}\phi_{\kappa}^*\phi_{\kappa}\right),
201: $$
202: is the vector density of current, $\tau$-component of which for
203: the wedge $P$ can be represented in the form
204: $\;\mathcal{J}^{\,\tau}=-e\phi_{\kappa}^*\!\!\lrx{\mathcal{D}}\!\!
205: \phi_{\kappa}$.
206: 
207: 
208: Near the light cone $x_+x_-=0$ the solution (\ref{bm}) reduces to
209: \cite{BE1} \eq{lc}{
210: \begin{array}{r}\ds
211: \phi_{\kappa}(x)\sim \frac1{\sqrt{2}}e^{-\pi /4\,\mathcal{E}}
212: \frac{\Gamma(-i\kappa)}{\Gamma(1/2+i/2\,\mathcal{E}-i\kappa)}(-mx_-)^{i\kappa}
213: +\\ \\ \ds + \frac1{\sqrt{2}}e^{\pi\kappa/2-\pi/4\,\mathcal{E}}
214: \frac{\Gamma(i\kappa)}{\Gamma(1/2+i/2\,\mathcal{E})}
215: \left(-\frac{\mathcal{E}}{2}mx_+\right)^{-i\kappa}.
216: \end{array}}
217: 
218: The first term at the r.h.s. is obviously a wave propagating to
219: the right, while the second term is a wave propagating to the
220: left, both waves travelling with the speed of light. Near the
221: horizon the charge densities carried by the right- and left-going
222: waves are respectively given by \eq{j+}{\begin{array}{cc}\ds
223: \mathcal{J}^{\,\tau}_{(-)}=-ie\phi_{\kappa}^*\lrx{D}_{(-)}\phi_{\kappa}=
224: \,e\frac{1+e^{2\pi\kappa-\pi/\mathcal{E}}}{e^{2\pi\kappa}-1},\label{j-}\\
225: \\ \ds \mathcal{J}^{\,\tau}_{(+)}
226: =-ie\phi_{\kappa}^*\lrx{D}_{(+)}\phi_{\kappa}=
227: -e\frac{1+e^{-\pi/\mathcal{E}}}{1-e^{-2\pi\kappa}}.
228: \end{array}}
229: 
230: It is easily seen that $\mathcal{J}^{\,\tau}_{(-)}>0$ and
231: $\mathcal{J}^{\,\tau}_{(+)}<0$ if $\kappa>0$. This means that the
232: right-going wave describes the flux of particles created by the
233: electric field, while the left-going wave describes the flux of
234: created antiparticles together with the incoming one. Thus, these
235: charge densities should be written in the form, compare
236: \cite{NN,FIAN} $\mathcal{J}^{\,\tau}_{(-)}=en_{\kappa}^{(P)}$,
237: $\mathcal{J}^{\,\tau}_{(+)}=-e\left(1+n_{\kappa}^{(P)}\right)$,
238: where $n_{\kappa}^{(P)}$ is the number of pairs in the $\kappa$-th
239: mode, created by the electric field from vacuum in the $P$-wedge.
240: Obviously, we have: \eq{kappa>0}{
241: n_{\kappa}^{(P)}=\frac{1+e^{2\pi\kappa-\pi/\mathcal{E}}}{e^{2\pi\kappa}-1},
242: \quad\kappa>0.} If $\kappa<0$, $\mathcal{J}^{\,\tau}_{(-)}<0$ and
243: $\mathcal{J}^{\,\tau}_{(+)}>0$. Therefore in this case the
244: right-going wave describes the flux of antiparticles, while the
245: left-going wave describes the flux of particles, and the
246: corresponding charge densities should be written in the form
247: $\mathcal{J}^{\,\tau}_{(-)}=-e\left(1+n_{\kappa}^{(P)}\right)$,\,
248: $\mathcal{J}^{\,\tau}_{(+)}=e n_{\kappa}^{(P)}$ with \eq{kappa<0}{
249: n_{\kappa}^{(P)}=\frac{1+e^{-\pi/\mathcal{E}}}{e^{2\pi|\kappa|}-1},
250: \quad\kappa<0.} Thus we conclude that for $\kappa>0$ particles
251: created in the $P$-wedge travel to the $R$-wedge and the
252: antiparticles to the $L$-wedge, while for negative values of
253: $\kappa$ we have the opposite situation.
254: 
255: \textbf{4.} Now let us consider what happens in the $R$-wedge. In
256: terms of Rindler coordinates (\ref{Rnd}) the KFG equation
257: (\ref{KFG}) after separation of the time-like variable $\eta,\;
258: \phi_{\kappa}=e^{-i\kappa\eta}\varphi_{\kappa}(\rho),$ formally
259: coincides with the stationary Schr\"odinger equation with respect
260: to the independent variable $u=\ln{\rho}$  \eq{RW_eq}{
261: \left[-\frac{d^2}{du^2}+
262: U_{\kappa}\right]\varphi_{\kappa}=\kappa^2\varphi_{\kappa},} where
263: the effective potential reads \eq{eff_pot}{
264: U_{\kappa}(\rho)=(1-\mathcal{E}\kappa)\rho^2-\frac14\,\mathcal{E}^2\rho^4,
265: \quad \rho=e^u.} If $\kappa\ge 1/\mathcal{E}$, the effective
266: potential is monotonously decreasing with $\rho^2$. If
267: $\kappa<1/\mathcal{E}$, then $U_{\kappa}(\rho)$ is a potential of
268: barrier type with maximum $U_{\kappa}^{(m)}$ at $\rho=\rho_m$
269: \eq{max}{ \rho_m^2=2\mathcal{E}^{-1}(\mathcal{E}^{-1}-\kappa),
270: \quad U_{\kappa}^{(m)}=(\mathcal{E}^{-1}-\kappa)^2.} However real
271: solutions for the equation $U_{\kappa}(\rho)=\kappa^2$ exist only
272: for $\kappa<1/2\,\mathcal{E}$. It means that the values
273: $\kappa>1/2\,\mathcal{E}$ correspond to above-barrier scattering,
274: while for $\kappa<1/2\,\mathcal{E}$ the Eq.(\ref{RW_eq}) describes
275: the sub-barrier tunnelling. The classical turning points of the
276: potential for the sub-barrier situation are located at
277: \eq{tp}{\rho_{\pm}=\left\{\begin{array}{ll}\mathcal{E}^{-1}
278: (1\pm\sqrt{1-2\kappa\, \mathcal{E}}\,),& \quad
279: 0<\kappa<1/2\,\mathcal{E}\,, \\
280: \mathcal{E}^{-1}(\sqrt{1+2|\kappa|\,\mathcal{E}}\pm 1), &\quad
281: \kappa<0.\end{array}\right.}We will again base our analysis on the
282: solution (\ref{bm}) though analytically continued to the
283: $R$-wedge. It consists of a superposition of in- and outgoing
284: waves near horizons ($u\rightarrow -\infty$) while at the right
285: spatial infinity of the wedge ($u\rightarrow \infty$) reduces to a
286: semiclassical right-going wave. The transmission $D_{\kappa}$ and
287: reflection $R_{\kappa}$ coefficients can be obtained by the
288: standard quantum mechanical procedure and are equal to \eq{DR}{
289: D_{\kappa}=\frac{1-e^{-2\pi\kappa}}{1+e^{\pi/\mathcal{E}-2\pi\kappa}}\,,\quad
290: R_{\kappa}=1-D_{\kappa}=
291: \frac{1+e^{-\pi/\mathcal{E}}}{1+e^{2\pi\kappa-\pi/\mathcal{E}}}\,.}
292: The scattering interpretation of the coefficients (\ref{DR}) is
293: consistent for all positive values of $\kappa$. However, it
294: follows from Eq.(\ref{DR}) that $D_{\kappa}<0$ and $R_{\kappa}>1$
295: if $\kappa<0$. This phenomena is known as the Klein paradox
296: \cite{Klein} (see also \cite{NN,FIAN}) and is explained by the
297: effect of pair creation by the external field. The fact that pairs
298: are created only with $\kappa<0$ is easy to understand if one
299: notes that, as it follows from Eqs.(\ref{tp}), the work of the
300: external field at the sub-barrier region is equal to
301: $$A=m\,\mathcal{E}(\rho_+-\rho_-)=\left\{
302: \begin{array}{ll}
303: 2m\sqrt{1-2\kappa\,\mathcal{E}},&
304: \quad0<1/\kappa<1/2\,\mathcal{E},
305: \\  2m, & \quad\kappa<0,
306: \end{array}
307: \right.$$ and is sufficient for pair creation only for $\kappa<0.$
308: According to Ref.\cite{FIAN} (see also \cite{Hund}) the number of
309: created pairs in this situation is given by the absolute value of
310: the transmission coefficient \eq{n_R}{
311: n_{\kappa}^{(R)}=\theta(-\kappa)|D_{\kappa}|=\theta(-\kappa)
312: \frac{e^{2\pi|\kappa|}-1}{e^{\pi/\mathcal{E}+2\pi|\kappa|}+1}.} It
313: is important that this result is valid in the $R$-wedge only under
314: assumption that the external field creates pairs from {\it
315: vacuum}. In the second quantized theory it means that there exists
316: the amplitude of vacuum -- vacuum transition $_r\bra 0|0\ket_l$,
317: where $|0\ket_l$ is the state of the field without right-going
318: particles on the left of the barrier and $|0\ket_r$ is the state
319: without left-going particles on the right of it. It is clear that
320: in our case the state $|0\ket_l$ could be prepared only if no
321: particles arrive to the $R$-wedge from outside. In other words,
322: the validity of the spectrum (\ref{n_R}) requires zero boundary
323: condition for the field in $R$-wedge at $u\rightarrow -\infty$, or
324: $\rho\rightarrow 0$ (compare with \cite{PRD}). However there is no
325: reason for implementation of such boundary condition in MS. We
326: will consider now what is happening in the $R$-wedge with regard
327: for the fact that the $R$-wedge is a part of MS but not a separate
328: spacetime.
329: 
330: \textbf{5.} The total picture can be reconstructed from what we
331: already know about $P$- and $R$-wedges. Physical phenomena in the
332: $F$- ($L$-) wedge are identical to those in $P$- ($R$-) wedge due
333: to the symmetry of the problem with respect to time reversal (or
334: space inversion). This total picture is shown in
335: Fig.~\ref{scheme}.
336: \begin{figure}[hhttbb]
337: \epsfxsize=13cm \centerline{\epsffile{figure2.eps}}
338: \caption{Spacetime location of particle creation regions and
339: direction of particle (empty arrows), antiparticle (filled arrows)
340: fluxes for different signs of $\kappa$.} \label{scheme}
341: \end{figure}
342: In $P$- and $F$-wedges the external field creates pairs
343: characterized by $\kappa$ of both signs. The particles created in
344: the $F$ wedge stay within it, while the particles created in the
345: $P$-wedge go to the $R$- and $L$-wedges. If $\kappa>0$, the
346: particles get to the $R$-wedge, while the antiparticles to the
347: $L$-wedge. The incoming particles (antiparticles) are partially
348: reflected there by the effective potential (\ref{eff_pot}) and
349: arrive to the $F$-wedge. Those particles (antiparticles) which
350: penetrate through the potential go to the right infinity of the
351: $R$-wedge (left infinity of the $L$-wedge), where they can be
352: detected by a remote uniformly accelerated (Rindler) observer who
353: is moving along a hyperbolic world line $\rho={\rm const}$. In
354: particular, the Rindler observer in the right wedge will detect
355: created particles and will not observe antiparticles. Since the
356: particles behind the scattering barrier in the $R$-wedge are
357: continuously accelerated by the electric field, they are
358: travelling with almost the speed of light. It is clear that the
359: number of particles with $\kappa>0$ which are crossing the world
360: line of a remote Rindler observer per unit his local time is given
361: by \eq{J_p}{
362: J_{\kappa}=D_{\kappa}n_{\kappa}^{(P)}=e^{-\pi/\mathcal{E}},\quad
363: \kappa>0.}
364: 
365: If $\kappa<0$, then the particles created in the wedge $P$ get to
366: the $L$-wedge, while the antiparticles to the $R$-wedge. All of
367: them are reflected by the effective potential (\ref{eff_pot}) and
368: follow to the $F$-wedge. But the electric field creates pairs with
369: $\kappa<0$ directly in the $R$ and $L$ wedges. The particles
370: created in the wedge $R$ (antiparticles in $L$) move to the right
371: (left) spatial infinity, while the antiparticles (particles) go to
372: the $F$-wedge. Hence, independently of the sign of $\kappa$ the
373: right Rindler observer always detects only particles. But if at
374: $\kappa>0$ these particles originate from the $P$-wedge, at
375: $\kappa<0$ they are created at the sub-barrier region in the
376: $R$-wedge itself. The important point is that the number of
377: particles created in the wedge $R$ (and $L$) in the current
378: situation differs from the expression (\ref{n_R}) because the
379: latter is related to particle production from vacuum. However,
380: pairs in the wedge $R$ are created in the presence of
381: antiparticles arrived from the $P$-wedge. Since scalar particles
382: satisfy the Bose statistics, presence of antiparticles results in
383: stimulation of pair production in the $R$-wedge. Therefore the
384: number of particles with $\kappa<0$ actually produced in the wedge
385: $R$ and equal to the number of particles detected by the remote
386: Rindler observer per unit his local time is given by (see, e.g.,
387: Ref.~\cite{FIAN}) \eq{J_n}{
388: J_{\kappa}=n_{\kappa}^{(R)}\left(1+n_{\kappa}^{(P)}\right)=
389: e^{-\pi/\mathcal{E}},\quad \kappa<0,} with the quantity
390: $n_{\kappa}^{(R)}$ defined in Eq.(\ref{n_R}). Thus, the spectrum
391: of particles measured by the Rindler observer is independent of
392: quantum number $\kappa$.
393: 
394: It is worth noting that this result can also be obtained directly
395: from the solution (\ref{bm}). Indeed, the asymptotic form
396: (\ref{past}) of this solution can be analytically  continued to
397: the right spatial infinity of the $R$-wedge along the path in the
398: complex space $\{x_{+},x_{-}\}$, which goes round the horizon
399: $x_{+}=0$ at sufficiently large distance, such that the
400: semiclassical approximation remains valid along all the path. It
401: is easy to see, that after such continuation the classical action
402: $S_{\kappa}(x_{+},x_{-})$ acquires constant imaginary part $\Delta
403: S_{\kappa}=i\pi/2\,\mathcal{E}$. Since the solution (\ref{bm})
404: corresponds to an incoming antiparticle at far past in wedge $P$
405: and anti\-particles do not penetrate through the potential
406: $U_{\kappa}$, see Fig. \ref{scheme}, the asymptotic of solution
407: (\ref{bm}) describes at the right spatial infinity of the
408: $R$-wedge only particles created by the electric field. Hence, the
409: asymptotic of the current in this region is equal to $eJ_{\kappa}$
410: and after simple calculations we arrive to the result
411: $J_{\kappa}=\exp\{-2\,{\rm Im}\,S_{\kappa}\}$ in full agreement
412: with Eqs. (\ref{J_p}), (\ref{J_n}).
413: 
414: Our analysis can be generalized to the 3-dimensional case. Indeed,
415: it is clear that the 3-dimensional KFG equation after separation
416: of variables orthogonal to the direction of the electric field
417: reduces to the equation for the 1-dimensional problem with the
418: effective mass $\sqrt{m^2+p_{\perp}^2}$ instead of $m$. Hence the
419: number of particles with quantum numbers
420: $(\kappa,\mathbf{p}_{\perp})$ detected by remote Rindler observer
421: per unit local time $\eta$ can be obtained from Eqs. (\ref{J_p}),
422: (\ref{J_n}) by substitution $m \rightarrow \sqrt{m^2+p_{\perp}^2}$
423: \eq{3d}{ J_{\kappa,\mathbf{p}_{\perp}}=
424: \exp\left(-\frac{\pi\left(m^2+p_{\perp}^2\right)}{eE}\right).}
425: 
426: \textbf{6.} Finally, we summarize our results and present the
427: conclusions. In this paper we have analyzed the process of pair
428: creation by a constant homogeneous electric field in MS viewed
429: from a uniformly accelerated reference frame. We have shown that
430: the spectrum of particles (antiparticles) measured by a remote
431: uniformly accelerated observer is precisely the same as the one
432: measured by a conventional inertial observer (see, e.g., Refs.
433: \cite{NN,FIAN}). This means that inertial forces cannot create
434: particles. In particular, no particles can be observed at spatial
435: infinity in the absence of the electric field with respect to both
436: reference frames. In our opinion, this conclusion is an additional
437: explicit demonstration of non-existence of the so-called Unruh
438: effect \cite{Unruh}. Previously, we have come to the same
439: conclusion in Refs. \cite{PRD,Ann,det} on the basis of quite
440: different arguments.
441: 
442: Let us emphasize that the number of pairs (\ref{n_R}) created in
443: the $R$-wedge, when it is considered as a separate spacetime with
444: a special condition imposed at its boundary, is absolutely
445: different from the one (\ref{J_p}), (\ref{J_n}) measured by a
446: remote uniformly accelerating observer in MS where communication
447: between the wedges is possible. This result provides additional
448: evidence for the necessity of boundary conditions for consistent
449: field quantization on incomplete manifolds, see Refs.
450: \cite{Ann,PRD}.
451: 
452: The authors wish to thank L.B. Okun for constant interest to our
453: work. We appreciate support from the Russian Fund for Basic
454: Research and Ministry of Education of Russian Federation.
455: 
456: \begin{thebibliography}{99}
457: \bibitem{Schwinger}J. Schwinger, Phys. Rev. 82 (1951) 664.
458: \bibitem{N}A.I. Nikishov, Zh. Eksp. Teor. Fiz. 57 (1969) 1210
459: (Sov. Phys. JETP  30 (1969) 660).
460: \bibitem{NN}N.B. Narozhny, A.I. Nikishov, Yad. Fiz. 11
461: (1970) 1072 (Sov. Journ. Nucl. Phys. 11 (1970) 596).
462: \bibitem{VSP}V.S. Popov, Zh. Eksp. Teor. Fiz. 62 (1972) 1248
463: (Sov. Phys. JETP  35 (1972) 569).
464: \bibitem{GMF}A.A. Grib, V.M. Mostepanenko, V.M. Frolov, Teor.
465: Mat. Fiz. 13 (1972) 377.
466: \bibitem{TMF}N.B. Narozhny, A.I. Nikishov, Teor. Mat. Fiz. 26
467: (1976) 16 (Theor. Math. Phys. 26 (1976) 9).
468: \bibitem{PLI}N.B. Narozhny, A.I. Nikishov, in \tit{Issues in Intense-Field
469: Quantum Electrodynamics}, Proc. Lebedev Phys. Inst. Vol. 168, p.
470: 226. Ed. by V.L. Ginzburg (Commack, N.Y. Nova Science, 1987).
471: \bibitem{Unruh}W.G. Unruh. Phys. Rev. D 14 (1976) 870.
472: \bibitem{BD}N.D. Birrell and P.C.W. Davies, \tit{Quantum
473: Fields in Curved Space} (Cambridge University Press, Cambridge,
474: 1982).
475: \bibitem{FIAN}A.I. Nikishov, \tit{Issues in Intense-Field Quantum
476: Electrodynamics}, Trudy Lebedev Phys. Inst. Vol. 111, (Moskva,
477: Nauka, 1979).
478: \bibitem{BE1}H. Bateman and A. Erdelyi,
479: \tit{Higher Transcendential Functions}, Vol. 1 (Mc Graw-Hill, New
480: York, 1953).
481: \bibitem{PRD}N.B. Narozhny, A.M. Fedotov, B.M. Karnakov, V.D. Mur,
482: and V.A. Belinskii, Phys. Rev. D 65 (2002) 025004.
483: \bibitem{Klein}O. Klein, Zs. Phys. 53 (1929) 157; F. Sauter, ibid.
484: 69 (1931) 742; 73 (1931) 547; A.I. Nikishov, Nucl. Phys. B21 (1970) 346.
485: \bibitem{Hund}F. Hund, Zs. Phys. 117 (1940) 1.
486: \bibitem{Ann}N. Narozhny, A. Fedotov, B. Karnakov, V. Mur, and
487: V. Belinskii, Ann. Phys. (Leipzig) 9 (2000) 199.
488: \bibitem{det}A.M. Fedotov, N.B. Narozhny, V.D. Mur, and V.A. Belinskii,
489: Phys. Lett. A, 305 (2002) 211.
490: \end{thebibliography}
491: \end{document}
492: 
493: 
494: