physics0605038/HEI.TEX
1: \documentclass{article}
2: %\documentclass[12pt,a4]{scrartcl}
3: %\usepackage{german}
4: \usepackage{amssymb}
5: \usepackage{amsmath}
6: \usepackage[dvips]{graphicx}
7: \usepackage{float}
8: \usepackage{afterpage}
9: \usepackage{eufrak}
10: 
11:   \begin{document}
12:   \title{Consequences of Dirac's Theory of the Positron}
13:   \author{W. Heisenberg and H. Euler in Leipzig$^1$}
14:   \date{22. December 1935}
15:   \maketitle
16: 
17: 
18: \begin{abstract}
19: According to Dirac's theory of the positron,
20: an electromagnetic
21: field tends to create pairs of
22: particles which leads to
23: a change of Maxwell's
24: equations in the vacuum. These changes are
25: calculated in the special case that no real electrons or
26: positrons are present and the field
27: varies little over a Compton wavelength.
28: The resulting effective Lagrangian of the field reads:
29: \begin{equation*}
30: \begin{split}
31: \scriptstyle
32: \mathfrak{L}=
33: & \scriptstyle \frac{1}{2}(\mathfrak{E}^2 -\mathfrak{B}^2) + \frac{e^2}{\hbar c}
34: \int_0^{\infty} e^{-\eta}\frac{d\eta}{\eta ^3}\biggl\{\scriptstyle i\eta ^2 (\mathfrak{EB}) \cdot
35: \frac{\ \cos{(\frac{\eta}{|\mathfrak{E}_k|} \sqrt{\mathfrak{E}^2-\mathfrak{B}^2+2i(\mathfrak{EB})})} + \text{conj.}}
36: {\cos{(\frac{\eta}{|\mathfrak{E}_k|} \sqrt{\mathfrak{E}^2-\mathfrak{B}^2+2i(\mathfrak{EB})})} - \text{conj.}}\\
37: & \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad
38: \scriptstyle +|\mathfrak{E}_k | ^2+\frac{\eta ^2}{3}(\mathfrak{B}^2-\mathfrak{E}^2)
39: \biggr\}\\
40: & \qquad \qquad \scriptstyle \mathfrak{E},\mathfrak{B}\quad \text{ field strengths}\\
41: & \qquad \qquad \scriptstyle |\mathfrak{E}_k |= \frac{m^2 c^3}{e \hbar}=\frac{1}{137}\frac{e}{(e^2/m c^2)^2} =\quad \text{ critical field strengths}\\
42: \end{split}
43: \end{equation*}
44: The expansion terms in small fields (compared to $\mathfrak E$ ) describe light-light scattering.
45: The simplest term is already known from perturbation theory.
46: For large fields, the equations derived here
47: differ strongly from
48: Maxwell's equations. Our  equations
49: will be compared to those proposed by Born.
50: \\ [3mm]
51: German title:
52: ``Folgerungen aus der Diracschen Theorie des Positrons"
53: Zeitschr. Phys. {\em 98\/}, 714 (1936). \\[2mm]
54: Translation:
55: W. Korolevski and H. Kleinert,
56: Institut f\"ur Theoretische Physik,
57: Freie Universit\"at Berlin, Arnimallee 14, D-14195 Berlin, Germany. \\
58: emails: {\tt walja.k@web.de} and {\tt kleinert.physik.fu-berlin.de}
59: \end{abstract}
60: The fact that electromagnetic radiation can be transformed into matter and vice versa leads to
61: fundamentally new features in quantum electrodynamics.
62: One of the most important consequences is that, even in the vacuum, the
63: Maxwell equation have to be exchanged by more complicated formulas. In general,
64: it will be not possible to separate processes in the vacuum from those involving matter
65: since electromagnetic fields can create matter if they are strong enough. Even if they
66: are not strong enough to create matter they will, due to the virtual possibility
67: of creating matter, polarize the vacuum and therefore change the Maxwell equations.
68: This polarization of the vacuum to be studied below
69: will give rise to a
70: distinction between the vectors $\mathfrak B$, $\mathfrak E$
71: on the one hand and $\mathfrak D$, $\mathfrak H$ on the other, where%
72: %\footnotetext[1]{}
73: 
74: 
75:  \includegraphics[scale=0.6]{f1.eps}
76: The polarizations $\mathfrak P$ and $\mathfrak M$ can be arbitrary functions of
77: the field strengths at the same place, their derivatives, and the field strengths in the
78: surroundings of the observed position. If the field strengths are small
79: (which means, as we shall see small compared to $e^2/\hbar c$-times of the field strength
80: at the boundary of the electron), then $\mathfrak P$ and $\mathfrak M$ can be approximately
81: considered as linear functions of $\mathfrak E$ and
82: $\mathfrak B$. In this approximation, Uehling\footnote{E. A. Uehling, Phys. Rev. {\bfseries 48}, 55, 1935. \label{Ueh}}
83: and Serber\footnote{R. Serber, Phys. Rev. {\bfseries 48 }, 49, 1935. \label{Serber}} have calculated
84: the modifications of Maxwell's theory. Another interesting case is obtained by not assuming
85: small field strengths but instead slowly varying fields (i.e., the fields are nearly constant
86: over the length $\hbar/mc$).
87: Then one obtains $\mathfrak P$ and $\mathfrak M$ as functions of $\mathfrak E$
88: and $\mathfrak B$ at the same position. The derivatives of $\mathfrak E$ and $\mathfrak B$ do not appear
89: in the approximation. The expansion of $\mathfrak P$ and $\mathfrak M$ in powers of $\mathfrak E$ and $\mathfrak B$
90: will contain only odd powers, as will be seen in the calculation. The
91: expansion terms of
92: third order are
93: phenomenologically related to light-light scattering
94: and are already known\footnote{H. Euler, B. Kockel, Naturwissensch. {\bfseries 23}, 246, 1935.\label{EulerKockel}}.
95: The goal of this work is to find the functions
96: $\mathfrak P(\mathfrak E,\mathfrak B)$ and $\mathfrak M(\mathfrak E,\mathfrak B)$
97: for slowly varying
98: field strength. It is sufficient to calculate to energy density $U(\mathfrak E,\mathfrak B)$.
99: From the energy density one can derive the
100: fields using the Hamiltonian method; one introduces the
101: Lagrangian $\mathfrak L(\mathfrak E,\mathfrak B)$ and obtains
102: \includegraphics[scale=0.6]{f2.eps}
103: The Lagrangian is determined by (3) and $\mathfrak D$ and $\mathfrak H$ are determined by (2).
104: Due to relativistic invariant the Lagrangian can only depend
105: on the two invariants ${\mathfrak E}^2-{\mathfrak B}^2$ and $({\mathfrak EB})^2$ (compare \ref{EulerKockel}).
106: The calculation of $U(\mathfrak E,\mathfrak B)$ can be reduced to the question of how much energy density is associated with the matter fields in a background of constant fields $\mathfrak E$ and
107: $\mathfrak B$. Before solving this problem, the mathematical scheme of the positron
108: theory\footnote{W. Heisenberg, ZS. f. Phys. {\bfseries 90}, 209, 1934\label{Heisenberg}} will be presented. This will also correct some errors in earlier formulas.
109: 
110: \section{The Mathematical Schema of the Theory of the Positron}
111: 
112: Starting point of the theory is the Dirac density matrix, which is given
113: in naive
114: wave theory by
115: \begin{equation}
116: (x' t' k' |R |x'' t'' k'') =   \underset{\scriptscriptstyle \text{ occupied states}}{\sum (n)}
117: \psi _n ^{*} (x''t''k'') \: \psi _n (x't'k')\tag{4}
118: \end{equation}
119: and in this quantum theory of wave fields by\footnote{In the earlier paper [\ref{Heisenberg}] the values on the right hand side are interchanged}
120: \begin{equation}
121: (x' t' k' |R |x'' t'' k'') =  \psi  ^{*} (x''t''k'') \: \psi (x't'k')\tag{5}
122: \end{equation}
123: Apart from this matrix an important role is played by the matrix $R_S$,
124: formally defined by
125: \begin{equation}
126: (x' t' k' |R_S |x'' t'' k'') = \frac{1}{2} \biggl (  \underset{\scriptscriptstyle \text{occupied states}}{\sum (n)} -
127: \underset{\scriptscriptstyle \text{unoccupied states}}{\sum (n)}\quad  \biggr )
128: \psi _n ^{*} (x''t''k'') \psi _n (x't'k')\tag{6}
129: \end{equation}
130: and in the quantum theory by
131: \begin{equation}
132: (x' t' k' |R_S |x'' t'' k'') = \frac{1}{2} \biggl[ \psi  ^{*} (x''t''k'') \: \psi (x't'k') -
133: \psi (x't'k') \: \psi ^* (x''t''k'') \biggr] \tag{7}.
134: \end{equation}
135: The matrix $R_S$ as a function of the differences $x_\lambda' - x_\lambda'' = x_\lambda$
136: and $t'-t''=t$ becomes singular on the light cone. If we set
137: 
138: \includegraphics[scale=0.6]{f8.eps}
139: furthermore the potentials $A_0=-A^0$; $A_i=A^i$, and  for the Dirac matrices
140: $\alpha^0=-\alpha_0=1$; $\alpha^i=\alpha_i$, we derive
141: 
142: \includegraphics[scale=0.6]{f9.eps}
143: where\footnote{In the earlier paper [\ref{Heisenberg}] the exponent has wrongly a negative sign.}
144: 
145: \includegraphics[scale=0.6]{f10.eps}
146: (The sum of repeated Latin indices runs from 1 to 3, and of Greek indices from 0 to 3).
147: The integral is to be calculated along the straight line from $P'$ to $P''$.
148: 
149: The density matrix $r$ responsible for the behavior of matter is obtained from $R_S$ by the equation
150: 
151: \includegraphics[scale=0.6]{f11.eps}
152: where $S$ is given by
153: 
154: \includegraphics[scale=0.6]{f12.eps}
155: Here $S_0$ is the matrix $R_S$ in field-free and matter-free space, where
156: $\bar{a}$, $\bar{b}$, and $C$ are defined\footnote{In the earlier paper [\ref{Heisenberg}] there is a mistake that leads to another value of C. There, the character $\gamma$ stands for the logarithm of the Euler constant, in contrast to the common usage.} by
157: 
158: \includegraphics[scale=0.6]{f13.eps}
159: where $\gamma$ is the Euler constant $\gamma=1.781...$
160: 
161: The four-vector for the current density and the energy momentum tensor are obtained from $r$ by
162: 
163: \includegraphics[scale=0.6]{f14.eps}
164: In the quantum theory of wave fields it is useful to expand the wave function into an
165: orthogonal system:
166: 
167: \includegraphics[scale=0.6]{f15.eps}
168: The operators $a_n$ can be represented in the form
169: 
170: \includegraphics[scale=0.6]{f16.eps}
171: where $\Delta_n$ changes the number $N_n$ to $1-N_n$ and
172: $V_n=\prod_{t\leq n}(1-2N_t)$.
173: 
174: We further set
175: 
176: \includegraphics[scale=0.6]{f16b.eps}
177: The Hamiltonian of the whole system reads in these variables:
178: 
179: \includegraphics[scale=0.6]{f17.eps}
180: According to the powers of the elementary change $e$, we distinguish the terms:
181: 
182: \includegraphics[scale=0.6]{f18.eps}
183: 
184: \section{Calculation of the Energy Density in Intuitive Wave Theory}
185: 
186: Since the Lagrangian associated with the corrected Maxwell equations
187: is a function of the two invariants ${\mathfrak E}^2-{\mathfrak B}^2$ and $({\mathfrak EB})^2$,
188: it is sufficient to calculate the energy density of the matter field as a
189: function of two independent field quantities. For example, it will be sufficient
190: to calculate the energy density in a constant electric and a
191: constant parallel magnetic field. In these constant fields we have to
192: analyze the state of the matter field in the absence of matter,
193: i.e., in the state of lowest energy. In an intuitive wave theory based on equations (4) and (6),
194: the state of lowest energy is given when all electron states of negative energy
195: are occupied and all states of positive energy are empty. In the presence of only
196: a magnetic field, the stationary states of an electron can be divided into
197: those of negative and positive energy. Hence the state of the lowest energy
198: of the matter field can be derived in the same way as for a field-free space.
199: 
200: The situation is different in an electric field. In this case the potential
201: energy grows linearly in one space direction, so all energy values from $-\infty$ to $+\infty$
202: are possible. The eigenfunctions associated with different eigenvalues are transformed
203: into each other by a spatial translation. Hence a classification of energy values into
204: positive and negative is not unique.
205: 
206: This difficulty is physcially related to the fact that in an electric field,
207: pairs of positrons and electrons are created. The exact analysis of this problem
208: was performed by Sauter\footnote{F. Sauter, ZS. f. Phys. {\bfseries 69}, 742, 1931 \label{Sauter}}.
209: 
210: \begin{figure}[h]
211:         \centering
212:         \includegraphics[scale=0.3]{fig1.eps}
213: \end{figure}
214: 
215: In Fig. 1, the potential energy $V(x)$ and the lines $V(x)+mc^2$
216: and $V(x)-mc^2$ are plotted against the
217: coordinate (the electric field is parallel $x$-axis).
218: The calculations of Sauter show that the eigenfunction associated to the eigenvalue
219: $E_0$, for example, is large only in the regions I and II. In the region II, they decrease exponentially.
220: Therefore, a wave function that begins being large in region I decreases slowly in region III where the
221: transmission coefficient through region II (which plays the role of a Gamow-wall)
222: calculated by Sauter has the order of magnitude $e^{-\frac{m^2c^3}{\hbar e |{\mathfrak E}|}\pi}$.
223: If we define $|{\mathfrak E}_k|=\frac{m^2c^3}{\hbar e}$ as the critical field strength,
224: we can also write $e^{-\frac{|{\mathfrak E}_k|}{|{\mathfrak E}|}\pi}$. As long as $|{\mathfrak E}|\ll|{\mathfrak E}_k|$,
225: pair creation is so rare that it can be practically ignored. Then it must be possible to find solutions
226: of the Dirac equation playing the role of eigenfunctions, which are large in region I but stay small
227: of the order of $e^{-\frac{|{\mathfrak E}_k|}{|{\mathfrak E}|}\pi}$ in region III.
228: Conversely, it must be possible to find solutions which are large in region III
229: and small in region I. After that we can characterize the state of the lowest energy by
230: all electron states being occupied whose eigenfunctions are large only in region III,
231: while the others are unoccupied. The energy
232: density at $x_0$ is calculated from
233: the differences of the electron energies
234: with $E_0$ [compare Eq. (31)].
235: By switching off the electric field adiabatically, the
236: so characterized state of the system goes over into the state of the field-free space, in which only the
237: negativ-energy electron states are occupied.
238: 
239: 
240: 
241: Our calculations follow those of Sauter. If an external magnetic field $\mathfrak B$
242: and an electric field $\mathfrak E$ are present, both pointing in $x$-direction, the Dirac equation reads:
243: 
244:  \includegraphics[scale=0.6]{f19.eps}
245: The movement in $y$- and $z$-directions can be separated from that in $x$-direction
246: by the ansatz:
247: 
248: \includegraphics[scale=0.6]{f20.eps}
249: We introduce a new operator $K$ by
250: 
251:  \includegraphics[scale=0.6]{f21.eps}
252: It satisfies:
253: 
254:  \includegraphics[scale=0.6]{f22.eps}
255: This equation can be understood as a wave equation determing eigenfunctions $u_n(y)$.
256: If we set
257: 
258:  \includegraphics[scale=0.6]{f23a.eps}
259: we find (22) which becomes essentially the Schroedinger equation of the oszillator,
260: 
261:  \includegraphics[scale=0.6]{f23.eps}
262: ($H_n(\eta)$ is the $n$-th Hermite polynom). The eigenvalues are:
263: 
264: \includegraphics[scale=0.6]{f24.eps}
265: The operator $K$ anticommutates with $\alpha_1$ in the wave function (19), which can also
266: be written as
267: 
268: \includegraphics[scale=0.6]{f25.eps}
269: We can perform a canonical transformation in $\chi$, so that $\sigma_x$ becomes diagonal,
270: and $K$ and $\alpha_1$ read:
271: 
272: \includegraphics[scale=0.6]{f26.eps}
273: Here the two matrices refer to another index independent of the spin orientation
274: (they refer to the ``$\varrho$''-coordinate). One can treat $\sigma_x$ as if it was
275: an ordinary number ($\sigma_x=\pm 1$) and obtains with the convenient shorthand notation
276: 
277: \includegraphics[scale=0.6]{f27.eps}
278: the equations
279: 
280: \includegraphics[scale=0.6]{f28.eps}
281: where $f$ and $g$ are the two components of the function $\chi$ (associated with the $''\varrho''$-index).
282: The equations (28) are formally identical with the Equations (12) from Sauter''s paper (op. cit. \ref{Sauter}).
283: They differ from those, however, in the meaning of $k$ and that the system (28) actually has to
284: be written down twice; once for $\sigma_x=+1$
285: and once for $\sigma_x=-1$. Sauter obtains two sets of solutions:
286: 
287: \includegraphics[scale=0.6]{f29.eps}
288: The integrals are to be taken along a path coming from $+\infty$ circulating around both singular points
289: $+i/2$ and $-i/2$ in a positive manner, and returning to $+\infty$.
290: 
291: Since we ignore pair creation in the calculation we consider as eigenfunctions only those parts of the functions
292: $f$ and $g$ which vanish in one half of the space. Thus we set
293: 
294:         \includegraphics[scale=0.6]{f30.eps}
295: The new functions $f^1_1$ etc. do not correspond exactly to the stationary states but
296: represent wave packets for which it is unlikly to diffuse into the initially empty
297: area. For the construction of the density matrix we consider the states $f^1_1$, $g^1_1$, ...
298: as being occupied, the states $f^2_1$, $g^2_1$, ... as unoccupied.
299: We have doubled the number of states by the process (30). Thus, if we take all $f^1_1$, $g^1_1$, $f^1_2$, $g^1_2$
300: as occupied and $f^2_1$, $g^2_1$, $f^2_2$, $g^2_2$ as unoccupied, we shall obtain the density
301: matrix twice.
302: 
303: To find the energy density of the vacuum according to the method described in the first section we
304: would have to calculate the density matrix associated with a
305: finite distance of the two positions $\mathfrak{r'}$
306: and $\mathfrak{r''}$. From this we have to subtract the singular matrix $S$.
307: Then obtain
308: the energy density by taking
309: the limit $\mathfrak{r'}=\mathfrak{r''}$. For the following calculation
310: it is more convenient to set
311:  from the beginning $\mathfrak{r'}=\mathfrak{r''}$, but to restrict the summation over
312: the stationary states to those of finite energy only. Aquivalently, we could make the summation convergent
313: by an auxiliary factor $e^{-\mathrm{const}[E^2-(mc^2)^2]}$.
314: If the constant number in the exponent goes to zero, some of the terms in the energy density
315: become singular, and they will be compensated by corresponding elements in the matrix $S$.
316: The remaining regular elements yield the desired result.
317: 
318: Bevor writing down the density matrix, we have to normalize the eigenfunctions.
319: We can imagine the space of the eigenfunctions in $x$- and $z$-direction being
320: confined to a large length $L$ (the above eigenfunctions $u_n(y)$ are already normalized).
321: Then we have from the $z$-direction a normalization factor $1/\sqrt{L}$; from the
322: $x$-direction, the asymptotic behavior of the Sauter eigenfunctions [compare Eq. (22)]
323: yields the factor $2\frac{1}{\sqrt{L}}e^{-\frac{k^2\pi}{4}}$. The summation over all
324: states is to be done over all momenta which have the form
325: 
326: $$p_z=\frac{h}{L}m+\mathrm{const}$$
327: and over all energies of the form $E=\frac{hc}{L/2}m+\mathrm{const}$.
328: These two sums can be converted into integrals. When neglecting the factor $\frac{1}{\sqrt{L}}$
329: the differential is $\frac{dp_x}{h}\frac{dE}{2hc}$. If we calculate the density at position
330: $x_0$, the energy of a state is the difference $E-e|{\mathfrak E}|x_0$. We therefore obtain for
331: the energy density corresponding to the matrix $R_S$ (compare Section 1) [for the meaning of $a$ compare (33)]
332: 
333:  \includegraphics[scale=0.6]{f31.eps}
334: which due to (23a) goes over into
335: 
336:         \includegraphics[scale=0.6]{f32.eps}
337: From this expression we have to derive the limit $\alpha\overrightarrow{}0$.
338: We introduce the following convenient abbreviations:
339: 
340:         \includegraphics[scale=0.6]{f33.eps}
341: These are dimensionsless and represent the ratio of the field strengths with the critical field strength
342: $ |\mathfrak {E} _k | $, i.e., with the ``$1/137 $th fraction of the field strength at the radius of the electron''.
343: With the equation (29) one finally finds
344: 
345: \includegraphics[scale=0.6]{f34.eps}
346: where $h(s_1 s_2)$ is
347: 
348: \includegraphics[scale=0.6]{f34b.eps}
349: The integration over $\xi $ yields:
350: 
351: \includegraphics[scale=0.6]{f35.eps}
352: The first of the two $s$-integrals, i.e., the one over $s_1$, can be performed by deforming the integration
353: path in such a manner that only a loop around the pole $s_1 =s_2 -\alpha $ is left. Replacing $s_2 $ by $ s=s_2 - \alpha $
354: in the result, one obtains:\\
355: 
356: \includegraphics[scale=0.6]{f36.eps}
357: where
358: 
359: \includegraphics[scale=0.6]{f37.eps}
360: The integration is to be carried out along a path shown in Fig. 2. The contour comes from $+\infty $ passes
361: between the four poles, and returns back to $+\infty $. Alternatively, one may integrate along the imaginary axis
362: from $+\infty $ to $-\infty $.
363: 
364: \begin{figure}[h]
365: \centering
366: \includegraphics[scale=0.3]{fig2.eps}
367: \end{figure}
368: \pagebreak
369: 
370: The main contribution to the integral comes from the regions between the poles. One may expand the logarithm
371: in the exponent in power series of $ \alpha $, and obtains~\\
372: 
373: \includegraphics[scale=0.6]{f38.eps}                  ~\\
374: For the following calculations, it is assumed that the electric field is small against the critical field
375: $|\mathfrak {E} _k | $, i.e. $a\ll 1 $ and thus \\
376: 
377: \includegraphics[scale=0.6]{f38b.eps}      ~\\
378: Then the expression $-\frac{k^2 \alpha}{4s^2 +1}$ in the exponent must be retained; the higher terms in the exponent
379:  may be considered as small, while we shall eventually take $\alpha\overrightarrow{}0$. They can therefore be included
380: by an expansion. Thus, one obtains an expression for $f(k)$ of the form \\[5mm]
381: 
382: \includegraphics[scale=0.6]{f39.eps} \\[5mm]
383: 
384: Before performing the integral, it is convenient to calculate the summation over $n $ and $\sigma $; it is carried
385: out according to the scheme                   \\
386: 
387: \includegraphics[scale=0.6]{f40.eps}     \\~\\
388: It turns out that the higher terms of Euler summation formula contribute nothing to the end result. Finally,
389: setting $\frac{\alpha}{a}=\varepsilon $, one gets for $\varepsilon \to 0$ (where $\gamma$ denotes the Euler constant $ \gamma = 1.781$).
390: 
391: \pagebreak
392: \includegraphics[scale=0.6]{f41.eps}
393: (The coefficients $c_m $ are to be determined further below).
394: 
395: The parts corresponding to the singular matrix $ S $ still have to be subtracted from the result. One obtains
396: easily the field-independent part of the singular energy density, which is to subtract by repeating the above
397: calculation with plane waves:
398: 
399: \includegraphics[scale=0.6]{f41b.eps}
400: It is more difficult to calculate the field-dependent part of $S$. Since equation (13) contains only squares of
401: field strengths in $\bar{a}$ and $\bar{b}$, this would also for $U_S $. Further, the constant C in equation
402: (13) is chosen so that for constant fields, polarisation proportional to the field does not
403: occur. From this it follows that the terms which are quadratic in the field are to be subtracted, and only the
404: higher terms remain. This implies, anticipating an expansion in powers of $b$ for $b\ll 1 $, that the first line of
405: the right-hand equation (41) is to subtract on the whole. We proved this result by assuming that the points
406: $ \mathfrak {r} '$ and $\mathfrak {r} ''$ of the density matrix are different in the $x$ direction
407: $ [\mathfrak {r} ' -\mathfrak{r} ''=(x,0,0)]$, and by setting at the end $\alpha =\epsilon =0 $. After subtracting the terms
408: which come from the matrix $ S $, only the above discussed part of $U $ remains. For the electric
409: field, this calculation turned out to be very difficult. Starting with the total energy density, which is
410: composed of the usual Maxwell energy density $ \frac{1}{8\pi}(\mathfrak{E} ^2 +\mathfrak{B} ^2) $ and
411: Dirac energy density $ U-U_S $,
412: we go on with the Lagrangian using the relation (3)
413: 
414: \includegraphics[scale=0.6]{f41c.eps}
415: One obtains
416: 
417: \includegraphics[scale=0.6]{f42.eps}
418: For small magnetic fields, the power series expansion of $ b $ may be found applying to equation (42) once more
419: the Euler summation formula. Furthermore, since $ \mathfrak{L}$ depends only on both invariants $ a^2 -b^2 $ and $ a^2 b^2 $, from which it follows that $ \mathfrak{L} (a, b)$: $ \mathfrak{L} (a, 0)= \mathfrak{L} (0, ia) $, the missing
420: coefficients $ c_m $, whose direct calculation would be very involved, may be determined indirectly from this relation. By calculating $ c_2 $ and $c_3 $ we have checked that the direct calculation of $ c_m $ yields the same result; though we didn't find the general proof for it. In this manner, one obtains initially for small fields ($ a \ll 1$, $b \ll 1$):
421: 
422: \includegraphics[scale=0.6]{f43.eps}
423: The opposite limit ($ a \ll 1$, $b \gg 1 $) yields
424: 
425:  \includegraphics[scale=0.6]{f44.eps}
426: We tried to derive an integral representation of $ \mathfrak{L}$ in order to better see the general behavior of
427: $ \mathfrak{L}$ for an arbitrary field. This is possible when using the usual integral representation of the
428: zeta function. One obtains
429: 
430:   \includegraphics[scale=0.6]{f45.eps}
431: From the last form it is easy to see that $ \mathfrak {L} $ only depends on both invariants $ \mathfrak {E} ^2
432: - \mathfrak {B} ^2 $ and $ (\mathfrak {EB})^2 $. The cos terms allow an expansion in squares of the argument
433: $ (b +ia)^2 = b^2 - a^2 + 2i(ab) $ and $ (b +ia)^2 = b^2 - a^2 - 2i(ab) $. Since the overall result is real,
434: it may be represented as power series in $ b^2 - a^2 $ and $ (ab)^2 $, which may be replaced by
435: $\frac{\mathfrak{B} ^2 -\mathfrak {E} ^2}{|\mathfrak {E} _k | ^2} $ and
436: $\frac{\mathfrak{EB} ^2 }{|\mathfrak {E} _k | ^4}  $ respectively.
437: $\biggl( |\mathfrak {E} _k | ^2 = \frac {m^2 c^3}{e \hbar} \biggr) $.
438: Thus the Lagrangian for arbitrarily oriented fields reads:
439: 
440: \includegraphics[scale=0.6]{f45a.eps}
441: The first expansion term of equation (43) agrees with the results of Euler and Kockel (op. cit. \ref{EulerKockel}).
442: 
443: The question of convergence of this power series expansion deserves closer investigation. The integral (45)
444: converges for every value of $ b$ when $ a=0 $. But when $ a \neq 0 $, the integral looses its meaning for
445: $ \eta = \frac{\pi }{a},\,  \frac{2 \pi }{a}, \cdots $, where $\cot a\eta $ becomes infinite. According to this, the power
446: series expansion in $ a $, from which we started, can only be semiconvergent. We can give a definite meaning
447: to the integral (45) by selecting an integration path that avoids the singular values
448: $ \frac{\pi }{a},\,  \frac{2 \pi }{a} $. But then the integral (45) acquires additional imaginary terms
449: that can't be interpreted physically. Their meaning becomes clear if we estimate their size.
450: The integral (45) around the pole $ \eta = \frac {\pi }{a} $ has the value
451: $ -\frac {2i}{\pi }\cdot 4 a^2 mc^2 (\frac {mc}{h})^3 \cdot e^{-\frac{\pi }{a}}$ (for $ b=0 $). This is the order of the terms
452: which are associated with the pair creation in the electric field. Thus, the integral is to be interpreted
453: similar to the integration of the resonance denominator in perturbation theory. One may assume that a damping
454: term corresponding to the frequency of the resonance process is responsible for the convergence. Further, one
455: may assume that the result, which is obtained by avoiding the singular values, is correct except for the terms
456: whose order corresponds to the frequency of the resonance process.
457: 
458: The deviations from the Maxwell theory remain very small according to (43) and (44) as long as $ \mathfrak{E} $
459: and $ \mathfrak{B} $ are small against the electric field present at a distance of
460: $ \sqrt{137} \cdot \frac{e^2}{mc^2} $ from the center of the electron. But also in the case that the magnetic field
461: exceeds this value, the additional terms to the Maxwell equations remain small (proportional to the original terms).
462: They have the relative order $ \frac {1}{3\pi} \frac{e^2}{\hbar c}$ as long as $ \log {b} $ is near unity. Thus,
463: the deviations from the usual Coulomb force between the protons (for example) come from the terms (43) and (44),
464: and remain always very small. However, in this estimate, one must take into account that for a Coulomb field, the
465: neglected terms containing the derivatives of the field strengths are more important than those contained
466: in equations (43) and (44).
467: 
468: \section{Importance of the Result in the Quantum Theory of Wave Fields}
469: The results derived in the last section may not be transfered immediately to the quantum theory of wave fields.
470: In fact, one may show that the state of matter in a homogeneous field in the quantum theory of wave fields is not
471: described by the above equations.
472: Even if we start with the state of matter discussed in the last section as an ``undisturbed state'', there are
473: matrix elements of the perturbation theory which belong to the simultaneous formation of a light quantum and a pair.
474: Even, if the energy does not suffice for a real formation of these particles, such matrix elements give rise to a
475: perturbative energy of second order. This energy comes about through the virtual
476: possibility of formation and disappearance of the light quantum and a pair, and the calculation provides a
477: divergent result for them. The appearance of these perturbation terms may be made plausible by observing that
478: the circular trajectories in a magnetic field are not really stationary states, but that the electrons in these
479: states can radiate. In the classical theory of wave fields, this radiation can be ignored - this being crucial
480: for the physical content of the calculation of the last section. Indeed, in the final solution the charge density
481: and current density of matter vanish, so that radiation does not occur. In the quantum theory of wave fields
482: on the other hand, this radiation does appear in form of the divergent perturbative energy of second order.
483: 
484: One may also understand that a perturbative energy of exactly the same type occurs also in the field free vacuum
485: (``self-energy of the vacuum''). Such self-energies are found always when calculating the contribution
486: of second order or higher to the energy, which arises from virtual transition to another state and return to
487: the initial state. These self-energies have so far always been ignored. For instance, the cross
488: section of the
489: Compton scattering may be calculated by performing a perturbation calculation to second order. Considering
490: the terms of the forth order, one would get a contribution of the described type, which does not yield a
491: convergent result. The calculation of the light-light scattering is carried out to the forth order
492: (the lowest order that yields a contribution to the relevant process). The contributions of the sixth order
493: would already diverge. The previous success of these calculations, for instance the Klein-Nishina formula, seems
494: to show that omitting the divergent contributions of the higher order leads to the correct results. If this is
495: the case, then it follows from the above discussion that the results of Section 2 can be transferred to the
496: quantum theory of wave fields. This is also physically plausible, since the occurrence of the above radiation
497: terms would remain incomprehensible according to the correspondence principle. Each single term in the expansion
498: of the energy density in powers of $ \mathfrak{E}$ and $ \mathfrak{B}$ can concretely be associated with a
499: scattering process, whose cross section is determined by it. The terms of forth order, for instance, stand for
500: the usual light-light scattering, the terms of the sixth order determine the cross
501: section of the process
502: whereby three light quanta are scattered on each other etc. Irrespective of the question whether it is physically
503: acceptable to neglect higher order terms, each expansion term in the result of the last
504: section agree with a direct
505: calculation of the corresponding scattering process in the quantum theory of wave fields if the perturbation
506: calculation is only performed to the lowest order that yields a contribution to the corresponding process.
507: In both calculations, the contributions of the terms are neglected which correspond to the formation and
508: disappearance of the light quantum and a pair. [The agreement of the terms of forth order with the terms obtained
509: by the direct calculation of light-light scattering (compare \ref{EulerKockel}) is therefore a test for the
510: correctness of the calculation.]
511: For this reason, it is not ruled out that even the results for $|\mathfrak{B}| \geqq  |\mathfrak{E} _k| $
512: can be applied to experience. This is, however, certainly not true for $|\mathfrak{E}| \gtrsim  |\mathfrak{E} _k| $;
513: The pair creation destroys the basis for the above calculations when the electric fields are large.
514: 
515: \section{The Physical Consequences of the Result}
516: The results derived in the second section are very similar to Born's%
517: \footnote{M. Born, Proc. Roy. Soc. London (A) {\bfseries 143 }, 410, 1933;
518: M. Born and L. Infeld, Proc. Roy. Soc. London (A) {\bfseries 144 }, 425, 1934; {\bfseries 147}, 522, 1934;
519: {\bfseries 150}, 141, 1935.}
520: derivation of Maxwell equations.
521: Also Born obtains a more complicated function of the two invariants, $\mathfrak{E} ^2 -\mathfrak{B}^2 $
522: and $(\mathfrak{EB}) ^2$ in addition to the classical Lagrangian $\mathfrak{E} ^2 -\mathfrak{B}^2 $.
523: Incidentally, because of the actual value of $\frac{e ^2}{\hbar c} $, this function agrees with (43) in the order of
524: magnitude of the lowest expansion terms (compare \ref{EulerKockel}). Though it is also important to emphasize
525: the differences of both results. Born has chosen the modified Maxwell equation as the starting point of the theory,
526: whereas this change in the Dirac theory is a really indirect consequence of the virtual possibility of the pair
527: creation. In addition, Dirac's theory predicts also terms, which contain higher powers of the field strengths (compare
528: \ref{Ueh}, \ref{Serber} ).
529: Thus, especially the question of the self-energy of the electrons cannot be decided only with the help
530: of these changes. The result of Born's theory imply that the changes of the Maxwell equations calculated here
531: suffice to remove difficulties with the infinite self energy is an important indication of the further
532: development of the theory.
533: 
534: In this context, we must also ask the question whether the results about light-light scattering etc. derived
535: from the Dirac theory may be considered to be final or whether it is expected that a future theory will lead to
536: another results. Without doubt, the theory of the positron and the present quantum electrodynamics must be
537: considered as temporary. Especially, the rules of finding the $S$ matrix seem to be arbitrary in this theory
538: (the inhomogeneity of Dirac's equation). The only reason for the matrix $S$ introduced in [\ref{Heisenberg}]
539: is the relative mathematical simplicity (together with some postulates on the formulation of the conservation laws).
540: Thus, a deviation of the future theory from the present theory may well be possible. Since such variations may have a
541: big influence on the Maxwell equations, the present theory is only reliable in its order of magnitude
542: and the qualitative form of these changes. Up to now, it is hardly possible to make definite statements as to
543:  the final form of the Maxwell equations in future theory, since it is essential to consider all the processes of
544: particles with very high energy (for instance, the occurrence of ``showers'').
545: 
546: 
547: 
548: %\begin{thebibliography}{99}
549: %
550: %\bibitem{UH}
551: %E.A. Uehling etc
552: %\end{thebibliography}
553: 
554: \end{document}
555: