nucl-th0503082/ee.tex
1: \tolerance = 10000
2: \documentstyle[preprint,prc,aps,epsf]{revtex}
3: \begin{document}
4: \draft
5: \title{
6: Approximate treatment of electron Coulomb distortion in 
7: quasielastic $(e,e')$ reactions}
8: \author{ K.S. Kim,  L.E. Wright, and Yanhe Jin }
9: \address{
10: Institute of Nuclear and Particle Physics,
11: Department of Physics and Astronomy, Ohio University, Athens, 
12: Ohio 45701}
13: \author{ D.W. Kosik}
14: \address{Department of Physics, Butler University, Indianapolis, 
15: IN 46208 }
16: 
17: %\receipt{March 28, 1996}
18: \maketitle
19: \begin{abstract}
20: %12345678901234567890123456789012345678901234567890123456789012345678
21: In this paper we address the adequacy of various approximate methods 
22: of including Coulomb distortion effects in $(e,e')$ reactions by 
23: comparing to an exact treatment using Dirac-Coulomb distorted waves.
24: In particular, we examine approximate methods and analyses of 
25: $(e,e')$ reactions developed by Traini $et$ $al.$ using a high energy
26: approximation of the distorted waves and phase shifts due to Lenz and
27: Rosenfelder.
28: This approximation has been used in the separation of longitudinal 
29: and transverse structure functions in a number of $(e,e')$ 
30: experiments including the newly published $^{208}Pb(e,e')$ data from 
31: Saclay.
32: We find that the assumptions used by Traini and others are not valid 
33: for typical $(e,e')$ experiments on medium and heavy nuclei, and 
34: hence the extracted structure functions based on this formalism are 
35: not reliable.
36: We describe an improved approximation which is also based on the 
37: high energy approximation of Lenz and Rosenfelder and the analyses 
38: of Knoll and compare our results to the Saclay data.
39: At each step of our analyses we compare our approximate results to 
40: the exact distorted wave results and can therefore quantify the 
41: errors made by our approximations.
42: We find that for light nuclei, we can get an excellent treatment of 
43: Coulomb distortion effects on $(e,e')$ reactions just by using a 
44: good approximation to the distorted waves, but for medium and heavy 
45: nuclei simple additional $ad-hoc$ factors needs to be included.
46: We describe an explicit procedure for using our approximate analyses 
47: to extract so-called longitudinal and transverse structure functions 
48: from $(e,e')$ reactions in the quasielastic region.
49: \end{abstract}
50: \pacs{25.30.Fj 25.70.Bc}
51: \narrowtext
52: 
53: \section{Introduction}
54: %12345678901234567890123456789012345678901234567890123456789012345678
55: Medium and high energy electron scattering has long been acknowledged 
56: as a useful tool in the investigation of nuclear structure and 
57: nuclear properties, especially in the quasielastic region.
58: In the plane wave Born approximation (PWBA), where electrons are 
59: described as Dirac plane waves, the cross section for inclusive 
60: quasielastic $(e,e')$ processes can be written simply as
61: \begin{equation}
62: \frac{d^2\sigma}{d\Omega_e d\omega}= \sigma_{M}
63: \{ \frac{q^4_\mu}{q^4}  S_L(q,w) + [ \tan^2 \frac{\theta_e}{2} -
64: \frac{q^2_\mu}{2q^2} ]  S_T(q,w) \} \label{pwsep}
65: \end{equation}
66: where $q_\mu ^2 = \omega^2-{\bf q}^2$ is the four-momentum transfer,
67: $\sigma _{M}$ is the Mott cross section given by $\sigma_{M} = 
68: (\frac{\alpha }{2E} )^2  \frac{\cos^2
69: \frac{\theta}{2}}{\sin^4 \frac{\theta}{2}}$, and $S_L$ and $S_T$ 
70: are the longitudinal and transverse structure functions which depend 
71: only on the momentum transfer $q$ and the energy transfer $\omega$.
72: By keeping the momentum and energy transfers fixed while varying the 
73: electron energy $E$ and scattering angle $\theta_e$, it is possible 
74: to extract the two structure functions with two measurements.
75: However, when the electron wavefunctions are not Dirac plane waves, 
76: but rather are distorted by the static Coulomb field of the target 
77: nucleus, such a simple formulation as given in Eq.~(\ref{pwsep}) is 
78: no longer possible and, in general, the cross section does not 
79: separate into the sum of longitudinal and transverse structure 
80: functions with coefficients which only depend on the electron 
81: kinematics. 
82: Clearly the size of the Coulomb distortion effects depends on the 
83: charge of the target nucleus and on the energy of the electrons. 
84: In general, Coulomb effects are not too large near the peaks of 
85: cross sections, but can have greatly magnified  effects when one 
86: extracts ``structure functions'' by subtracting one cross section 
87: from another.
88: For example, in a recent distorted wave calculation~\cite{jin94} 
89: of $^{16}O(e,e'p)$ in the quasielastic region,  we found Coulomb 
90: effects on the extracted spectroscopic factors to be approximately 
91: 3\%, while the effects on the extracted fourth structure function 
92: was approximately 15\%.
93: 
94: However, some sort of extraction of structure functions, albeit 
95: approximate, is still very appealing since structure functions are 
96: sensitive to different aspects of the underlying knockout process or 
97: the final state interaction. Furthermore, $(e,e')$ reactions in the 
98: quasielastic region are particularly appealing in that the cross 
99: section is the sum of many knockout processes and some of the 
100: various Coulomb effects may be partially averaged out.
101: In fact, this is the case.
102: A rather simple approximation known as the Effective Momentum 
103: Approximation (EMA) where the electron momenta $p_{i,f}$ are 
104: modified by the value of the Coulomb potential at the center of the 
105: nucleus goes quite far in reproducing the Coulomb distortion effects 
106: for light nuclei.
107: However, the EMA for heavy nuclei does not adequately reproduce the 
108: DWBA cross section in the quasielastic region. 
109: 
110: During the past decade, $(e,e')$ cross sections have been 
111: measured~\cite{mez85}--~\cite{blat86} for a number of nuclei in the 
112: quasielastic region and either plane wave  or EMA was used to 
113: extract longitudinal and transverse structure functions.
114: These extracted structure functions were compared to the predictions 
115: of a simple Fermi-gas model, and in some cases, there appeared to be 
116: large suppression (up to about 40\%) of the longitudinal structure 
117: functions.
118: There were also disagreements with the extracted transverse 
119: structure functions and the predictions of the Fermi-gas model, but 
120: these were expected since exchange currents, pion production and 
121: other processes that are primarily induced by transverse photons 
122: were not included in the Fermi-gas model.
123: It should be noted that the Fermi-gas model is a rather crude 
124: description of a nucleus.
125: In particular, the shape of the structure functions at fixed 
126: momentum transfer plotted as a function of energy transfer is not 
127: well described.
128: On the other hand, we found that a ``single-particle'' relativistic 
129: model using relativistic Hartree wavefunctions coupled with the full 
130: DWBA treatment of Coulomb distortion for the electrons was in good 
131: agreement with the measured cross sections in the quasielastic 
132: region for $^{40}Ca$~\cite{jin92} and in reasonably good agreement 
133: for $^{238}U$.
134: Furthermore, the longitudinal structure function extracted for 
135: $^{40}Ca$ was well produced in magnitude and shape by this 
136: model~\cite{yates}.
137: 
138: More recently, Jourdan~\cite{jourdan} has examined the world data 
139: set for inclusive quasielastic scattering on $^{12}C$, $^{40}Ca$, 
140: and $^{56}Fe$ at momentum transfers of 300, 380 and 570 MeV/c, and 
141: has evaluated the Coulomb sum rule.
142: No evidence of suppression is found for the highest q value 
143: (570 MeV/c) where the sum rule is most model independent.
144: Jourdan used our Coulomb corrections in arriving at this conclusion.
145: Thus, on the basis of our direct comparison with the measured cross 
146: sections of $^{40}Ca$, and of Jourdan's results, the published 
147: analyses~\cite{saclay} for $^{208}Pb(e,e')$ which claim up to 50\% 
148: suppression of the longitudinal structure function is surprising.
149: However, the approximate treatment of Coulomb distortion used in the 
150: analysis is not accurate, and leads to doubts about the extracted 
151: structure functions.
152: As a separate matter, we question the  nuclear model used in making 
153: the claim of suppression.
154: 
155: In this paper we investigate the possibility of including Coulomb 
156: distortion effects in $(e,e')$ reactions in the quasielastic region 
157: for medium and heavy nuclei in an approximate way.
158: We have an advantage as compared to previous workers in that we have 
159: an exact treatment of the static Coulomb distortion of the target 
160: nucleus via a Distorted Wave Born Approximation (DWBA) calculation 
161: to which we can compare~\cite{jinphd}.
162: In Section II we will discuss various approximations that permit a 
163: ``plane-wave-like'' approach to the treatment of Coulomb distortion 
164: and compare the approximate results to the exact DWBA results in a 
165: step by step way.
166: We obtain an approximate potential due to the electron current which 
167: describes Coulomb distortion quite well.
168: In Section III, we apply this potential with further approximations 
169: to the particular case of inclusive quasielastic processes.
170: Finally we compare our calculations to the Saclay data~\cite{saclay} 
171: for the inclusive reaction $^{208}Pb(e,e')$.
172: 
173: In addition, we give an explicit procedure for extracting 
174: longitudinal and transverse structure functions from inclusive cross 
175: section data in the quasielastic region from medium and heavy nuclei, 
176: and make some general conclusions.
177: 
178: \section{Approximations}
179: \subsection{Approximation of the Electron Potential}
180: 
181: The four potential arising from the electron current $j_{\mu}$ is 
182: simply given in terms of the retarded Green's function by
183: \begin{equation}
184: A_{\mu}({\bf r}) = \int  G({\bf r}_{e}, {\bf r}) j_{\mu}({\bf r}_{e}) 
185: d{\bf r}_{e}
186: \label{pot0}
187: \end{equation}
188: where
189: \begin{eqnarray}
190: G({\bf r}_{e}, {\bf r}) = {\frac {e^{{\imath} {\omega} {\mid {\bf r}
191: _{e} - {\bf r} \mid}}} {\mid {\bf r}_{e} - {\bf r} \mid}} \nonumber 
192: \end{eqnarray}
193: and the electron current is given in terms of the initial and final 
194: electron wavefunctions by
195: \begin{equation}
196: j_{\mu}({\bf r}_{e}) = \overline{\psi}_f({\bf r}_{e}) \gamma_\mu 
197: \psi_i({\bf r}_{e}).
198: \end{equation}
199: For scattering processes this current extends over all space and 
200: thus the integral in Eq.~(\ref{pot0}) is not straightforward.
201: 
202: Knoll proposed~\cite{knol} a way of replacing the integral in 
203: Eq.~(\ref{pot0}) by a series of differential operators by using the 
204: transformation,
205: \begin{equation}
206: \int V({\bf r} -{\bf r'})f({\bf r'}) d{\bf r'} = e^{{\imath}{\bf q'}
207: {\cdot}{\bf r}}
208: \tilde{V}(-{\bf q'} +\imath {\bf \nabla}) e^{-{\imath}{\bf q'}
209: {\cdot}{\bf r}} f({\bf r})
210: \end{equation}
211: between the function $V({\bf r}-{\bf r'})$ and its Fourier transform
212: $\tilde{V}({\bf q'})$.
213: Applying this result to Eq.~(\ref{pot0}) and making a Taylor series 
214: expansion, we obtain the following series expansion for the potential,
215: \begin{equation}
216: A_{\mu} ({\bf r}) = {\frac {4 {\pi}} {q'^{2}-{\omega}^{2}}} [1 + 
217: ({\frac {q'^{2} + {\nabla}^{2}} {q'^{2} - {\omega}^{2}} }) + 
218: ({\frac {q'^{2} + {\nabla}^{2}} {q'^{2} - {\omega}^{2}} })^{2} +
219: {\ldots} ] j_{\mu} ({\bf r}). \label{pot}
220: \end{equation}
221: Note that while the momentum variable $q'$ is arbitrary, the choice 
222: affects the convergence of the series.
223: In particular, for the case of Dirac plane waves for the electron 
224: the only dependence of the electron current on $\bf r$ is simply 
225: $e^{{\imath}{\bf q}{\cdot}{\bf r}}$ and choosing ${\bf q}'={\bf q}$ 
226: results in the vanishing of all the terms except the first, and 
227: Eq.~(\ref{pot}) reduces to the well known M${\ddot{o}}$ller potential,
228: \begin{eqnarray}
229: A_{\mu}({\bf r})={\frac {4{\pi}} {q^{2}-{\omega}^{2}} } {\bar{u}}
230: ({\bf p}_{f}){\gamma}_{\mu}u({\bf p}_{i}) e^{{\imath}{\bf q}{\cdot}
231: {\bf r}}=a_{\mu}e^{{\imath}{\bf q}{\cdot}{\bf r}}, \nonumber
232: \end{eqnarray}
233: where $u$ is the familiar Dirac plane wave spinor.
234: 
235: As a test of this approximate procedure for calculating the potential,
236:  we calculated the electron charge distribution in the presence of 
237: the static Coulomb potential arising from the ground state charge 
238: distribution of $^{208}Pb$ using the partial wave solutions of the 
239: Dirac equation and evaluated the zeroth, zeroth plus first and zeroth
240: plus first and second terms in Eq.~(\ref{pot}).
241: Using these potentials we calculated the inelastic scattering cross 
242: section induced by a surface nuclear charge transition density
243: \begin{equation}
244: \rho^{if}_{n}({\bf r}) = {\frac {1} {R^{2}_{n}} } 
245: {\delta}(r-R_{n})Y^{M}_{L}({\hat r})
246: \end{equation}
247: where $R_{n}$ is the nuclear radius.
248: The ground state density of the nucleus was  described by a Fermi 
249: distribution of radius $R=6.65$ fm and total charge $Z=82$.
250: The cross sections were calculated at initial energy $E_{i}=400$ 
251: MeV and final energy $E_{f}=300$ MeV with energy transfer 
252: ${\omega}=100$ MeV.
253: In agreement with Knoll~\cite{knol}, we found that the first and 
254: second correction terms are sufficient to fill up the minima and the 
255: contribution of the second correction term was less than $2 \%$ for 
256: momentum transfer $q {\geq}\;350$ MeV/c.
257: We conclude that this high momentum approximation provides an 
258: alternative procedure for calculating the potential arising from 
259: inelastic electron scattering processes.
260: However, this procedure does require the numerical solution of the 
261: Dirac equation using partial waves which does take some computational 
262: time.
263: One advantage, however, is that the radial functions only have to be 
264: calculated out to about three times the nuclear radius.
265: 
266: We also applied this procedure to various approximate solutions of 
267: the Dirac equation and noted that the first and second order  terms 
268: were not well controlled if one approximates the electron current.
269: Clearly one should not approximate a function and then differentiate 
270: it.
271: In the following when we examine approximate electron wavefunctions, 
272: and hence approximate electron currents, we will only use the zeroth 
273: term in Eq.~(\ref{pot}), but will be sensitive to the choice of $q'$.
274: 
275: \subsection{High Energy Wavefunction Approximation}
276: 
277: The incoming Coulomb distorted electron scattering wave function 
278: that satisfies appropriate boundary conditions for an electron with 
279: spin $s_i$ can be expressed in the form of a partial wave sum 
280: by~\cite{uber} 
281: \begin{equation} 
282: \Psi_{i}^{s_{i}}({\bf r}) = 4\pi{\sqrt {\frac{E_{i} + m}{2E_{i}}} }
283: \sum_{\kappa , \mu}e^{i\delta_\kappa} i^lC^{\ l\ \ 1/2\ \ j}_
284: {\mu-s_{i}\ s_{i}\ \mu}\ {Y_l^{\mu-s_{i}}}^*({\bf \hat p}_{i}) 
285: \psi^\mu_\kappa ({\bf r}),\label{iwfn}
286: \end{equation}
287: where the spinor $\psi_{\kappa}^{\mu}$ is an eigenstate with
288: angular momentum quantum numbers $\kappa$ and $\mu$ given explicitly 
289: by 
290: \begin{equation}
291: \psi_\kappa^\mu({\bf r}) =\left[\begin{array}{l} f_\kappa
292: (r)  \chi^\mu_\kappa({\bf \hat r}) \\ ig_\kappa(r)  
293: \chi^\mu_{-\kappa}
294: ({\bf \hat r}) \end{array}\right], \label{wf1} 
295: \end{equation}
296: where the spin-angle functions are
297: \begin{equation}\chi^\mu_\kappa({\bf \hat r}) = \sum_{s} 
298: C^{\ l\ \ 1/2\ \ j}_{\mu-s\ s\ \mu}\ Y^{\mu-s}_l({\bf \hat r}) \chi_s. 
299: \label{wf2}
300: \end{equation}
301: The Dirac quantum number $\kappa$ determines the angular momentum 
302: labels for both $l$ and $j$.
303: Note that if we ignore the mass of the electron (valid for scattering 
304: angles away from extreme forward and backward angles),  the electron 
305: helicity is a good quantum number, and we only require positive 
306: $\kappa$ solutions.
307: The radial functions $f_\kappa$ and $g_\kappa$ and their
308: corresponding phase factors $\delta_\kappa$ are obtained by 
309: numerically solving the Dirac radial equation for a finite 
310: spherically symmetric nuclear charge distribution. 
311: The outgoing distorted wave function $\psi^{s_f}_f$ is found from 
312: Eq.~(\ref{iwfn}) by making the replacements $(i\delta_\kappa 
313: \rightarrow -i\delta_\kappa)$, $(s_i \rightarrow s_f)$,
314: $(E_{i} \rightarrow E_{f})$, and $({\bf p}_{i} \rightarrow 
315: {\bf p}_{f})$.
316: Note that by setting the phases $\delta_\kappa$ to zero and 
317: replacing $f_\kappa$ and $g_\kappa$ by the spherical Bessel 
318: functions $j_l(pr)$ and $sign(\kappa)j_{\bar l}(pr)$,  where $\bar 
319: l = l(-\kappa)$, the partial wave sum in Eq.~(\ref{wf1}) can be 
320: summed to give the Dirac plane wave solution,
321: \begin{equation}
322: \Psi_i({\bf r}) = {\sqrt {\frac {E_{i}+m} {2E_{i}}}} 
323: \left( \begin{array}{c}{\chi}_{s} \\ {\frac {{\mbox{\boldmath
324: $\sigma$}}{\cdot}{\bf p}} {E+m}}{\chi}_{s}
325: \end{array} \right)e^{i{\bf p}{\cdot}{\bf r}}, \label{plwv}
326: \end{equation}
327: where the electron spin label has been suppressed.  Thus, one way 
328: of obtaining a  plane-wave-like Coulomb distorted wavefunction, is 
329: to approximate the radial functions  $f_\kappa$ and $g_\kappa$  by 
330: spherical Bessel functions, and to approximate the scattering phase 
331: shifts  by an operator that can be pulled out of the partial wave 
332: sum.
333: Using these two ideas,  Lenz and Rosenfelder\cite{lenz} obtained an 
334: approximate scattering wavefunction for high energy electrons which  
335: includes Coulomb distortion for a finite nucleus in an approximate 
336: way more than twenty years ago.
337: 
338: In the high energy limit with good helicity $(E\gg m_{e})$, the 
339: approximate wavefunction of Lenz and Rosenfelder~\cite{lenz}, and 
340: Knoll~\cite{knol} can be written as
341: \begin{equation}
342: \Psi^{(\pm)} ({\bf r}) = {\eta} (r) e^{{\pm} {\imath} \delta 
343: ({\bf J}^{2})} e^{{\imath} {\bf p} {\cdot} {\bf r} {\eta}(r)} u_{p} 
344: \label{hea}
345: \end{equation}
346: where $u_{p}$ is the Dirac plane wave spinor and
347: \begin{equation}
348: {\eta}(r)={\frac 1 {pr}}\int_{0}^{r} [p-V(r')]dr'
349: +\mbox{correction term}.
350: \end{equation}
351: Lenz and Rosenfelder approximated the dependence of the phases on 
352: the operator ${\bf J}^2$ by $\delta({\bf J}^{2}) = \delta_{1/2} + 
353: b({\bf J}^2 -3/4)$ where $b$ depends on the Coulomb potential.
354: This particular approximation for the phase shifts is not necessary 
355: since  any function of the operator ${\bf J}^2$ still allows the 
356: partial wave sum to be carried out.
357: Lenz and Rosenfelder claimed that this equation is valid for 
358: ${\frac {{\mid}V(r){\mid}} {p}}{\ll}1$ and $j+1/2{\ll}pR$.
359: The second condition being primarily determined by the expression 
360: used for the phase shifts.
361: 
362: The approximate radial functions are given by spherical Bessel 
363: functions with modified argument,
364: \begin{eqnarray}
365: f_{\kappa}(r)&=&{\frac {x} {pr}}j_{l}(x) \nonumber \\
366: g_{\kappa}(r)&=&{\frac {x} {pr}}sign(\kappa)j_{{\bar l}}(x)
367: \end{eqnarray}
368: where $x=p^{\prime}(r)r$ and
369: \begin{equation}
370: p^{\prime}(r)=p-{\frac 1 r}{\int_{0}^{r}V(r)dr}+\frac{\Delta}{r} 
371: \label{loc}
372: \end{equation}
373: where $V(r)$ is the spherically symmetric Coulomb potential of the 
374: target nucleus and $\Delta$ is a correction term in the argument of 
375: order $\frac{\kappa^2}{(pr)^2}$ and is normally neglected.
376: 
377: To avoid having an $r$-dependent momentum, Traini $et$ 
378: $al.$~\cite{trai} made the further approximation that $p^{\prime}(r)
379: {\cong}p^{\prime}(0)=p-V(0)$ where $V(0)$ is the static Coulomb 
380: potential evaluated at the origin.
381: This approximation of the radial wavefunction leads to what is known 
382: as the Effective Momentum Approximation(EMA).
383: An approximation to the radial wavefunction where we keep the 
384: $r$-dependence in the momentum but still neglect $\Delta$  will be 
385: referred to as the Local Effective Momentum Approximation(LEMA).
386: In looking at these approximate wavefunctions, one should keep in 
387: mind that it is the spatial region around the nuclear surface that 
388: contributes most significantly to electron induced transitions.
389: That is, partial waves with angular momenta of order $pR$ where $R$ 
390: is the nuclear radius play a large role in the transition amplitude.
391: Thus, the approximation that the angular momentum is significantly 
392: less than $pr$ is not valid.  With this point in mind, we sought an 
393: $ad-hoc$ approximation for the correction $\Delta$.
394: We found that the following expression  describes the radial 
395: wavefunction at larger radial distances quite well,
396: \begin{eqnarray}
397: \Delta({\alpha}Z,E,{\kappa}^{2})=-{\alpha}Z({\frac {{16}{\kappa}} 
398: {E}})^{2}=a{\kappa}^{2}. \nonumber
399: \end{eqnarray}
400: The parameter $a=-{\alpha}Z({\frac {16} {E}})^{2}$ where the number 
401: 16 is given in MeV and was determined by comparing to the exact 
402: distorted radial wavefunction.
403: We will label this case  LEMA + $\Delta$.  Note that we have chosen 
404: the parametrization of $\Delta$ so that it contains $\kappa^2$ 
405: which can be expressed in terms of the operator $J^2$ so that we 
406: will still be able to sum the partial wave series.
407: 
408: To investigate these approximations we compare the radial 
409: wavefunctions calculated using EMA, LEMA, and LEMA + $\Delta$ to 
410: the exact Coulomb distorted waves (DW) for various angular momentum 
411: states for electron scattering from $^{208}Pb$.
412: The radius of $^{208}Pb$ is approximately 6.5 fm.
413: Figs.~(\ref{radfun5}) and ~(\ref{radfun15}) show the comparison 
414: between three approximate radial wavefunctions and the exact 
415: distorted radial wavefunction for different energies ($E=200$ MeV, 
416: $400$ MeV and $600$ MeV) and different $\kappa$ values($\kappa=5$ 
417: and $\kappa=15$).
418: The EMA wavefunction is acceptable at small radial distances, 
419: particularly much less than the nuclear radius, but at larger radii 
420: the approximate radial wavefunction is shifted too much to the left 
421: indicating the potential $V(0)$ at the origin is too large.
422: The LEMA wavefunction is a much better approximation to the exact 
423: wavefunction than EMA for medium and high energies out to radii 
424: beyond the nuclear radius, but at lower energies ( less than 250 
425: MeV), it also deviates significantly from the DW result.  
426: The wavefunction with the LEMA +$\Delta$ correction agrees with the 
427: distorted one almost perfectly above  400 MeV.
428: 
429: \begin{figure}[p]
430: \newbox\figa
431: \setbox\figa=\hbox{
432: \epsfysize=130mm
433: \epsfxsize=160mm
434: \epsffile{radfun5.eps}}
435: \noindent\hspace{0mm}\vspace{20mm}\box\figa
436: \caption{Radial wavefunctions in ${^{208}}Pb$ for ${\kappa}=5$. 
437: The top figure is for energy $E=200$ MeV, the middle for energy 
438: $E=400$ MeV and the bottom for energy $E=600$ MeV. The solid line 
439: is the exact DW wavefunction while the dash-dotted line is 
440: LEMA+$\Delta$, the dashed line is LEMA and the dotted line is EMA.}  
441: \label{radfun5}
442: \end{figure} 
443: 
444: At the lower energy this wavefunction is a bit to the right  of the 
445: distorted wavefunction, but the discrepancy is acceptably small.
446: Our conclusion is that LEMA is much better than EMA and may be 
447: acceptable for some reactions. The approximation LEMA + $\Delta$ 
448: furnishes quite a good description of the distorted wave radial 
449: functions out to several nuclear radii.
450: 
451: \begin{figure}[p]
452: \newbox\figb
453: \setbox\figb=\hbox{
454: \epsfysize=130mm
455: \epsfxsize=160mm
456: \epsffile{radfun15.eps}}
457: \noindent\hspace{0mm}\vspace{20mm}\box\figb
458: \caption{Radial wavefunctios in ${^{208}}Pb$ for ${\kappa}=15$. 
459: The same as Fig.~(\ref{radfun5}).}
460: \label{radfun15}
461: \end{figure}
462: 
463: The other important ingredient in Coulomb distortion considerations 
464: are the phase shifts ${\delta}_{\kappa}$.
465: In order to sum the partial wave series in Eq.~(\ref{iwfn}), the 
466: phase shifts in  Eq.~(\ref{hea}) were expressed as a function of the 
467: total angular momentum operator $J^{2}$.
468: The eigenvalues for this operator are $j(j+1)$ when operating on the 
469: partial waves, so the question is how well are the exact phase 
470: shifts reproduced by the expression~\cite{knol},
471: \begin{eqnarray}
472: {\delta}_{j}&=& {\delta}_{1/2}+b[j(j+1)-{\frac 3 4}]\nonumber \\
473: &=&{\delta}_{1/2}+b[{\kappa}^{2}-1] \label{phs1}
474: \end{eqnarray}
475: where we used the relation $j = \mid\kappa\mid-1/2$.
476: For a uniform charge distribution of radius R,
477: \begin{equation}
478: {\delta}_{{\frac 1 2}}=Z{\alpha}({\frac 4 3}-{\ln {2pR}})+b 
479: \end{equation}
480: and
481: \begin{equation}
482: b=-{\frac {3Z{\alpha}} {4{p(0)^{\prime}}^{2}R^{2}}}.
483: \end{equation}
484: We investigated this ${\kappa}^{2}$ approximation for the phase 
485: shifts for the Coulomb potential with ${^{208}}Pb$ and 400 MeV 
486: electrons.
487: The phase shifts are in good agreement with the exact phase shifts 
488: for small $\kappa$ values, but for large $\kappa$ values, the 
489: approximate phase shifts are much too large in magnitude.
490: The breakdown of the approximation occurs for ${\kappa}{\approx}pR$ 
491: as expected.
492: However, as noted earlier, it is these orbitals that play a dominant 
493: role in electron induced reactions from the nucleus.
494: In order to avoid this violation, we assume that the approximate 
495: phase shifts contain a ${\kappa}^{4}$-term expressed as
496: \begin{eqnarray}
497: {\delta}_{\kappa}&=&a_{0}+a_{2}{\kappa}^{2}+a_{4}{\kappa}^{4} 
498: \nonumber \\
499: &=&b_{0}+b_{2}[j(j+1)-{\frac 3 4}]+b_{4}[j(j+1)-{\frac 3 4}]^{2} 
500: \label{phs2}
501: \end{eqnarray}
502: where $b_{0}=a_{0}+a_{2}+a_{4}$, $b_{2}=a_{2}+2a_{4}$, and 
503: $b_{4}=a_{4}$.
504: Unfortunately, we do not have a simple analytical expression for 
505: the coefficients, so $a_{0}$, $a_{2}$ and $a_{4}$ are  fitted to the 
506: exact phase shifts calculated with a distorted wave code.
507: As is evident in Fig.~(\ref{phase}), including the ${\kappa}^{4}$ 
508: terms leads to a much better description of the exact phases.
509: 
510: 
511: \begin{figure}[p]
512: \newbox\figc
513: \setbox\figc=\hbox{
514: \epsfysize=130mm
515: \epsfxsize=160mm
516: \epsffile{phase.eps}}
517: \noindent\hspace{-5mm}\vspace{30mm}\box\figc
518: \caption{Comparison of the exact and ${\delta}({\kappa}^{2})$ fits 
519: to the phases in ${^{208}}Pb$ for the energy $E=400$ MeV and 
520: ${\kappa}_{max}=35$. 
521: The diamonds are the exact phases and the dashed line is the 
522: ${\kappa}^{2}$-fit to the phases and the solid line includes the 
523: ${\kappa}^{4}$ term in the fit.}
524: \label{phase}
525: \end{figure} 
526: 
527: Clearly higher powers of $\kappa^2$ could also be included if needed.
528: Of course, using the description of the phases given in 
529: Eq.~(\ref{phs2}) rather than the one in Eq.~(\ref{phs1}) requires 
530: calculation of the exact Coulomb phases for both the incoming and 
531: outgoing electron energy and then determining the coefficients $a_n$ 
532: by a fitting procedure.
533: However, the solution for the phases is straightforward and very 
534: rapid for modern computers, so this poses no real practical problem.
535: 
536: We have calculated various multipoles of the scalar potential in the 
537: partial wave formalism and confirmed that the approximate radial 
538: function and the approximate phases are in close agreement with the 
539: exact potential.
540: These two approximations permit the summation of the partial wave 
541: series of Eq.~(\ref{iwfn}) so that the electron wavefunction for 
542: incoming or outgoing waves can be written as
543: \begin{equation}
544: \Psi^{({\pm})}({\bf r})={\frac {p'(r)} {p}}e^{{\pm}i{\delta}({\bf
545: J}^{2})}e^{i(1+{\frac {\Delta} {p'(r)r}}){\bf p}'(r){\cdot}
546: {\bf r}}u_{p}.
547: \label{w-k1}
548: \end{equation}
549: The operator ${\bf J}^{2}$ is the square of the total angular 
550: momentum operator $({\bf J}={\bf L}+{\bf S})$, and previous 
551: workers~\cite{trai,giupac} have made a power series expansion of 
552: the phase term and applied successive terms to the plane-wave-like 
553: part of the wavefunction.
554: The problem with this procedure is that the phase shifts are greater 
555: than one for $\kappa$ values of importance and this series is very 
556: slowly converging. 
557: It is straightforward to check that keeping only the first three 
558: terms in the exponential expansion leads to significant errors. 
559: We choose not to make this expansion, but to approximate $J^2$ by 
560: the orbital angular momentum squared $L^2$ and further to replace 
561: $L^2$ by its classical value 
562: $({\bf r}{\times}{\bf p}(r)^{\prime})^{2}$. 
563: Clearly we are neglecting the spin dependence of the phase shifts 
564: by this classical approximation, but since the processes we are 
565: interested in are dominated by angular momentum values around 10 or more, 
566: we expect the spin dependence to be negligible. 
567: We confirmed this estimate by comparisons of our partial wave calculation
568: of $(e,e'p)$ where the full spin dependence is included to a three
569: dimensional numerical integration using the classical approximation.
570: In like manner, we also replace the $\kappa^2$ dependence in 
571: $\Delta$ by $({\bf r}{\times}{\bf p}(r)^{\prime})^{2}$.
572: Finally, the approximate Coulomb distorted wavefunction of 
573: Eq.~(\ref{w-k1}) with the $\Delta$ correction is given explicitly 
574: by~\cite{kimphd}
575: \begin{equation}
576: \Psi^{({\pm})}({\bf r})={\frac {p^{\prime}(r)} {p}}e^{{\pm}i{\delta}
577: (({\bf r}{\times}{\bf p}^{\prime}(r))^{2})}e^{ia({\hat p}^{\prime}
578: (r){\cdot}{\hat r})({\bf r}{\times}{\bf p}^{\prime}(r))^{2}}
579: e^{i{\bf p}^{\prime}(r){\cdot}{\bf r}}u_{p}.
580: \label{w-k2}
581: \end{equation}
582: with ${\delta}(({\bf r}{\times}{\bf p}'(r))^{2})=b_{0}+b_{2}
583: ({\bf r}{\times}{\bf p}'(r))^{2}+b_{4}({\bf r}{\times}
584: {\bf p}'(r))^{4}$.
585: Using this wavefunction and the first term of Eq.~(\ref{pot}) we 
586: obtain the following four potential, which includes in an 
587: approximate way, the Coulomb distortion of the target nucleus
588: \begin{equation}
589: A^{\mu}({\bf r})={\frac {4{\pi}} {4p_{i}p_{f}{\sin^4 {\frac 
590: {{\theta}_{e}} {2}}}}}e^{i[{\delta}_{i}(({\bf r}{\times}{\bf p}'
591: _{i}(r))^{2})+{\delta}_{f}(({\bf r}{\times}{\bf p}'_{f}(r))^{2})]}
592: e^{i({\Delta}_{i}-{\Delta}_{f})}e^{i{\bf q}'(r){\cdot}{\bf r}}
593: {\bar u}_{f}{\gamma}^{\mu}u_{i} \label{apppot}
594: \end{equation}
595: where $\Delta=a({\bf r}{\times}{\bf p}'(r))^{2}({\hat r}{\cdot}
596: {\hat p}'(r))$ and ${\bf q}'(r)={{\bf p}'}_{i}(r)-{{\bf p}'}_{f}(r)$.
597: The approximate potential of Eq.~(\ref{apppot}) is similar to the 
598: plane wave result except for the phase factors and the radial 
599: dependence in the momentum transfer.
600: Unfortunately, the spatial dependence in the phase factors makes a 
601: multipole decomposition of this potential impractical.
602: However, since it is an analytical function it is straightforward 
603: to calculate interaction matrix elements by performing the three 
604: dimensional integration over ${\bf r}$ numerically.
605: 
606: \section{Application to the Inclusive Process}
607: 
608: For the inclusive cross section $(e,e')$, the longitudinal and 
609: transverse structure functions in Eq.~(\ref{pwsep}) are bi-linear 
610: products of the Fourier transform of the components of the nuclear 
611: transition current density integrated over outgoing nucleon angles.
612: Furthermore, it is the Dirac structure of the M\"{o}ller 
613: potential which leads to the cross section containing one term with 
614: only longitudinal components of the current and a second containing 
615: only transverse components.
616: However, the Dirac structure of the approximate potential in 
617: Eq.~(\ref{apppot}) is the same as the plane wave result.
618: Therefore, even with Coulomb distribution included, albeit in an 
619: approximate way, the separation of the cross section into a 
620: longitudinal term and a transverse term persists. 
621: Explicitly, the structure functions for knocking out nucleons from a 
622: shell with angular momentum $j_{b}$ are given by
623: \begin{eqnarray}
624: S_{L}(q,{\omega})&=&\sum_{{\mu}_{b}s_{P}}{\frac {{\rho}_{P}} 
625: {2(2j_{b}+1)}} \int {\mid}N_{0}{\mid}^{2}d{\Omega}_{P} \\
626: S_{T}(q,{\omega})&=&\sum_{{\mu}_{b}s_{P}}{\frac {{\rho}_{P}} 
627: {2(2j_{b}+1)}} \int
628: ({\mid}N_{x}{\mid}^{2}+{\mid}N_{y}{\mid}^{2})d{\Omega}_{P}
629: \end{eqnarray}
630: where the nucleon density of states ${\rho}_{P}={\frac {pE_{p}} 
631: {(2\pi)^{2}}}$, the $z$-axis is taken to be along ${\bf q}$, and 
632: ${\mu}_{b}$ and $s_{P}$ are the z-components of the angular momentum 
633: of the bound and continuum state particles.
634: The Fourier transfer of the nuclear current $J^{\mu}({\bf r})$ is 
635: simply,
636: \begin{equation}
637: N^{\mu}=\int J^{\mu}({\bf r})e^{{\imath}{\bf q}{\cdot}{\bf r}}d^{3}r.
638: \end{equation}
639: The continuity equation has been used to eliminate the $z$-component 
640: ($N_{z}$) via the equation $N_{z}=-{\frac {\omega} {q}}N_{0}$.
641: Note that when we use the approximate electron four potential along 
642: with current conservation to eliminate the z-component of the current 
643: we run into a problem since the momentum transfer ${\bf q}^{\prime}$ 
644: depends on r both in magnitude and direction.
645: In addition, the phase factors depend on ${\bf r}$.
646: To avoid generating additional terms we assume the direction of 
647: ${\bf q}^{\prime}(r)$ is along the asymptotic momentum transfer 
648: ${\bf q}$ which defines the ${\hat z}$-axis, and neglect the 
649: dependence on ${\bf r}$ in the phases and in ${\bf q}^{\prime}(r)$, 
650: when taking the divergence of ${\bf N}$.
651: With this further approximation, current conservation implies 
652: ${\omega}N_{0}+{\bf q}^{\prime}(r){\cdot}{\bf N}=0$.
653: Using the approximate potential of Eq.~(\ref{apppot}), the cross 
654: section with for the inclusive reaction $(e,e')$ can be written as
655: \begin{equation}
656: \frac{d^2\sigma}{d\Omega_e d\omega}= \sigma_{M}
657: \{ \frac{q^4_\mu}{q^4}  S_L(q',w) + [ \tan^2 \frac{\theta_e}{2} -
658: \frac{q^2_\mu}{2q^2} ]  S_T(q',w) \} 
659: \end{equation}
660: and the transform of the transition nuclear current elements which 
661: appears in $S_{L}$ and $S_{T}$ are given by
662: \begin{eqnarray}
663: N_{0}&=&\int ({\frac {q'_{\mu}(r)} {q_{\mu}}})^{2} 
664: ({\frac {q} {q'(r)}})^{2}e^{i{\delta}_{f}([{\bf r}{\times}
665: {\bf p}_{i}^{\prime}(r)]^{2})}e^{i{\delta}_{f}([{\bf r}{\times}
666: {\bf p}_{f}^{\prime}(r)]^{2})}e^{i({\Delta}_{i}-{\Delta}_{f})}
667: e^{i{\bf q}^{\prime}(r){\cdot}{\bf r}}J_{0}({\bf r})d^{3}r 
668: \label{apn0}\\
669: {\bf N}_{T}&=&\int e^{i{\delta}_{i}([{\bf r}{\times}
670: {\bf p}_{i}^{\prime}(r)]^{2})}e^{i{\delta}_{f}([{\bf r}{\times}
671: {\bf p}_{f}^{\prime}(r)]^{2})}e^{i({\Delta}_{i}-{\Delta}_{f})}
672: e^{{\imath}{\bf q}^{\prime}(r){\cdot}{\bf r}}J_{T}({\bf r})d^{3}r 
673: \label{apnt}
674: \end{eqnarray}
675: 
676: As noted earlier, a multipole expansion of the approximate potential 
677: is not practical, and since the inclusive reaction $(e,e')$ requires 
678: a sum over all occupied neutron and proton shells, numerical 
679: integration is very time consuming.
680: In order to have a more practical procedure we choose to make 
681: additional approximations.
682: Firstly, we neglect all of the phase factors($\delta$ and $\Delta$) 
683: in Eqs.~(\ref{apn0}) and ~(\ref{apnt}) but retain the $r$-dependence 
684: in ${\bf q}^{\prime}$.
685: This returns us to the approximation we call LEMA.
686: For light to medium nuclei this approximation is in good agreement 
687: with the full DWBA result regarding the shape as a function of 
688: energy transfer, but has a small discrepancy in magnitude. 
689: The magnitude is corrected with an overall factor of 
690: $({\frac {p_{i}'(0)} {p_{i}}})^{2}$ in the cross section.
691: However, for heavier nuclei, we noticed that LEMA with the magnitude 
692: factor was a very good 
693: approximation to the DWBA results for large electron scattering 
694: angle where the transverse term dominates, but deviated significantly 
695: for forward electron angles where the longitudinal term has a 
696: significant contribution.
697: Thus it appears that the phase factors play a significant role in 
698: the longitudinal term for large Coulomb distortion.
699: 
700: We also noticed that for forward electron angles the low energy 
701: side of the DWBA quasielastic peak looks similar to the plane wave 
702: result.
703: Thus, it appears that in the longitudinal term the phase factors are 
704: partially cancelling the effect of the effective momentum 
705: ${\bf q}^{\prime}(r)$ on the low $\omega$ side of the quasielastic
706:  peak.
707: We examined a number of simple $ad-hoc$ modifications to the longitudinal
708: term as described by LEMA and 
709: the magnitude factor in 
710: order to reproduce the DWBA forward angle results for $(e,e')$ 
711: reactions on $^{208}Pb$ in the quasielastic region.
712: Based on a number of trials we propose the following ``Fourier'' transform for the charge 
713: component of the current 
714: \begin{equation}
715: N^{LEMA^{\prime}}_{0}=({\frac {p_{i}'(0)} {p_{i}}})
716: \int e^{i{\bf q}''(r){\cdot}{\bf r}}J_{0}
717: ({\bf r})d^{3}r \label{lemal'}
718: \end{equation}
719: where $({\frac {p_{i}'(0)} {p_{i}}})$ is the magnitude 
720: enhancement and ${\bf q}''(r)={\bf p}_{i}''(r)-{\bf p}_{f}''(r)$, 
721: $p''(r)=p-{\frac {\lambda} {r}}\int_{0}^{r}V(r')dr'$ and the 
722: factor $\lambda$, which depends on the energy transfer $\omega$, 
723: is given by $\lambda=({\omega}/{\omega}_{0})^{2}$ with 
724: ${\omega}_{0}={\frac {q^{2}} {1.4M}}$.
725: Clearly for small $\omega$, ${\bf q}''$ approaches the asymptotic 
726: value, while for $\omega$ past the quasielastic peak which is located 
727: approximately at ${\omega}_{0}$ in the cross section, the effective 
728: momentum differs significantly from the asymptotic value.
729: We have tested this $ad-hoc$ prescription for 
730: $0.3{\leq}{\omega}/{\omega}_{0}{\leq}2.0$ for a range of energies 
731: and nuclei and find excellent agreement with the DWBA result.
732: The transverse ``Fourier'' transform only contains the normalization  
733: factor and is given by
734: \begin{equation}
735: {\bf N}^{LEMA^{\prime}}_{T}=({\frac {p_{i}'(0)} {p_{i}}})
736: \int e^{i{\bf q}'(r){\cdot}{\bf r}}J_{T}({\bf r})
737: d^{3}r. \label{lemat'}
738: \end{equation}
739: We will refer to the cross section for the inclusive reaction 
740: calculated with these two structure functions as the 
741: LEMA$^{\prime}$ result.
742: Clearly $N^{LEMA^{\prime}}_{0}$ and $N^{LEMA^{\prime}}_{T}$ 
743: represent a modified Fourier transform 
744: of the nuclear transition current.
745: The approximation known as the EMA replaces ${\bf q}'(r)$ with 
746: ${\bf q}'(0)$ wherever it appears in Eqs.~(\ref{apn0}) and 
747: ~(\ref{apnt}) for $N_{0}$ and $N_{T}$ and the phases are neglected 
748: as usual.
749: We find that for light nuclei the EMA is adequate, but it leads to 
750: large errors for nuclei as heavy as $^{208}Pb$.
751: 
752: In our analyses of quasielastic scattering, we use relativistic 
753: bound and continuum single particle wavefunctions.
754: For the inclusive reaction we use continuum solutions for the 
755: outgoing, but unobserved, nucleon which are in the same Hartree 
756: potential as the bound state orbitals~\cite{jin92}.
757: This choice insures charge conservation and gauge invariance.
758: Thus, in this relativistic single particle model the nuclear 
759: current matrix element is
760: \begin{equation}
761: J^{\mu}({\bf r})=e{\bar {\Psi}}_{P}{\hat J}^{\mu}{\Psi}_{b}
762: \end{equation}
763: where we use the free nucleon current operator 
764: \begin{equation}
765: {\hat J}^{\mu}=F_{1}{\gamma}^{\mu}+F_{2}{\frac {i{\mu}_{T}} 
766: {2m_{N}}}{\sigma}^{{\mu}{\nu}}q_{\nu}
767: \end{equation}
768: where $\mu_{T}$ is the nucleon anomalous magnetic moment 
769: (for proton $\mu_{T}=1.793$ and for neutron $\mu_{T}=-1.91$).
770: The form factor $F_{1}$ and $F_{2}$ are related to the electric 
771: and magnetic form factors $G_{E}$ and $G_{M}$ by
772: \begin{eqnarray}
773: G_{E}&=&F_{1}+{\frac {{\mu}_{T}q_{\mu}^{2}} {4M^{2}}}F_{2} \\
774: G_{M}&=&F_{1}+{\mu}_{T}F_{2}
775: \end{eqnarray}
776: We choose the standard result:
777: \begin{equation}
778: G_{E}=G_{M}/({\mu}_{T}+1)=(1-q^{2}_{\mu}/0.71)^{-2}
779: \end{equation}
780: where in this formula $q_{\mu}^{2}$ is in units of GeV$^{2}$.
781: 
782: In Fig.~(\ref{cros}), we compare various approximations to the DWBA 
783: result as a function of the energy transfer $\omega$ for two cases, 
784: incident electron energy $E_{i}=310$ MeV, scattering angle 
785: $\theta_{e}=143^{o}$ and $E_{i}=485$ MeV, scattering angle 
786: $\theta_{e}=60^{o}$ data sets.
787: The dotted line is the PWBA result, the diamonds are the DWBA result, 
788: the dash-dotted line is the EMA result, and the solid line is the 
789: LEMA$^{\prime}$ result.
790: We notice that the EMA result is always lower than the DWBA result 
791: although the peak position is approximately in the right place.
792: This lack of agreement with the EMA is not too surprising based on 
793: our previous examination of the wavefunction.
794: Replacing the average value of the Coulomb potential between the 
795: origin and the position $r$ by the value at the origin is too large 
796: an error for $r$ near the nuclear surface where most of the 
797: interaction takes place.
798: Since the EMA is such a bad approximation to the full DWBA result, 
799: it should not be used as a basis for including the Coulomb phase terms in 
800: $(e,e')$ or $(e,e'p)$ reactions.
801: Over the whole region, the LEMA$^{\prime}$ result is in excellent 
802: agreement with the DWBA result apart from the extreme wings of the 
803: quasielastic peak.
804: The difference around the peak is less than 2 $\%$ and side parts are 
805: about 5 $\%$.
806: 
807: \begin{figure}[p]
808: \newbox\figda
809: \setbox\figda=\hbox{
810: \epsfysize=120mm
811: \epsfxsize=155mm
812: \epsffile{cros.eps}}
813: \noindent\hspace{5mm}\vspace{10mm}\box\figda
814: \caption{The differential cross section for ${^{208}}Pb(e,e')$ 
815: at two different electron energies and scattering angles. 
816: The dotted line is the PWBA result, the diamonds are the DWBA result, 
817: the dash-dotted line is the EMA result, and the solid line is the 
818: LEMA$^{\prime}$ result.}
819: \label{cros}
820: \end{figure}
821: 
822: We illustrate our approximations for other kinematics by calculating 
823: the cross section at fixed momentum transfer $q=425$ MeV/c with 
824: three different electron scattering angles (${\theta}_{e}=60^{0}$,
825: $90^{0}$, and $143^{0}$) as shown Fig.~(\ref{croqfx}). 
826: The EMA result is always lower than the DWBA and the LEMA$^{\prime}$ 
827: results and is shifted toward large energy transfer by about 10 MeV. 
828: LEMA$^{\prime}$ again reproduces the DWBA cross sections quite well.
829: 
830: \begin{figure}[p] 
831: \newbox\figga
832: \setbox\figga=\hbox{
833: \epsfysize=120mm
834: \epsfxsize=155mm
835: \epsffile{qfx.eps}}
836: \noindent\hspace{5mm}\vspace{5mm}\box\figga
837: \caption{The differential cro10s section for $^{208}Pb$ at constant
838: momentum transfer $q=425$ MeV/c, but three different electron 
839: scattering angles.}
840: \label{croqfx}
841: \end{figure}
842: 
843: \section{Separation Procedure and Structure Functions}
844: 
845: Since the LEMA$^{\prime}$ cross section for $(e,e')$ has the same 
846: structure as the plane wave result, a Rosenbluth-type separation 
847: can be used to extract a longitudinal and a transverse contribution 
848: even in the presence of large Coulomb distortions.
849: In fact,  experimental evidence for this was given in one of the 
850: early papers on
851: quasielastic scattering~\cite{blat86} where cross section 
852: measurements on $^{238}U$ at three different electron scattering 
853: angles, but with the same energy and momentum transfer, fell on a 
854: straight line when a Rosenbluth plot was made.
855: Our LEMA$^{\prime}$ approximation gives a theoretical explanation 
856: for this observation.
857: Of course the separated structure functions are no longer bi-linear 
858: products of the simple Fourier transforms of the current components 
859: integrated over outgoing nucleon directions.
860: Inclusion of Coulomb distortion within LEMA$^{\prime}$ results in an 
861: $r$-dependent Fourier momentum variable which differs for the 
862: longitudinal and transverse case.
863: We have only been able check LEMA$^{\prime}$ for our particular model 
864: of the quasielastic process, so we can not prove that it applies to 
865: other models.  
866: However, based on previous work we know our model describes $(e,e'p)$ 
867: and $(e,e')$ quite well in the quasielastic region and its treatment 
868: of the spatial dependence in the nuclear charge and current is 
869: very realistic.
870: Thus we believe that the Coulomb corrections that we calculate 
871: will be appropriate for any realistic nuclear model.
872: We conclude that LEMA$^{\prime}$ is a good approximation for 
873: the inclusive cross section $(e,e')$ in the quasielastic region.
874: 
875: To illustrate the ``Rosenbluth'' separation we write 
876: $S = S_L + x S_T$ where S is the total structure function given by
877: \begin{equation}
878: S=({\frac {q} {q_{\mu}}})^{4} {\frac 1 {{\sigma}_{M}}}
879: {\frac {d^{2}{\sigma}} {d{\Omega}_{e}d{\omega}}}
880: \end{equation}
881: and $x=[ {\tan^2} {\frac{{\theta}_{e}} {2}} - {\frac {q^{2}_{\mu}} 
882: {2q^{2}}} ]/{\frac {q^{4}_{\mu}} {q^{4}}}$.
883: 
884: In Fig.~(\ref{sep}), we compare the structure functions extracted 
885: using our calculations of $(e,e')$ on $^{208}Pb$ using various 
886: treatments of Coulomb distortion for a momentum transfer of 
887: $q=425$ MeV/c and three different values of energy transfer around 
888: the peak of the cross section.
889: In each case we find a very good fit to a straight line.
890: Furthermore, we see that the intercept and slope extracted using the 
891: full DWBA and LEMA$^{\prime}$ calculations agree within $2\%$ in 
892: all cases.
893: The EMA and plane wave (PWBA) results clearly are in disagreement 
894: with the DWBA results.
895: \begin{figure}[p] 
896: \newbox\figxa
897: \setbox\figxa=\hbox{
898: \epsfysize=200mm
899: \epsfxsize=155mm
900: \epsffile{sep.eps}}
901: \noindent\hspace{5mm}\vspace{10mm}\box\figxa
902: \caption{Rosenbluth separation plot of the cross section for 
903: $^{208}Pb$ at $q=425$ MeV/c.
904: The solid line is the DWBA result, dashed line for the 
905: LEMA$^{\prime}$, the dotted line for the PWBA, and the dash-dotted 
906: line for the EMA. The unit of the total structure function is 
907: MeV$^{-1}$.}
908: \label{sep}
909: \end{figure}
910: 
911: Given this result there are several ways to analyse experimental 
912: $(e,e')$ data.
913: If one has a model for the process under investigation, one can 
914: calculate the $r$-dependent Fourier transforms of Eqs.~(\ref{lemal'}) 
915: and ~(\ref{lemat'}) and compare to  the experimentally measured 
916: cross section.
917: Or, one could make a Rosenbluth separation to obtain the 
918: LEMA$^{\prime}$ structure functions as a function of energy 
919: transfer $\omega$ and asymptotic momentum transfer $q$.
920: Although these structure functions clearly have some dependence on 
921: electron energy and scattering angle in addition to their 
922: dependence on $q$ and $\omega$, in our model of the quasielastic 
923: process this dependence is not very strong and we recommend that 
924: in comparing a theoretical model with the extracted LEMA$^{\prime}$ 
925: structure functions that electron energies and angles in the same 
926: range as used in the experiment be used.
927: 
928: Lacking a model for the process being measured, one could take a  
929: simple model with suitable geometry and calculate the ratio of the 
930: cross section calculated with PWBA to one calculated with 
931: LEMA$^{\prime}$ and renormalize the measured cross section data by 
932: multiplying by this ratio.
933: The resulting pseudo PWBA ``data'' could then be separated by a 
934: Rosenbluth procedure and would produce structure functions in terms 
935: of the Fourier transforms of the current components.
936: Clearly this procedure introduces some error if the model amount of 
937: longitudinal and transverse contributions are not close to the 
938: amounts in the process being measured since LEMA$^{\prime}$ treats 
939: them somewhat differently.
940: However, this difference is not too large  and seems to be the best 
941: one can do.
942: 
943: Finally one could assume that the theoretical model has the correct 
944: kinematics and spatial dependence, but that the magnitude of the 
945: longitudinal and/or transverse portions of the model is not correct.
946: That is, one could use the model to calculate the longitudinal and 
947: transverse contributions to the cross section and then multiply each 
948: by a scale factor to be determined by making a least squares fit to 
949: the cross section data.
950: 
951: \section{Comparison With Experimental Data }
952: 
953: In Fig~(\ref{clesa}), we compare our theoretical results based on 
954: the relativistic ``single particle'' model~\cite{jin92} calculated 
955: with LEMA$^{\prime}$ to the Saclay data~\cite{saclay} for several 
956: electron angles (${\theta}_{e}=143^{0}, 90^{0}$, $60^{0}$ and 
957: $35^{0}$).
958: The solid line is our model result for the cross section while the 
959: dotted line shows the longitudinal contribution to the cross section.
960: Clearly the longitudinal contribution is quite small except for the 
961: forward electron scattering angle of $35^{0}$ where it represents 
962: slightly over $50\%$ of the cross section.
963: Pion production is not included in our model, so the behavior at 
964: large $\omega$ is not expected to agree with the data.
965: Clearly the agreement between the data and the calculation is not 
966: very good, which can be contrasted with quite good agreement between 
967: our calculations and the $(e,e')$ data from Bates on 
968: $^{40}Ca$~\cite{yates}.
969: Further, these may be a suggestion of longitudinal suppression if 
970: coupled with a transverse enhancement. 
971: \begin{figure}[p]
972: \newbox\figha
973: \setbox\figha=\hbox{
974: \epsfysize=130mm
975: \epsfxsize=130mm
976: \epsffile{sacros.eps}}
977: \noindent\hspace{20mm}\vspace{10mm}\box\figha
978: \caption{The differential cross section for ${^{208}}Pb(e,e')$. 
979: The solid line is the LEMA$^{\prime}$ result and the dotted line 
980: shows the longitudinal contribution. The data are from Ref.[8]}
981: \label{clesa}
982: \end{figure}
983: 
984: We selected all of the Saclay data with energy transfer between 100 
985: MeV and 200 MeV which had a clearly defined quasielastic peak in the 
986: cross section(218 data points), and used our model with 
987: LEMA$^{\prime}$ to calculate the longitudinal and transverse 
988: contributions to the cross section for each kinematical point in 
989: the data set.
990: We performed a linear least squares fit by scale factors in front of 
991: the longitudinal and transverse contributions.
992: This fit produced a factor in front of the longitudinal term of 0.69 
993: and in front of the transverse term of 1.25.
994: However, the fit is not very good since the ${\chi}^{2}$ per data 
995: point is 60.  The only conclusion we can make is that we do not find 
996: a $50\%$ suppression of the longitudinal contribution in $^{208}Pb$.
997: Furthermore, it appears to us that the experimental data do not 
998: scale quite correctly.
999: In Fig.~(\ref{fitcros}) we show four sets of experimental data along 
1000: with our model.
1001: Two of the sets are at forward angles with quite similar kinematics 
1002: and two are at backward angles with similar kinematics.
1003: One of the forward angle data sets falls significantly below our 
1004: model while the other is above.
1005: It would be difficult to modify the longitudinal and transverse 
1006: strength in our model to get such a dramatic shift in behavior.
1007: For the larger angle case, the contribution of the longitudinal 
1008: terms are negligible and again the kinematics do not vary so much 
1009: between the two cases.
1010: The 310 MeV result lies far above our calculation while the 
1011: 262 MeV result is only slightly above our calculation. 
1012: \begin{figure}[p]
1013: \newbox\figha
1014: \setbox\figha=\hbox{
1015: \epsfysize=130mm
1016: \epsfxsize=130mm
1017: \epsffile{fitcros.eps}}
1018: \noindent\hspace{20mm}\vspace{10mm}\box\figha
1019: \caption{The inconsistent comparison with LEMA$^{\prime}$ and
1020: Saclay data.}
1021: \label{fitcros}
1022: \end{figure}
1023: 
1024: \section{Conclusion}
1025: 
1026: We have developed a simple way of including Coulomb distortion 
1027: in the M\"oller potential for inelastic electron scattering 
1028: from medium and heavy nuclei.
1029: In this paper we have applied a simplified version of this 
1030: approximation to   $(e,e')$ reactions from medium and heavy nuclei 
1031: in the quasielastic region.   
1032: The previously used Effective Momentum Approximation (EMA) disagrees 
1033: with the distorted wave analysis  for nuclei as heavy as $^{208}Pb$, 
1034: while the Local Effective Momentum Approximation with an enhancement factor and an $ad-hoc$ 
1035: correction to the longitudinal term (LEMA$^{\prime}$) reproduces the 
1036: full DWBA calculation very well.
1037: Our most important finding is that LEMA$^{\prime}$ allows the cross 
1038: section to be separated into a longitudinal and a transverse 
1039: contribution.
1040: However,  the resulting structure functions depend on $r$-dependent 
1041: Fourier transforms of the transition current components.
1042: We recommend several different procedures for using LEMA$^{\prime}$ 
1043: for the analysis of experimental data including one where distortion 
1044: effects can be applied to the experimental data and ``plane-wave'' 
1045: structure functions can then be extracted from the corrected data.
1046:  
1047: We analysed the quasielastic $(e,e')$  data from Saclay  on 
1048: $^{208}Pb$ using a relativistic single particle model and 
1049: LEMA$^{\prime}$.
1050: We do not find agreement with the data and even if we vary the 
1051: longitudinal and transverse contributions in our model do not agree 
1052: with the data.
1053: This is to be contrasted with our excellent agreement~\cite{yates} 
1054: with quasielastic data on $^{40}Ca$ and rough agreement with 
1055: quasielastic data~\cite{blat86} on $^{238}U$.
1056: We strongly recommend  that additional  $(e,e')$ experiments on 
1057: medium and heavy nuclei in the quasielastic region be carried out.
1058: The treatment of Coulomb corrections is no longer a serious 
1059: hindrance to the analysis of such experiments.
1060: 
1061: \begin{center}
1062: ACKNOWLEDGMENTS
1063: \end{center}
1064: 
1065: We thank the Ohio Supercomputer Center in Columbus for many hours
1066: of Cray Y-MP time to develop this calculation and to perform the
1067: necessary calculations. This work was supported in part by the U.S.
1068: Department of Energy under Grant No. FG02-87ER40370.
1069: 
1070: \newpage
1071: \begin{references}
1072: \bibitem{jin94}Yanhe Jin, D. S. Onley, and L. E. Wright, Phys. 
1073: Rev. C{\bf 50}, 168(1994).
1074: \bibitem{mez85}Z. E. Meziani, $et$ $al.$, Nucl. Phys. A{\bf446}, 
1075: 113(1985); Phys. Rev. Lett., {\bf 52}, 2130(1984); 
1076: {\bf 54}, 1233(1985).
1077: \bibitem{dead86}M. Deady, $et$ $al.$, Phys. Rev. C{\bf 33}, 
1078: 1897(1986); {\bf 28}, 631(1983).
1079: \bibitem{blat86}C. C. Blatchley, J. J. LeRose, O. E. Pruet, 
1080: P. D. Zimmerman, C. F. Williamson, and M. Deady, Phys. Rev. 
1081: C{\bf 34},1243(1986).
1082: \bibitem{jin92}Yanhe Jin, D. S. Onley, and L. E. Wright, Phys. Rev. 
1083: C{\bf 45}, 1333(1992).
1084: \bibitem{yates} T.C. Yates,  C.F. Williamson, W.M. Schmitt, 
1085: M. Osborn, M. Deady, Peter D. Zimmerman, C. C. Blatchley, 
1086: Kamal K. Seth,  M. Sarmiento, B. Parker, Yanhe Jin, L.E. Wright, 
1087: and D.S. Onley,  Phys. Lett. B{\bf 312}, 382(1993).
1088: \bibitem{jourdan}J. Jourdan, Phys. Lett. B{\bf 353}, 189(1996).
1089: \bibitem{saclay}A. Zghiche, $et$ $al.$, Nucl. Phys. A{\bf 573}, 
1090: 513(1994).
1091: \bibitem{jinphd}Yanhe Jin, D. S. Onley, and L. E. Wright, Phys. Rev. 
1092: C{\bf 45}, 1311(1992); Yanhe Jin, Ph.D., Dissertation, 
1093: Ohio University, 1991.
1094: \bibitem{knol}J. Knoll, Nucl. Phys. A{\bf 201}, 289(1973); 
1095: A{\bf 223}, 462(1974).
1096: \bibitem{uber}H. $\ddot{U}$berall, {\it Electron Scattering from  
1097: Complex Nuclei} (Academi, Press. New York, 1971).
1098: \bibitem{lenz}F. Lenz, thesis, Univ. of Freiburg, (1971); 
1099: F. Lenz and R. Rosenfelder, Nucl. Phys. A{\bf 176}, 513(1971).
1100: \bibitem{hol}G. H$\ddot{o}$hler, E. Pietarinen, 
1101: I. Sabba-Stevanescu, F. Borkowski, G.G. Simon, V. H. Walter, 
1102: and R. D. Wendling, Nucl. Phys. B{\bf 114}, 505(1976).
1103: \bibitem{trai}M. Traini, S. Turck-Chi$\acute{e}$ze, and 
1104: A. Zghiche, Phys. Rev. C{\bf 38}, 2799(1988); 
1105: Phys. Lett. B{\bf 213}, 1(1988).
1106: \bibitem{giupac}C. Giusti and F. D. Pacati, Nucl. Phys. 
1107: A{\bf 473}, 717(1987).
1108: \bibitem{kimphd}Kyungsik Kim, Ph.D., Dissertation, 
1109: Ohio University, 1996.
1110: \end{references}
1111: 
1112: \end{document}
1113: