1: \documentclass[11pt]{article}
2:
3: %\pagestyle{empty}
4: \oddsidemargin -5mm
5: \textwidth 170mm
6: \textheight 222mm
7: \columnwidth\textwidth
8: \usepackage{graphicx}
9: \usepackage{epsfig}
10:
11: \begin{document}
12: \title{Coulomb Corrections for Coherent Electroproduction of
13: Vector Mesons: Eikonal Approximation}
14: \author{Andreas Aste, Kai Hencken, Dirk Trautmann\\
15: Institut f\"ur Physik der Universit\"at Basel, 4056 Basel,
16: Switzerland}
17: \date{July 8, 2004}
18:
19: \maketitle
20:
21: \begin{center}
22: \begin{abstract}
23: Virtual radiative corrections due to
24: the long range Coulomb forces of heavy nuclei with
25: charge $Z$ may lead to sizeable corrections to the Born cross
26: section usually used for lepton-nucleus scattering processes.
27: An introduction and presentation of the most important
28: issues of the eikonal approximation is given.
29: We present calculations for forward electroproduction
30: of rho mesons in a framework suggested by the VDM
31: (vector dominance model), using the eikonal approximation.
32: It turns out that Coulomb corrections may become relatively large.
33: Some minor errors in the literature are corrected.
34: \vskip 0.2 cm
35: \noindent {\bf{PACS.}} 25.30.-c Lepton induced reactions - 25.30.Rw
36: Electroproduction
37: reactions - 13.40.-f Electromagnetic processes and properties - 25.30.Bf
38: Elastic electron scattering
39: \vskip 0.2 cm
40: \noindent Eur. Phys. J. A (2004)\\
41: \noindent DOI 10.1140/epja/i2003-10182-3
42: \end{abstract}
43: \end{center}
44:
45: \newpage
46:
47: \section{Introduction}
48: Due to the small size of the elemental electric charge
49: $e=\sqrt{4 \pi \alpha}$,
50: it is for most electromagnetic elementary particle reactions
51: fully sufficient to calculate only in Born approximation.
52: But this is no longer true
53: when heavy nuclei are involved, like e.g. lead, where the relevant
54: perturbation expansion parameter $\alpha Z \sim 0.6$ is not small.
55: It is then possible that even at high energies, the ratio of the exact and
56: the Born cross section does not approach unity.
57: One prominent example for this observation
58: is the photoelectric effect, where a single photon
59: knocks out an electron from the $K$ shell, or electron positron pair
60: production by a single photon incident on a nucleus, where the
61: electron is captured into the $K$ shell \cite{Aste,Agger}.
62: For small $\alpha Z$
63: and high photon energy $\epsilon_\gamma$, the cross section
64: for these processes is given by the Sauter formula \cite{Sauter}
65: \begin{equation}
66: \sigma_0 \simeq 4 \pi \alpha^6 Z^5 \lambda_C^2 \frac{1}{\epsilon_\gamma} ,
67: \end{equation}
68: where $\lambda_C$ is the electron Compton wavelength.
69: Applying the Born approximation in the usual sense, i.e. by the
70: use of plane waves for the positron, was attempted by Hall and
71: Oppenheimer \cite{Oppenheimer} already in 1931 in the case of the
72: photoelectric effect.
73: But a central difficulty in the treatment of the photoeffect
74: arose from the distorted wave function of the
75: ejected electron. The bound state wave function depends on
76: $\alpha Z$ and only terms of relative order $\alpha Z$ survive in
77: the Born matrix element. Terms of this order also
78: come from the continuum wave function.
79: Using exact wave functions for the continuum electron,
80: it turns out that the cross section calculated from plane
81: waves for the ejected electron is already wrong for small
82: $\alpha Z$.
83: It is therefore not
84: astonishing that Coulomb corrections can become very large even
85: at high energies.
86: For the derivation of the Sauter formula, nonrelativistic
87: bound state wave functions and approximate Sommerfeld-Maue wave
88: functions for the ejected electron were used.
89: Boyer \cite{Boyer} improved the calculation
90: by using exact bound state wave functions and Sommerfeld-Maue
91: wave functions for the continuum wave function.
92: The calculational advantage of using Sommerfeld-Maue wave functions
93: relies in the fact that the exact continuum wave functions
94: are only available as an expansion in partial waves, and
95: for higher energies summation over a large number of terms is
96: necessary in order to compute exact matrix elements.
97: But Sommerfeld-Maue wave functions are valid for the Coulomb
98: potential of point-like charges, and hence do not take into account
99: the finite size of the nucleus. We therefore pursue a similar
100: strategy, namely the eikonal distorted wave approximation,
101: where the advantage is that also finite size effects can be
102: taken into account. This is clearly necessary for nuclear
103: processes, whereas the relevant
104: physical length scale of the photoelectric
105: effect is given by $\sim (\alpha Z)^{-1} \lambda_C$ and much larger than
106: a typical nuclear radius.
107:
108: In the following part of this paper, we revisit the eikonal
109: approximation by giving a condensed introduction to the
110: subject, where we also point out the most important facts
111: related to possible improvements concerning
112: the range of validity of the method in its basic form.
113: In sect. 3, we apply our calculation to coherent
114: vector meson electroproduction.
115: It turns out that the Coulomb corrections are indeed
116: large for heavy nuclei, and that it is necessary to use
117: a rather accurate description of the electrostatic
118: nuclear potential. We correct some errors in the literature
119: and rederive the eikonal result given in \cite{Kopeliovich}.
120:
121: \section{Eikonal Approximation}
122: For highly relativistic particles in a potential $V$ with
123: asymptotic momentum $\vec{p}$,
124: one may neglect the mass of the
125: particle ($|E-V|, |\vec{p} \, | \gg m$),
126: such that the energy-momentum relation reduces to
127: ($\hbar=c=1$)
128: \begin{equation}
129: (E-V)^2=\vec{p}^{\, 2} + m^2 \, \rightarrow \, E-V=p, \, p=|\vec{p} \,|.
130: \label{classical}
131: \end{equation}
132: The classical relation (\ref{classical}) for the (initial) momentum of the
133: particle $\vec{p}_i=p_i \hat{p}_i$ can be taken into account
134: approximately in quantum theory by modifying the plane wave
135: describing the initial state of the particle by the eikonal phase $\chi_1(\vec{r})$
136: \begin{equation}
137: e^{i\vec{p}_i\vec{r}} \, \rightarrow \, e^{i \vec{p}_i\vec{r}+i\chi_1(\vec{r})} \, ,
138: \end{equation}
139: \begin{equation}
140: \chi_1(\vec{r})=-\int \limits_{-\infty}^{0} V(\vec{r}+
141: \hat{p}_i s) ds \, .
142: \end{equation}
143: If we choose the initial momentum parallel to the $z$-axis
144: $\vec{p}_i=p_z^i \hat{{\bf{e}}}_z$,
145: the phase is
146: \begin{equation}
147: \chi_1(\vec{r})=-\int \limits_{-\infty}^{z} V(x,y,z') dz' \, .
148: \end{equation}
149: The $z$-component of the momentum then becomes
150: \begin{equation}
151: -i \partial_z e^{i p_z^i z+i\chi_1}=
152: (p_z^i-V)e^{i p_z^i z+i\chi_1},
153: \end{equation}
154: and further analysis shows that also the transverse momentum
155: modification is well approximated.
156: Analogously, for the final state wave function follows
157: \begin{equation}
158: e^{i \vec{p}_f\vec{r}-i\chi_2(\vec{r})},
159: \end{equation}
160: where
161: \begin{equation}
162: \chi_2(\vec{r})=-\int \limits_{0}^{\infty} V(\vec{r}+
163: \hat{p}_f s') ds' \, .
164: \end{equation}
165: For the sake of simplicity, we consider spinless particles first.
166: The spatial part of the free charged particle current
167: \begin{equation}
168: -ie[e^{-i \vec{p}_f\vec{r}} \vec{\nabla}
169: e^{i \vec{p}_i\vec{r}}-(\vec{\nabla}e^{-i \vec{p}_f\vec{r}})
170: e^{i \vec{p}_i\vec{r}}]= e(\vec{p}_i+\vec{p}_f)e^{i(\vec{p}_i-
171: \vec{p}_f)\vec{r}}
172: \end{equation}
173: becomes
174: \begin{displaymath}
175: -ie[e^{-i \vec{p}_f\vec{r}+i\chi_2(\vec{r})} \vec{\nabla}
176: e^{i \vec{p}_i\vec{r}+i\chi_1(\vec{r})}-
177: (\vec{\nabla}e^{-i \vec{p}_f\vec{r}
178: +i\chi_2(\vec{r})}) e^{i \vec{p}_i\vec{r}+i\chi_1(\vec{r})}]=
179: \end{displaymath}
180: \begin{equation}
181: e(\vec{p}_i+\vec{p}_f+\vec{\nabla} \chi_1-\vec{\nabla} \chi_2)
182: e^{i(\vec{p}_i-\vec{p}_f)\vec{r}+i\chi (\vec{r})
183: } \, , \label{scalarcurr}
184: \end{equation}
185: where $e$ is the charge of the particle and $\chi(\vec{r})=
186: \chi_1(\vec{r})+\chi_2(\vec{r})$.
187: The spatial part of the current now contains the
188: additional eikonal phase, and the prefactor
189: contains gradient
190: terms of the eikonal phase which represent essentially
191: the change of the electron momentum due to the attraction of
192: the electron by the nucleus.
193:
194: So far we have only considered the modification of the
195: phase of the wave function, and for many problems,
196: like the one treated in this paper, this is a sufficient
197: approximation. It has been applied to elastic high energy
198: scattering of Dirac particles in an early paper by Baker \cite{Baker}.
199: However, the method leads, e.g. for quasielastic nucleon knockout
200: scattering of electrons on lead with
201: initial energy $\epsilon_1=300$ MeV and energy transfer $\omega=100$ MeV,
202: to errors up to 50\% in the calculated cross sections.
203: The reason is that also the amplitude of
204: the incoming and outgoing particle
205: wave functions is changed due to the Coulomb attraction.
206: This fact is related to the classical observation
207: that an ensemble of negatively charged test particles approaching
208: a nucleus is focused due to its attractive potential.
209: For the sake of completeness, we give here a short discussion of this
210: effect.
211:
212: Knoll \cite{Knoll} derived the focusing from a high energy
213: partial wave expansion, following previous results given
214: by Lenz and Rosenfelder \cite{Lenz,Rosenfelder}.
215: E.g., for the incoming particle wave in the vicinity of
216: the nucleus he obtained an approximation for the electron wave
217: function ($u_p$ is the constant free electron spinor)
218: \begin{equation}
219: \psi_i=e^{i \delta_+} (p'/p) e^{i \vec{p'}\vec{r}}
220: \{1+a_1r^2-2a_2\vec{p'}\vec{r}+ia_1r^2 \vec{p'}\vec{r}+ia_2[
221: (\vec{p'} \times \vec{r})^2+
222: \vec{\sigma} (\vec{p'} \times \vec{r})] \} u_p, \label{expknoll}
223: \end{equation}
224: where $\sigma$ describes spin changing effects.
225: A similar formula holds for the final state wave function.
226: The so-called effective momentum
227: $\vec{p'}$ is parallel to $\vec{p}$ and is given by the classical
228: momentum of the electron in the center of the nucleus, i.e.
229: $p'=|E-V(0)|$.
230: The parameters $a_{1,2}$ depend on the shape of the nuclear electrostatic
231: potential. For a homogeneously charged sphere with radius $R_s$
232: they are given by $a_1=-\alpha Z/6 p' R_s^3, \,
233: a_2=-3 \alpha Z/4 p'^2 R_s^2$.
234: The increase of the amplitude of the wave while passing
235: through the nucleus is given by the $-2a_2\vec{p'}\vec{r}$-term.
236: The $a_1 r^2$-Term accounts for the decrease of the focusing
237: also in transverse direction.
238: Imaginary terms like $ia_2
239: (\vec{p'} \times \vec{r})^2$ describe the deformation of the
240: wave front near the center of the nucleus. They could be obtained
241: correspondingly by an expansion of the eikonal phase in that region,
242: and $\delta_+$ is the eikonal phase in the center of the nucleus
243: $\chi(0)$.
244:
245: Apart from the spin structure which is absent for scalar particles,
246: the eikonal approximation relies basically on the same strategy
247: for particles with or without spin, namely on the modification
248: of the wave function by the eikonal phase. But there are also
249: gradient terms present in the expression for the spinless current
250: (\ref{scalarcurr}), which lead to corrections to the current which
251: are of comparable magnitude as those which result from the wave
252: focusing. Such terms are not explicitly present in the
253: Dirac expression for the current (\ref{current1}) below.
254: It is instructive to have a closer look at the
255: structure of the electron current in a potential.
256: A Gordon decomposition of the electron current
257: \begin{equation}
258: j^\mu=e \bar{\Psi} \gamma^\mu \Psi \label{current1}
259: \end{equation}
260: can be performed by using the Dirac equation
261: $[i \gamma^\mu (\partial_\mu+ieA_\mu)-m] \Psi=0$.
262: The current can be split into a
263: convective current and a spin current
264: \begin{displaymath}
265: j_G^\mu=\Biggl\{
266: \frac{ie}{2m} \bigl[ \bar{\Psi} \partial^\mu \Psi -(\partial^\mu
267: \bar{\Psi}) \Psi \bigl] -\frac{e^2}{m} \bar{\Psi}\Psi A^\mu \Biggr\}
268: \end{displaymath}
269: \begin{equation}
270: +\frac{e}{2m} \partial_\nu [\bar{\Psi} \sigma^{\mu \nu} \Psi ] \, ,
271: \label{current2}
272: \end{equation}
273: which are separately conserved and gauge invariant.
274: For exact solutions of the Dirac equation, the two forms
275: of the current are of course equivalent.
276: Here, the convective current has exactly the same structure as the
277: current of a scalar particle in a potential.
278: Kopeliovich {\em{et al.}} \cite{Kopeliovich}
279: calculate Coulomb corrections by describing the leptons as
280: spinless particles. The discussion above clarifies when
281: such an approximation is justified (see also sect. 3).
282:
283: In the following, we will consider electrons with energies
284: of the order of several GeV. Then the impact of the focusing effect
285: on the matrix element will not be larger than about one percent even
286: for heavy nuclei like lead, where the electrostatic potential in the
287: center of the nucleus is ($Z=82$, $A=208$)
288: \begin{equation}
289: \sim \frac{3}{2} \frac{\alpha Z}{R_A} \sim 25 \, \mbox{MeV},
290: \end{equation}
291: and $R_A \sim 7.1$ fm is the equivalent radius of a
292: homogeneously charged sphere, which we will call
293: "nuclear radius" for short in the following.
294: E.g., for a scattering process where the initial and final
295: momentum of the electron is of the order of $10$ GeV/c,
296: the focusing enters the cross section by a factor of the order
297: $(p'/p)^4=(10.025/10)^4 \sim 1.01$, according to (\ref{expknoll}).
298:
299: \section{Coherent Electroproduction of Vector Mesons}
300: \subsection{Matrix Element}
301: Coherent electroproduction of vector mesons from virtual
302: photons plays an important role in the understanding of the
303: transition of soft diffractive models to quantum chromodynamics
304: \cite{Donnachie,Weise}. Models based on the
305: assumption of vector dominance have been applied successfully to
306: describe the data \cite{Yennie}.
307: Our intention is not to give a better model to describe the vector
308: meson production, but to explore the importance of Coulomb correction
309: effects. Therefore we use a schematic model, which captures some
310: essential features of the production process.
311: But it is clear that for a more realistic analysis a better
312: model should be used.
313:
314: Our model is the one proposed in \cite{Kopeliovich},
315: which is inspired by the vector dominance model and
316: which allows to derive a relatively simple form for the
317: vector meson production amplitude on a nucleus
318: with mass number $A$ and nuclear charge $Z$.
319:
320: We denote the energy momentum vectors of the initial and final electron
321: by $(\epsilon_{1,2},\vec{p}_{1,2})$ and
322: the momentum of the produced meson by
323: $\vec{p}_V$. $\vec{e}_V$ denotes the polarization vector of the meson.
324: The production amplitude is then given by
325: \begin{equation}
326: M(eA \rightarrow e'VA)= \int
327: \limits_{0}^{\infty} dx \, \vec{e}_V \cdot
328: \vec{f}(\vec{p}_1,\vec{p}_2,\vec{p}_V,x) ,
329: \end{equation}
330: where $\vec{f}=\vec{f}_1-\vec{f}_2$ and
331: \begin{displaymath}
332: \vec{f}_{1,2}(\vec{p}_1,\vec{p}_2,\vec{p}_V,x)=
333: \frac{1}{2 \omega_{1,2}} \frac{\partial}{\partial \omega_{1,2}}
334: \end{displaymath}
335: \begin{displaymath}
336: \times \int \frac{d^3 r}{r} \Bigl\{ \Bigl[ \epsilon_1 \vec{p}_2-\epsilon_2
337: \vec{p}_1 \Bigr] + \Bigl[ \epsilon_1 \vec{\nabla} \chi_2(\vec{r})
338: +\epsilon_2 \vec{\nabla} \chi_1(\vec{r}) \Bigr] \Bigr\}
339: \end{displaymath}
340: \begin{equation}
341: \times \exp[i \vec{\kappa} \vec{r} + i \chi_1(\vec{r})+i \chi_2
342: (\vec{r}) - i\omega_{1,2}r ]. \label{matrixel}
343: \end{equation}
344: For details concerning the derivation of eq. (\ref{matrixel}) we
345: refer to \cite{Kopeliovich} and the appendix of this paper.
346: We use a different sign convention for the eikonal phase than
347: \cite{Kopeliovich}.
348: The gradient terms are artefacts of the spinless treatment of
349: the electrons in \cite{Kopeliovich}
350: and therefore their physical significance in eq. (\ref{matrixel})
351: dubious at best.
352: However, from our discussion above follows that they can be neglected
353: for momentum transfer $Q^2=(\vec{p}_1-\vec{p}_2)^2-(\epsilon_1-\epsilon_2)
354: \gg R_A^{-2}$,
355: since they are of the order of $\alpha Z/R_A$, and
356: $| \epsilon_1 \vec{p}_2-\epsilon_2 \vec{p}_1 |= \sqrt{
357: \epsilon_1 \epsilon_2 Q^2}$.
358:
359: In (\ref{matrixel}), we have
360: \begin{displaymath}
361: \omega_1^2=(1-x)(\epsilon_1-\epsilon_2)^2-x(1-x)
362: \vec{p}_V^{\, 2} -2x/B,
363: \end{displaymath}
364: \begin{equation}
365: \omega_2^2=(1-x)((\epsilon_1-\epsilon_2)^2-m_V^2)^2-x(1-x)
366: \vec{p}_V^{\, 2} -2x/B \label{omegadef}
367: \end{equation}
368: and $\vec{\kappa}=\vec{p}_1-\vec{p}_2-x \vec{p}_V$.
369: Note that in eq. (50) in \cite{Kopeliovich}, the squares for
370: $\omega_{1,2}$ are missing and the terms containing $B$ are wrong.
371: We give therefore a detailed derivation of eqns. (\ref{matrixel},
372: \ref{omegadef})
373: in the appendix, which is missing in \cite{Kopeliovich}.
374: Note also the different sign in front of $\omega_{1,2}$ in
375: the exponent. The misprints will be corrected in the electronic
376: preprint version of the paper \cite{Kopeliovich2}.
377:
378: B is the slope parameter of the differential cross section.
379: For coherent electroproduction,
380: it is related to the mean charge nuclear radius squared
381: $\langle r^2_A \rangle_{rms}$ by (see e.g. \cite{Kopeliovich,Weise})
382: \begin{equation}
383: B=\frac{1}{3} \langle r^2_A \rangle_{rms} = \frac{1}{5} R_A^2.
384: \label{slope}
385: \end{equation}
386: At high energies and $\epsilon_{1,2} \gg \sqrt{Q^2}$,
387: the vectors $\vec{p}_1$, $\vec{p}_2$ and $\vec{p}_1-\vec{p}_2$ are
388: nearly parallel. Therefore we choose the $z$-axis along
389: $\vec{p}_1$ such that the vector $\vec{r}=(\vec{b},z)$ is given
390: by its $z$-component and the projection $\vec{b}$ to the
391: normal plane.
392: Using the relation
393: \begin{equation}
394: \int \limits_{-\infty}^{\infty} \frac{dz}{r} e^{i \vec{\kappa} \vec{r}-
395: i \omega r} = 2 K_0 \Bigl(b\sqrt{\kappa_L^2-\omega^2} \Bigr), \quad
396: \label{besselformula}
397: \end{equation}
398: where $K_0$ is the modified Bessel function, and $\kappa_{L,T}$
399: the longitudinal and transverse components of $\vec{\kappa}$
400: with respect to $\vec{p}_1$, we obtain ($\chi=\chi_1+\chi_2$)
401: \begin{equation}
402: \vec{f}_{1,2}=\frac{\epsilon_1 \vec{p}_2-\epsilon_2 \vec{p}_1}
403: {\omega_{1,2}} \frac{\partial}{\partial \omega_{1,2}}
404: \int \limits_0^\infty db \int \limits_0^{2 \pi}
405: d \varphi K_0 \Bigl( b \sqrt{\kappa_L^2-\omega_{1,2}^2} \Bigr)
406: e^{i \kappa_T b \cos(\varphi)+i \chi(b)}
407: \end{equation}
408: Eq. (\ref{besselformula}) is valid also for
409: $\kappa_L^2-\omega^2<0$, whereas in
410: \cite{Kopeliovich} the sign in front of $\omega$ is problematic.
411: From
412: \begin{equation}
413: \int \limits_0^{2 \pi}
414: d \varphi e^{i \kappa_T b \cos(\varphi)}=2 \pi J_0
415: (\kappa_Tb), \quad \frac{\partial}{\partial z} K_0(z)=-K_1(z)
416: \end{equation}
417: follows
418: \begin{equation}
419: \vec{f}_{1,2}=\frac{2 \pi (\epsilon_1 \vec{p}_2-\epsilon_2 \vec{p}_1)}
420: {\sqrt{\kappa_L^2-\omega_{1,2}^2}}
421: \int \limits_0^\infty db \, b K_1
422: \Bigl(b \sqrt{\kappa_L^2-
423: \omega_{1,2}^2} \Bigr) J_0(\kappa_T b) e^{i \chi(b)}.
424: \end{equation}
425: Kopeliovich {\em{et al.}} \cite{Kopeliovich}
426: performed calculations for a model potential
427: \begin{equation}
428: V_{mono}(r)=-\frac{\alpha Z}{r} e^{-\lambda r} (1-e^{-\mu r}),
429: \label{pot1}
430: \end{equation}
431: with an infrared cutoff $\exp(-\lambda r)$, and the last
432: factor corresponds to the monopole form of the nuclear form factor
433: via
434: \begin{equation}
435: \mu^{-2}=\frac{\langle r^2_A \rangle_{rms}}{6} .
436: \end{equation}
437: The eikonal phase can be calculated analytically for such
438: a potential. One obtains
439: \begin{displaymath}
440: \chi(\vec{r})=\chi_1(\vec{r})+\chi_2(\vec{r})=
441: -\int \limits_{-\infty}^{\infty} dz \, V_{mono}(\vec{b},z)=
442: \end{displaymath}
443: \begin{equation}
444: 2 \alpha Z \Bigl\{ K_0(\lambda b)-K_0[(\mu+\lambda)b] \Bigr\}.
445: \end{equation}
446:
447: We checked the result given in Fig. \ref{fig4} in
448: \cite{Kopeliovich} for small $\lambda$.
449: Our results show the same behavior, although we obtain slightly
450: smaller Coulomb corrections. This might be due to the fact that
451: the authors used formula (\ref{slope}) only for their Sommerfeld-Maue
452: calculations \cite{Kopeliovich2}.
453: We can reproduce a nearly identical curve
454: (Fig. \ref{fig1}),
455: if we reduce the slope parameter $B$ according to formula
456: (\ref{slope}) by a factor of 2.
457:
458: For the nuclear radius $R_A$
459: we used the formula
460: \begin{equation}
461: R_A=(1.128 \, \mbox{fm}) A^{1/3} + (2.24 \, \mbox{fm}) A^{-1/3},
462: \label{radiusrel}
463: \end{equation}
464: which is a good approximation for $A > 20$,
465: whereas the mass number and charge are related by
466: \begin{equation}
467: Z=\frac{A}{1.98+0.015 \, A^{2/3}}.
468: \end{equation}
469:
470: \begin{figure}[htb]
471: \centering
472: \includegraphics[width=12cm]{fig1.eps}
473: \caption{Coulomb correction for forward production
474: for the model potential used in \cite{Kopeliovich},
475: calculated for $\epsilon_1=100$ GeV,
476: $y=(\epsilon_1-\epsilon_2)/\epsilon_1=0.6$, and
477: electron scattering angle $1^o$ $(Q^2 \sim 1.2 \,
478: \mbox{GeV}^2)$,
479: with the slope parameter replaced by $B \rightarrow B/2$
480: (see text).}
481: \label{fig1}
482: \end{figure}
483:
484: \subsection{Improved Electrostatic Nuclear Potential}
485:
486: It is obvious that the potential given by eq. (\ref{pot1}) provides
487: only an inaccurate description of the electrostatic nuclear potential
488: near the nucleus.
489: We looked therefore for potentials which allow an analytic calculation
490: of the eikonal phase ($\chi$ and $\chi_{1,2}$ as well).
491: This is not a trivial problem, since the class of
492: meaningful analytic potentials
493: which allow to express the eikonal phase by known special
494: functions is rather restricted.
495:
496: A good choice is a potential energy for electrons of the form
497: \begin{equation}
498: (\alpha Z)^{-1}V_{model}(r)=-\frac{r^2+\frac{3}{2}R^2}{(r^2+R^2)^{3/2}}
499: -\frac{24}{25 \pi} \frac{R^2 R' r^4}{(r^2+R'^2)^4} \, ,
500: \label{pot_acc}
501: \end{equation}
502: which goes over into a Coulomb potential for $r \rightarrow
503: \infty$, and being close to the potential generated
504: by the relatively homogeneous spherical charge distribution of a nucleus.
505: The charge density
506: \begin{equation}
507: \rho(r)=-\frac{1}{e r} \partial^2_r (r V_{model}(r))
508: \end{equation}
509: corresponding to (\ref{pot_acc})
510: satisfies
511: \begin{equation}
512: \langle \rho \rangle = eZ \, ,
513: \end{equation}
514: \begin{equation}
515: \langle r^2 \rho \rangle = \frac{3}{5}R^2 eZ \, ,
516: \end{equation}
517: i.e. we can identify $R^2$ with the equivalent radius of the
518: homogeneously charged sphere which is given approximately by the formula
519: given above.
520: $R'$ serves as an additional fit parameter. A good choice is
521: $R'=0.5174 R$.
522:
523: Because the eikonal phase turns out
524: to be divergent for a Coulomb-like potential,
525: we regularize the eikonal phase by subtracting a screening potential
526: $V_{scr} \sim (r^2+a^2)^{-1/2}$ from (\ref{pot_acc}),
527: such that the potential
528: falls off like $r^{-2}$ for large $r$.
529: The divergence can then be absorbed
530: in a constant divergent phase $\sim \log(a)$ without
531: physical significance, when the limit $a \rightarrow
532: \infty$ is taken. It is instructive to calculate the
533: eikonal phase for the simple screened potential
534: \begin{equation}
535: V(r)=-\frac{\alpha Z }{\sqrt{r^2+R^2}} \, , \quad
536: V^a(r)=-\frac{\alpha Z }{\sqrt{r^2+R^2}}+
537: \frac{\alpha Z }{\sqrt{r^2+a^2}} \, .
538: \end{equation}
539: One obtains for a particle incident parallel to the z-axis
540: with impact parameter $b$ ($r^2=b^2+z^2$)
541: \begin{displaymath}
542: \chi_1^a=\alpha Z \int_{-\infty}^{z} dz \,
543: \Bigl( \frac{1}{\sqrt{r^2+R^2}}-
544: \frac{1}{\sqrt{r^2+a^2}} \Bigr)=
545: \end{displaymath}
546: \begin{equation}
547: \alpha Z \log \frac{(z+\sqrt{z^2+b^2+R^2})(b^2+a^2)}
548: {(z+\sqrt{r^2+a^2})(b^2+R^2)},
549: \end{equation}
550: and therefore for the regularized
551: eikonal phase
552: \begin{equation}
553: \chi_1=\lim_{a \to \infty} (\chi_1^a - \alpha Z \log(a))=
554: \alpha Z \log \Bigl( \frac{z+\sqrt{r^2+R^2}}{b^2+R^2} \Bigr) .
555: \end{equation}
556:
557: For a simple potential $\sim (r^2+R^2)^{-1/2}$, the
558: rms radius does not exist, since the corresponding
559: charge distribution does not fall off fast enough.
560:
561: For the potential $V_{model}$ (\ref{pot_acc}), the regularized eikonal phase
562: is given by
563: \begin{equation}
564: (\alpha Z)^{-1} \chi(b)=\log \Bigl( \frac{R^2+b^2}{R^2} \Bigr)
565: -\frac{R^2}{R^2+b^2}
566: -\frac{3}{50}\frac{R^2 R_1 (R_1^4+4 R_1^2 b^2 + 8 b^4)}
567: {(R_1^2+b^2)^{7/2}}
568: \end{equation}
569: The $z$-dependent formulae for $\chi_{1,2}$ are a bit lengthy,
570: but can be derived in a straightforward manner.
571:
572: In Fig. \ref{fig2}, we compare the potentials generated
573: by eq. (\ref{pot1}),
574: the potential $V_{hcs}$ of a homogeneously charged sphere with radius
575: $R_A$ and our model potential (\ref{pot_acc}) for
576: $^{208} Pb$ with identical mean squared radii in all three cases.
577:
578: \begin{figure}[htb]
579: \centering
580: \includegraphics[width=12cm]{fig2.eps}
581: \caption{Comparison of different model potentials
582: used in the calculations.}
583: \label{fig2}
584: \end{figure}
585:
586: \section{Results and Conclusions}
587: For the results presented in Figs. 3-5, we used the same kinematic
588: conditions as for Fig. \ref{fig1}.
589: Fig. \ref{fig3} compares the eikonal correction, using the simple
590: model potential $V_{mono}$ and the potential $V_{model}$ given by
591: eq. (\ref{pot_acc}).
592: It turns out that the Coulomb corrections are overestimated
593: by the model potential given by eq. (23). This is probably due
594: to the fact that the potential is too deep in the
595: central nuclear region. It is therefore mandatory to use
596: an accurate description for the nuclear Coulomb potential
597: in order to obtain reliable results for the Coulomb corrections.
598:
599: \begin{figure}[htb]
600: \centering
601: \includegraphics[width=12cm]{fig3.eps}
602: \caption{Comparison of Coulomb corrections for different
603: potentials, given by eq. (\ref{pot_acc}), solid line,
604: and eq. (\ref{pot1}), dashed line. Equal rms radii were
605: used in both cases.}
606: \label{fig3}
607: \end{figure}
608:
609: Fig. \ref{fig4} shows the Coulomb corrections for each element
610: with slope parameter $B$ according to eq. (\ref{slope}),
611: but with three different
612: rms charge radii for the electrostatic potential: the correct rms radius
613: $\sqrt \langle r^2_A \rangle_{rms}$,
614: a too large rms radius $2 \sqrt \langle r^2_A \rangle_{rms}$
615: and a too small rms radius $\frac{1}{2}\sqrt \langle r^2_A \rangle_{rms}$.
616: The figure clearly indicates that the distortion of the
617: electron waves is stronger for a small nucleus, whereas for
618: a larger nucleus, the eikonal phase varies less over the length
619: scale given by the nuclear radius (or slope parameter). The
620: initial and final state wave function resembles then more a plane
621: wave in the vicinity of the nucleus. The strong dependence of
622: the Coulomb corrections on the size of the nuclear radius clearly
623: indicates that the use of Sommerfeld-Maue wave functions
624: would lead to incorrect results.
625:
626: \begin{figure}[htb]
627: \centering
628: \includegraphics[width=12cm]{fig4.eps}
629: \caption{Coulomb corrections for the
630: correct rms radius (solid line), doubled radius
631: (dash-dotted), and radius divided by 2 (dashed).}
632: \label{fig4}
633: \end{figure}
634:
635: The dependence of the results on the slope parameter is displayed in
636: Fig. \ref{fig5} for $^{208} Pb$.
637: There we varied the slope parameter $B \rightarrow \lambda B$
638: for $\lambda=0.4 \ldots 1.6$, where $B$ is the theoretical value
639: given by eq. (\ref{slope}). The results show a clear dependence of the
640: Coulomb corrections on the model for the hadronic current.
641:
642: \begin{figure}[htb]
643: \centering
644: \includegraphics[width=12cm]{fig5.eps}
645: \caption{Coulomb corrections for different values $\lambda B$
646: of the slope parameter for $^{208} Pb$.}
647: \label{fig5}
648: \end{figure}
649:
650: Finally, Fig. \ref{fig6} shows the dependence of the Coulomb corrections
651: for $^{16}O$ and $^{208}Pb$, where we have varied artificially
652: the nuclear charge of the two elements,
653: while the correct charge radius of the two nuclei
654: and the corresponding slope parameter
655: were held fixed.
656:
657: \begin{figure}[htb]
658: \centering
659: \includegraphics[width=12cm]{fig6.eps}
660: \caption{Dependence of Coulomb corrections on the
661: nuclear charge, but with approximately correct nuclear radii
662: according to eq. (\ref{radiusrel}) and correct
663: slope parameter for lead and oxygen.}
664: \label{fig6}
665: \end{figure}
666:
667: The calculations presented in this paper treat only the case
668: where the scattering angle of the electron is small, such that
669: the expression for the vector meson production amplitude
670: can be reduced to a simple two-dimensional integral, which
671: can be solved without involving large computational efforts.
672: But it is also possible to perform
673: numerical calculations of the three-dimensional integral
674: representation of the amplitude on a modern workstation,
675: such that arbitrary scattering angles and more general models
676: of the hadronic current could be treated. This is the subject of
677: a forthcoming paper.
678:
679: The oscillatory behavior of the corrected cross section
680: shown in Fig. 3 in \cite{Kopeliovich} can not be reproduced by our
681: calculations. Fig. 4 shows clearly how important it is
682: to use the correct charge distribution of the nucleus for the
683: calculation of Coulomb corrections. Approximate wave functions
684: for pointlike nuclei are therefore not adequate.
685:
686: \section{Appendix A. Matrix Element for Electroproduction
687: of Vector Mesons}
688:
689: We start from the amplitude for coherent electroproduction of vector mesons
690: by electrons derived in \cite{Kopeliovich}, eq. (27), working with
691: the Born approximation first
692: \begin{equation}
693: M \propto \int \frac{d^3 q}{Q^2+i0} \Bigl[ \vec{j}(\vec{q})-j_0(\vec{q})
694: \frac{\vec{q}}{\nu} \Bigr] \vec{J}(Q,\vec{p}_V).
695: \end{equation}
696: where $\vec{j}$ is the spatial operator of the electron current,
697: $\vec{J}$ the hadronic current operator, $\nu=\epsilon_1-\epsilon_2$
698: and $Q^2=\vec{q}^{\, 2}-\nu^2$.
699: From current conservation
700: \begin{equation}
701: \nu j_0(\vec{q})=\vec{q} \cdot \vec{j} (\vec{q})
702: \end{equation}
703: we obtain
704: \begin{equation}
705: M \propto 2\frac{\epsilon_1 \vec{p}_2-\epsilon_2 \vec{p}_1}
706: {\epsilon_1-\epsilon_2} \int \frac{d^3q}{\vec{q}^{\, 2}-\nu^2}
707: \vec{J}(Q,\vec{p}_V).
708: \end{equation}
709: Using the hadronic model current ($\vec{\Delta}=\vec{q}-\vec{p}_V$ here)
710: \begin{equation}
711: \vec{J}(Q,\vec{p}_V)=\frac{\vec{e}_V m_V^2}{m_V^2+Q^2}
712: \frac{1}{1+B\vec{\Delta}^2/2}
713: \end{equation}
714: the amplitude becomes in Born approximation
715: \begin{equation}
716: M \propto 2\frac{(\epsilon_1 \vec{p}_2-\epsilon_2 \vec{p}_1)\vec{e}_V}
717: {\epsilon_1-\epsilon_2}
718: \int d^3 r \int d^3 q
719: \frac{1}{(\vec{q}^{\, 2}-\nu^2)(\vec{q}^{\, 2}-\nu^2+m_V^2)}
720: \frac{(2/B) m_V^2}{(2/B+\vec{\Delta}^2)}
721: e^{i(\vec{p}_1-\vec{p}_2-\vec{q})\vec{r}} \, .
722: \end{equation}
723: Going over to the eikonal approximation, we modify the electron current
724: by the eikonal phases. Therefore, we have to evaluate the integral
725: \begin{equation}
726: I=
727: m_V^2 \int d^3 r \int
728: \frac{d^3 q}{(\vec{q}^{\, 2}-\nu^2)(\vec{q}^{\, 2}-\nu^2+m_V^2)
729: (2/B+\vec{\Delta}^2)}
730: e^{i(\vec{p}_1-\vec{p}_2-\vec{q})\vec{r}+i \chi_1(\vec{r})+i \chi_2(\vec{r})}.
731: \end{equation}
732: Due to the trivial identity
733: \begin{equation}
734: \frac{m_V^2}{Q^2 (Q^2+m_V^2)}=\frac{1}{Q^2}-\frac{1}{Q^2+m_V^2},
735: \end{equation}
736: we can decompose integral $I=I_1+I_2$ according to
737: \begin{displaymath}
738: I_1=\int d^3 r \int \frac{d^3 q}{(\vec{q}^{\, 2}-\nu^2)
739: (2/B+\vec{\Delta}^2)} e^{i(\vec{p}_1-\vec{p}_2-\vec{q})\vec{r}+i \chi_1(\vec{r})+i \chi_2(\vec{r})},
740: \end{displaymath}
741: \begin{equation}
742: I_2=\int d^3 r \int \frac{d^3 q}{(\vec{q}^{\, 2}-\nu^2+m_V^2)
743: (2/B+\vec{\Delta}^2)} e^{i(\vec{p}_1-\vec{p}_2-\vec{q})\vec{r}+i \chi_1(\vec{r})+i \chi_2(\vec{r})}.
744: \end{equation}
745: With Feynman's trick
746: \begin{equation}
747: \frac{1}{\alpha \beta}=\int \limits_{0}^{1}
748: \frac{dx}{[\alpha x + \beta (1-x)]^2}
749: \end{equation}
750: follows for $I_1$ first
751: \begin{displaymath}
752: I_1=\int \limits_{0}^{1} dx
753: \int d^3 r \int d^3 q
754: \frac{e^{i(\vec{p}_1-\vec{p}_2-\vec{q})\vec{r}+i \chi_1(\vec{r})+i \chi_2(\vec{r})}}
755: {[(\vec{q}^{\, 2}-\nu^2)(1-x)+(2/B+(\vec{q}-\vec{p}_V)^2)x]^2}=
756: \end{displaymath}
757: \begin{equation}
758: \int \limits_{0}^{1} dx
759: \int d^3 r \int d^3 q
760: \frac{e^{i(\vec{p}_1-\vec{p}_2-\vec{q})\vec{r}+i \chi_1(\vec{r})+i \chi_2(\vec{r})}}
761: {[(\vec{q}-x \vec{p}_V)^{\, 2}-x^2 \vec{p}_V^2-(1-x) \nu^2 +2x/B+x
762: \vec{p}_V^{\, 2}]^2}.
763: \end{equation}
764: Shifting the integration variable $\vec{q} \rightarrow
765: \vec{q}+x \vec{p}_V$
766: leads to
767: \begin{displaymath}
768: I_1=\int \limits_{0}^{1} dx
769: \int d^3 r \int d^3 q
770: \frac{e^{i(\vec{p}_1-\vec{p}_2-x\vec{p}_V-\vec{q})
771: \vec{r}+i \chi_1(\vec{r})+i \chi_2(\vec{r})}}
772: {[\vec{q}^{\, 2}-(1-x) \nu^2+x(1-x)\vec{p}_V^{\, 2} +2x/B]^2}=
773: \end{displaymath}
774: \begin{equation}
775: \frac{1}{2}\int \limits_{0}^{1} dx
776: \int d^3 r \frac{1}{\omega_1} \frac{\partial}{\partial \omega_1}
777: \int d^3 q
778: \frac{e^{i(\vec{p}_1-\vec{p}_2-x\vec{p}_V-\vec{q})
779: \vec{r}+i \chi_1(\vec{r})+i \chi_2(\vec{r})}}
780: {\vec{q}^{\, 2}-\omega_1^2}.
781: \end{equation}
782: Finally, the identity
783: \begin{equation}
784: \int d^3 q \frac{e^{-i \vec{q} \vec{r}}}{\vec{q}^{\, 2}-\omega_1^2}=
785: \pi \frac{e^{-i \omega_1 r}}{r}
786: \end{equation}
787: immediately gives the final result in agreement with eq.
788: (\ref{matrixel})
789: \begin{equation}
790: I_1=\int \limits_{0}^{1} dx
791: \frac{\pi}{2 \omega_1} \frac{\partial}{\partial \omega_1}
792: \int \frac{d^3 r}{r} e^{i(\vec{p}_1-\vec{p}_2-x\vec{p}_V)\vec{r}
793: -i \omega_1 r+i \chi(\vec{r})}.
794: \end{equation}
795: For $I_2$ we must simply replace $\nu^2 \rightarrow \nu^2-m_V^2$.
796:
797: \begin{thebibliography}{99}
798: \bibitem{Aste} A. Aste, K. Hencken, D. Trautmann, G. Baur, Phys. Rev.
799: {\bf{A50}}, 3980 (1994).
800: \bibitem{Agger} C. K. Agger, A. H. Sorensen, Phys. Rev.
801: {\bf{A55}}, 402 (1997).
802: \bibitem{Sauter} F. Sauter, Ann. Phys. (Leipzig) {\bf{11}}, 454 (1931).
803: \bibitem{Oppenheimer} H. Hall, J. R. Oppenheimer, Phys. Rev. {\bf{38}},
804: 57-70 (1931).
805: %\bibitem{Hall} H. Hall, Revs. Modern Phys. {\bf{8}}, 358 (1936).
806: \bibitem{Boyer} R. H. Boyer, Phys. Rev. {\bf{117}}, 475 (1960).
807: \bibitem{Kopeliovich} B. Z. Kopeliovich, A. V. Tarasov, O.O. Voskresenskaya,
808: Eur. Phys. J. {\bf{A11}}, 345 (2001).
809: \bibitem{Kopeliovich2} B. Z. Kopeliovich, {\em{private communication}};
810: see also hep-ph/0105110.
811: \bibitem{Baker} A. Baker, Phys. Rev. {\bf{B134}}, 240 (1964).
812: \bibitem{Knoll} J. Knoll, Nucl. Phys. {\bf{A223}},
813: 462 (1974).
814: \bibitem{Lenz} F. Lenz, thesis, Freiburg, Germany, 1971.
815: \bibitem{Rosenfelder} F. Lenz, R. Rosenfelder, Nucl. Phys.
816: {\bf{A176}}, 513 (1971).
817: \bibitem{Donnachie}
818: A. Donnachie, P.V. Landshoff,
819: Phys. Lett. {\bf{B348}}, 213 (1995).
820: \bibitem{Weise} T. Renk, G. Piller, W. Weise,
821: Nucl. Phys. {\bf{A689}}, 869 (2001).
822: \bibitem{Yennie}
823: T.H. Bauer, R.D. Spital, D.R. Yennie, F.M. Pipkin,
824: Rev. Mod. Phys. {\bf{50}}, 261 (1978), Erratum-ibid. {\bf{51}}, 407 (1979).
825: \end{thebibliography}
826: \end{document}
827: