1: %\documentclass{elsart}
2: \documentclass[showpacs]{revtex4}
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5: \usepackage{epsfig}
6:
7: \newcommand{\be}{\begin{eqnarray}}
8: \newcommand{\ee}{\end{eqnarray}}
9: \newcommand{\dis}{\displaystyle}
10: \def\bra#1{\textstyle{\left\langle \, #1 \, \right\vert \>}}
11: \def\ket#1{\textstyle{\> \left\vert \>\! #1 \>\! \right\rangle}}
12:
13: \begin{document}
14: %\title{\hfill {\footnotesize FZJ--IKP(TH)--2005--21} \\
15: \title{
16: Extraction of scattering lengths from
17: final-state interactions}
18:
19: \author{A. Gasparyan$^1$, J. Haidenbauer$^2$, and C. Hanhart$^2$}
20:
21: \affiliation{
22: $^1$Institute of Theoretical and Experimental Physics,
23: 117259, B. Cheremushkinskaya 25, Moscow, Russia\\
24: $^2$Institut f\"ur Kernphysik (Theorie), Forschungszentrum J\"ulich,
25: D-52425 J\"ulich, Germany
26: }
27:
28: \begin{abstract}
29: A recently proposed method based on dispersion theory, that allows to extract
30: the scattering length of a hadronic two-body system from corresponding
31: final-state interactions, is generalized
32: to the situation where the Coulomb interaction is present.
33: The steps required in a concrete practical application are discussed in detail.
34: %
35: In addition a thorough examination of the accuracy of the proposed method
36: is presented and a comparison is made with results achieved with other
37: methods like the Jost-function approach based on the effective-range
38: approximation. Deficiencies of the latter method are pointed out.
39: %
40: The reliability of the dispersion theory method for extracting also the
41: effective range is investigated.
42: \end{abstract}
43: %
44: \pacs{11.55.Fv,13.75.-n,13.75.Ev,25.40.-h}
45:
46: \maketitle
47:
48: \section{Introduction}
49:
50: The scattering length
51: provides not only an important measure for the strength of the
52: interaction in a specific hadronic two-body system \cite{Joach}
53: but often allows to draw further more general and thus even more
54: interesting conclusions. For example in the case of the
55: proton-proton and neutron-neutron systems the corresponding
56: scattering lengths in the $^1S_0$ partial wave provide a
57: very sensitive test of charge symmetry in the strong interaction \cite{Miller}.
58: SU(3) symmetry can be tested by comparing the $^1S_0$ scattering
59: lengths in the neutron-proton and $\Sigma^+p$ systems, which should
60: fulfil the relation $a_{np} = a_{\Sigma^+p}$ in case SU(3) symmetry
61: holds rigorously \cite{Dover}.
62: In the chiral limit the $\pi N$ S-wave scattering
63: lengths vanish and therefore any deviation from that value
64: is a direct measure for how strongly this symmetry is broken. Here especially
65: the isoscalar component is of high interest due to its close link to the sigma
66: term of the nucleon \cite{xx}.
67: On a more phenomenological level the $\eta{N}$ scattering length is
68: interesting because the magnitude of its real part is directly
69: linked with speculations about the existence of $\eta$-mesonic
70: hadronic bound states such as $\eta ^3He$
71: \cite{Wycech,Belyaev1,Belyaev2,Rakityansky,Fix1,Niskanen,Niskanen1}
72: or $\eta ^4He$ \cite{Haider2}.
73:
74: Unfortunately, a direct determination of the scattering length is
75: only feasible in a few cases. It can be done with
76: scattering experiments sufficiently close to the reaction threshold
77: so that the effective range expansion can be utilized for
78: extracting the scattering length. But in practice such experiments are
79: only possible for charged and also (quasi) stable particles,
80: as it is the case with, e.g., $pp$, $\pi^+p$ or $K^+p$ scattering.
81: One can also extract the scattering length from a study of the hadronic
82: level shifts of atoms \cite{rus} like $\pi^-p$,
83: $\pi^-d$ or also $\bar p p$ \cite{Gotta}.
84: For the majority of the hadronic two-body systems
85: information about the scattering length is only accessible
86: via an investigation of the final-state interaction of systems
87: which have at least three particles in the final state.
88:
89: While in the former type of experiments the accuracy of the scattering
90: length is directly connected with the precision of
91: the data more detailed and often sophisticated considerations are necessary
92: in order to estimate additional uncertainies that arise when the scattering
93: length is extracted from final-state effects \cite{Gibbs}. However, in some
94: cases such an
95: estimation is facilitated by the fact that the reaction mechanism is known.
96: For example, the reaction $nd \to (nn)p$, one of the prime sources of the
97: $nn$ scattering length, can be analysed by means of rigorous Faddeev
98: calculations \cite{Trotter,Huhn,Deng}. Reactions involving the pion such as
99: $\pi d$, $\pi^- d\to \gamma nn$ or $\gamma d \to \pi^+ nn$, which can be used to extract the $\pi N$ and
100: $nn$ scattering lengths, respectively, can be tackled by chiral perturbation
101: theory in a well controlled way in the relevant near-threshold regime \cite{xxx,Lensky}.
102:
103: In a recent publication \cite{Gasparyan2003}
104: we argued that also large-momentum transfer reactions such as
105: $pp\to K^+p\Lambda$ \cite{C11_1,Bilger,jan} or $\gamma d\to K^+n\Lambda$
106: \cite{Renard,Kerbikov,Mecking,Adel,Li,Yama} are excellent candidates for extracting
107: information about the scattering lengths. In reactions with large
108: momentum transfer the production process is necessarily of short-ranged
109: nature. As a consequence the results are basically insensitive to details of
110: the production mechanism and therefore a reliable and general error estimation can
111: be given.
112: %
113: Indeed, in \cite{Gasparyan2003} a formalism based on dispersion theory was
114: presented that relates spectra from large-momentum transfer reactions,
115: such as $pp\to K^+p\Lambda$ or $\gamma d\to K^+n\Lambda$, directly to
116: the scattering length of the interaction of the final state particles.
117: The theoretical error of the method was estimated to be 0.3 fm or
118: even less, which is comparable to the error quoted in the context of the
119: determination of $a_{nn}$ \cite{Howell}, say.
120: This estimate was confirmed by comparing results obtained with
121: the proposed formalism to those of microscopic model calculations for the
122: specific reaction $pp\to K^+\Lambda p$.
123: %
124: The arguments but also the formalism of Ref. \cite{Gasparyan2003} are, of course,
125: valid for any production or decay process which is of short-ranged nature,
126: i.e. also for investigation of hadronic two-particle subsystems resulting from the
127: decay of the $J/\Psi$ or $B$ mesons \cite{Psi,Belle}.
128:
129: In the present paper we want to investigate further aspects of extracting scattering
130: lengths from final-state interactions which were not addressed in our
131: earlier work. One of those topics is the presence of the Coulomb interaction.
132: In many interesting hadronic two-particle systems both particles carry
133: charges, like in the already mentioned $\Sigma^+p$ channel whose scattering
134: length could be extracted from the reaction $pp\to K^0 \Sigma^+p$.
135: Then the production amplitude acquires additional singularities, due to the
136: long-range nature of the Coulomb forces, and the formalism developed
137: in Ref. \cite{Gasparyan2003} is no longer directly applicable.
138: We will derive the modifications that are necessary in order to adapt the
139: dispersion-relation method to the situation when the Coulomb force
140: is present in the final-state interaction. We also demonstrate in a toy model
141: calculation how one has to proceed in a concrete application to data.
142:
143: In addition we present a more detailed examination of the accuracy of the
144: method proposed in Ref. \cite{Gasparyan2003}.
145: A test based on one specific model calculation, namely for the
146: reaction $pp\to K^+\Lambda p$, has been already performed in that paper.
147: However, here we want to put this investigation on a broader basis by considering
148: final-state interactions of varying strengths, corresponding to a much larger range
149: of values of the scattering length. In addition we take a look at the effective
150: range $r_e$ as well which can be also extracted by the proposed dispersion-integral
151: method. For the effective range a sensible error estimation is not possible, as was
152: already pointed out in Ref. \cite{Gasparyan2003}, but it is still interesting to
153: examine in concrete applications whether meaningful results could be achieved.
154: Finally, and equally important, we want to
155: compare the present method with the performance of other, approximative treatments
156: of the final-state interaction that are commonly used in the literature to
157: extract information on the scattering length and also the effective range.
158: This concerns in particular the Jost-function approach \cite{book} based on the
159: effective-range approximation
160: and an even simpler approach that relies simply on utilizing the effective range
161: approximation itself \cite{migdal}. Thereby, we will show that the latter methods lead to
162: (partly drastic) systematic deviations
163: from the true values and therefore one has to be rather cautious in the
164: interpretation of results achieved with those methods.
165:
166: The paper is structured in the following way: In the subsequent section we
167: give a short review of the dispersion integral method for extracting the scattering
168: length from final-state interactions. In section 3 results of an examination of the
169: accuracy of this method are presented. Thereby,
170: we consider various (singlet and triplet) $S$ wave $YN$ and $NN$ interactions
171: and compare the scattering lengths and effective range extracted with the
172: dispersion-integral method from appropriately generated final-state effects
173: with the ones predicted by the models. We also apply two approximative methods
174: for treating final-state effects, namely the Jost-function approach
175: based on the effective range approximation (Jost-ERA)
176: as well as the effective range approximation itself,
177: and compare their performance with the one of our method.
178: In section 4 we generalize the dispersion-integral method to the case where
179: a repulsive Coulomb interaction is present in the final state. Test calculations for
180: a final-state interaction with Coulomb are then presented in sect. 5 and it
181: is discussed in detail how one has to procede in a practical application.
182: The paper ends with a short summary.
183:
184:
185:
186: \section{Formalism}
187: Our method, which goes back to an idea of Geshkenbein
188: \cite{Geshkenbein1969,Geshkenbein1998},
189: is based on using the dispersion relation technique.
190: Consider the production amplitude $A_S$ of a $2\to 3$ reaction. To be
191: concrete we discuss $pp\to
192: K^+p\Lambda$, or $\gamma d\to K^+n\Lambda$, with the
193: $\Lambda N$ system being in an $L=0$ partial wave and a specific spin state
194: $S$ ($^1S_0$ or $^3S_1$). This amplitude depends on the total energy
195: squared $s=(p_1+p_2)^2$, the invariant mass squared of the
196: outgoing $\Lambda N$ system $m^2=(p_N+p_\Lambda)^2$ and the
197: momentum transfer $t=(p_1-p_{K^+})^2$, where $p_1$, $p_2$, $p_N$, $p_\Lambda$,
198: and $p_{K^+}$ are the 4-momenta of the two initial particles, final nucleon,
199: lambda, and kaon, respectively. Then one can write down a dispersion relation
200: for this amplitude with respect to $m^2$ at fixed $s$ and $t$
201: \begin{eqnarray}
202: A_S(s,t,m^2)=\frac1\pi\int_{-\infty}^{\tilde m\, ^2} \frac{D_S(s,t,m' \, ^2)}{m' \, ^2-m^2}dm' \, ^2
203: +\frac1\pi\int_{m_0^2}^\infty \frac{D_S(s,t,m' \, ^2)}{m' \, ^2-m^2}dm' \, ^2,
204: \label{dispers}
205: \end{eqnarray}
206: where $\tilde m\, ^2$ is the upper boundary of the lefthand cut, $m_0^2 = (m_N+m_\Lambda)^2$, and
207: \begin{eqnarray}
208: D_S(s,t,m^2) = \frac{1}{2i}(A_S(s,t,m^2+i0)-A_S(s,t,m^2-i0))
209: \end{eqnarray}
210: is the discontinuity of the amplitude along the cuts.
211: We neglect here the contributions from possible kaon-baryon interactions.
212: In case they are not small, they still can be considered as constant (weakly
213: mass dependent) if one chooses the kinematics such that the excess energy of
214: the reaction is significantly larger than the typical range of the $\Lambda N$
215: interaction, cf. the discussion in Ref. \cite{Gasparyan2003}.
216: The index $S$ denoting the spin state will be suppressed in the following to
217: simplify the notation.
218:
219: For a purely elastic $\Lambda N$ system, the discontinuity along the righthand
220: cut would be given by
221:
222: \begin{eqnarray}
223: D(s,t,m^2) = A(s,t,m^2)e^{-i\delta}\sin{\delta},
224: \label{drhc}
225: \end{eqnarray}
226: where $\delta$ is the $\Lambda N$ ($^1S_0$ or $^3S_1$) scattering phase
227: shift.
228: Then the solution of Eq. \eqref{dispers} in the physical region reads
229: (see Refs. \cite{Muskhelishvili1953,Omnes1958,Frazer1959})
230: \begin{eqnarray}
231: A(s,t,m^2)=\exp\left[{\frac1\pi\int_{m_0^2}^\infty\frac{\delta(m' \, ^2)}{m' \, ^2-m^2-i0}dm' \, ^2}\right]
232: \Phi(s,t,m^2),
233: \label{dis0}
234: \end{eqnarray}
235: where
236: $\Phi(s,t,m^2)$ contains only lefthand singularities and therefore is a
237: slowly varying function of $m^2$. In order to ensure this requirement to be
238: fulfilled it is important that the momentum transfer $t$ is large.
239: We assume also that there is no bound state in the $\Lambda N$ system.
240:
241: Consider now a realistic situation where inelastic channels are present -- as it is the
242: case with $\Lambda N$ due to the coupling to the $\Sigma N$ channel, say.
243: Then one can write down a formula similar to Eq.~\eqref{dispers}, but with
244: the integration performed over a finite range of masses \cite{Gasparyan2003}:
245: \begin{eqnarray}
246: A(m^2)=\exp\left[{\frac1\pi\int_{m_0^2}^{m_{max}^2}\frac{\delta(m' \, ^2)}{m' \, ^2-m^2-i0}dm' \, ^2}\right]
247: \tilde \Phi(m^2) \label{C_int} \ ,
248: \label{A_def}
249: \end{eqnarray}
250: where $\tilde \Phi(m^2)$ is again a slowly varying function of $m^2$ given
251: the phase shift $\delta$ is sufficiently small in the vicinity of $m_{max}$ \cite{Gasparyan2003}.
252: The upper limit $m_{max}$ has to be chosen in such a way that the corresponding relative
253: momentum of the $\Lambda N$ system, $p_{max}$, is of the order of the typical
254: scale of the $\Lambda N$ interaction, i.e. in the order of $1/a$ or $1/r$.
255: The equation \eqref{A_def} can be solved with respect to the $\delta$ \cite{Gasparyan2003}:
256: \begin{eqnarray}
257: \nonumber
258: &&\frac{\delta(m^2)}{\sqrt{m^2-m_0^2}}= \\
259: & & -\frac1{2\pi}{\bf P}
260: \int_{m_0^2}^{m_{max}^2}\frac{\log{|A(m' \, ^2)/\tilde \Phi(m_{max}^2,m' \, ^2)|^2}}
261: {\sqrt{m' \, ^2-m_0^2} \ (m' \, ^2-m^2)}\sqrt{\frac{m_{max}^2-m^2}{m_{max}^2-m' \, ^2}}dm' \, ^2.
262: \label{almostfinal}
263: \end{eqnarray}
264: If one neglects the mass dependence of $\tilde \Phi(m^2)$ and uses the
265: relation between the partial cross section $\sigma_S$ and the amplitude
266: $$\frac{d^2\sigma_S}{dm' \, ^2dt} \propto p'|A_S(s,t,m'\,^2)|^2 \ ,$$
267: then one obtains the expression for the scattering length in terms
268: of observables
269: \begin{eqnarray}
270: \nonumber
271: a_S&=&\lim_{{m}^2\to m_0^2}\frac1{2\pi}\left(\frac{m_\Lambda+m_N}
272: {\sqrt{m_\Lambda m_N}}\right){\bf P}
273: \int_{m_0^2}^{m_{max}^2}dm' \, ^2
274: \sqrt{\frac{m_{max}^2-{m}^2}{m_{max}^2-m' \, ^2}}\\
275: & & \qquad \qquad \times \
276: \frac1{\sqrt{m' \, ^2-m_0^2} \ (m' \, ^2-{m}^2)}
277: \log{\left\{\frac{1}{p'}\left(\frac{d^2\sigma_S}{dm' \, ^2dt}\right)\right\}}
278: \ ,
279: \label{final}
280: \end{eqnarray}
281: and analogously for the effective range $r_e$.
282:
283: \section{Accuracy of the method and comparison with other approaches}
284:
285: The most important advantage of the method proposed by us \cite{Gasparyan2003}
286: is that a reliable estimate for the uncertainty of the extracted scattering
287: length can be given. The sources for the uncertainty are trifold: (i) A possible
288: influence of the final-state interaction in the other outgoing channels. For the
289: reaction $pp\to K^+\Lambda p$ considered in Ref.~\cite{Gasparyan2003} this
290: concerns the $K\Lambda$ and $KN$ systems. (ii) The adopted value for $m^2_{max}$, the
291: upper limit chosen for the dispersion integral in Eq. (\ref{final}).
292: (iii) A sensitivity to left-hand cuts of the production operation.
293: %
294: A detailed analysis of the issues (ii) and (iii) , based on general arguments,
295: presented in Ref.~\cite{Gasparyan2003} suggests that the error in the scattering
296: length should be typically in the order of 0.3 fm or less. The role of issue (i)
297: cannot be quanitified theoretically but has to be investigated by performing experiments
298: and corresponding analyses at different beam momenta \cite{Gasparyan2003}.
299:
300: In this section we want to present a thorough examination of the accuracy of the proposed
301: method and, in particular, to corroborate the error estimate, by means of concrete
302: model calculations. A test based on one specific model calculation, namely for the
303: reaction $pp\to K^+\Lambda p$, has been already performed in Ref.~\cite{Gasparyan2003}.
304: However, here we want to put this investigation on a broader basis by considering
305: final-state interactions of varying strengths, corresponding to a much larger range
306: of values of the scattering length. Furthermore, and equally important, we want to
307: compare the present method with the performance of other, approximative treatments
308: of the final-state interaction that are commonly used in the literature to
309: extract information on the scattering length and the effective range
310: and have been applied to $pp\to K^+\Lambda p$ \cite{Bale,Hinter}.
311:
312: %%%
313: One of those approximative treatments follows from the assumption that the
314: phase shifts are given by the first two terms in the effective range
315: expansion,
316: \begin{eqnarray}
317: p \ {\rm cot} (\delta(m^2)) = -{1 \over a}+{r_e\over 2} p \,^2 \ ,
318: \label{ere}
319: \end{eqnarray}
320: usually called the effective range approximation (ERA), over the whole energy
321: range. Here $p$ is the
322: relative momentum of the final-state particles under consideration
323: in their center of mass system, corresponding to the invariant mass $m^2$.
324: In this case the relevant integrals \eqref{dis0} can be evaluated in closed
325: form as \cite{book}
326: \begin{eqnarray}
327: A(m^2) \propto
328: \frac{(p^2+\alpha^2)r_e/2}{-1/a+(r_e/2)p^2-ip} \ ,
329: \label{arform}
330: \end{eqnarray}
331: where $\alpha = 1/r_e(1+\sqrt{1-2r_e/a})$.
332: Because of its simplicity Eq. (\ref{arform}) is often used for the treatment
333: of the final-state interaction (FSI).
334:
335: A further simplification can be made if one assumes that
336: $a\gg r_e$. This situation is practically realized in the $^1S_0$ partial
337: wave of the $NN$ system. Then
338: the energy dependence of the quantity in Eq. (\ref{arform})
339: is given by the energy dependence of the elastic amplitude
340: \begin{eqnarray}
341: A(m^2) \propto
342: \frac{1}{-1/a+(r_e/2)p^2-ip} \ ,
343: \label{MW}
344: \end{eqnarray}
345: as long as $p \ll 1/r_e$.
346: Therefore one expects that, at least for
347: small kinetic energies, $NN$ elastic scattering and meson production
348: in $NN$ collisions with a $NN$ final state exhibit the same energy dependence \cite{book,migdal},
349: which indeed was experimentally confirmed.
350: This treatment of FSI effects is often referred to as Migdal-Watson (MW) approach \cite{migdal}.
351:
352: In order to examine the reliability of the three methods described above we took different
353: $YN$ models from the literature \cite{Holz,Melni2,NijV,Haiden} and calculated the
354: production amplitude $A(m^2)$ utilizing the meson exchange model from Ref. \cite{model}.
355: Then this amplitude was used for
356: extracting the scattering length by means of the dispersion integral
357: Eq. (\ref{final}) or from the approximative prescriptions given by
358: Eqs. (\ref{arform}) and (\ref{MW}). For comparison we considered also
359: the $^1S_0$ partial wave of the $np$ system of the Argonne potential \cite{Argonne}.
360: In this case $A(m^2)$ was set equal to the scattering wave function $\Psi (p,r)$
361: at the origin, more precisely to $\Psi^-(p,0)^*$, which corresponds to the assumption
362: that the production operator is point-like.
363:
364: \begin{table}[h]
365: \caption{$S$-wave scattering lengths $a$ (in fm) for various $YN$ \cite{NijV,Melni2}
366: and $NN$ \cite{Argonne} potentials.
367: The results for the original models are compared with those obtained by
368: applying the dispersion integral method (\ref{final})
369: and the approximations Eq. (\ref{arform}) (Jost-ERA) and Eq. (\ref{MW}) (MW).
370: }
371: \vskip 0.2cm
372: \begin{tabular}[t]{|c|c|c|c|c|}
373: \hline
374: model& \ exact result \ & \ disp.int. \ & \ Jost-ERA \ & \ MW \ \\
375: \hline
376: \hline
377: J\"ulich 01 singlet& -1.02&-1.03&-1.28& -1.67\\
378: \hline
379: Nijmegen 97a singlet& -0.73&-0.75&-0.98&-1.33 \\
380: \hline
381: Nijmegen 97f singlet& -2.59&-2.57&-2.96& -3.35\\
382: \hline
383: J\"ulich 01 triplet& -1.89&-1.66&-2.05& -2.42\\
384: \hline
385: Nijmegen 97a triplet& -2.13&-1.98&-2.37& -2.75\\
386: \hline
387: Nijmegen 97f triplet& -1.69&-1.61&-2.00& -2.37\\
388: \hline
389: Argonne v14 singlet & -23.71&-23.54&-24.56& -24.79\\
390: \hline
391: \end{tabular}
392: \label{Table1}
393: \end{table}
394:
395: Some selective results
396: (for the Nijmegen NSC97 \cite{NijV} and J\"ulich 01 \cite{Melni2} $YN$ models
397: and the Argonne v14 \cite{Argonne} $NN$ potential) are summarized in Table \ref{Table1}.
398: The second column contains the
399: correct scattering length evaluated directly from the potential model.
400: One can see that the extraction of the scattering length via the
401: dispersion integral (\ref{final}) yields results pretty close to the
402: original values for all considered potentials. In fact, in most cases
403: the deviation is significantly smaller than the uncertainty of the method,
404: estimated in Ref. \cite{Gasparyan2003} to be 0.3 fm. The results of the
405: Jost-ERA approach Eq. (\ref{arform})
406: exhibit a systematic offset in the order of 0.3 fm. The situation is much
407: worse for the Migdal-Watson approach Eq. (\ref{MW})
408: where a similar offset is found though
409: now in the order of 0.6 fm. As a consequence, the extracted values
410: differ by 50 \% or more from the correct scattering lengths. Only for
411: the $^1S_0$ $np$ partial wave the disagreement is still in the order of 5 \%.
412: Here the reliability of the Jost-ERA and Migdal-Watson approaches are
413: comparable. This is in agreement with the expectations mentioned above.
414:
415: The systematic offset inherent in the Jost-ERA approach as well as in
416: the Migdal-Watson prescription can be best seen in Fig.~\ref{fig0}, where we
417: shown the difference between the scattering lengths predicted by various
418: models and the values
419: extracted via the dispersion integral (circles), the Jost-ERA method (squares)
420: and the Migdal-Watson prescription (triangles).
421:
422: While the Jost-ERA approach might still be a reasonable tool for getting a
423: first rough estimate of the scattering length for a particular two-body interaction
424: one should be rather cautious when using it for more quantitative analyses.
425: In particular, its application in a combined fit to elastic scattering data and
426: invariant mass spectra, e.g. to $\Lambda p$ and $pp\to K^+\Lambda p$,
427: is rather problematic and can easily cause misleading results. Because of the
428: offset in the scattering length in applications to final-state effects it is clear
429: that a combined fit cannot converge to a unique (the ``true'') $\Lambda p$
430: scattering length. Only the elastic data will favour values close to the
431: ``true'' scattering length whereas the production data tend to support larger
432: (negative) values. This is obvious from the corresponding Jost-ERA
433: results presented in Table \ref{Table1} and also from Fig.~\ref{fig0}.
434: %
435: We believe that the analysis of Hinterberger and Sibirtsev presented
436: in Ref. \cite{Hinter} is an instructive exemplification of this dilemma.
437: Employing the
438: Jost-ERA approach to low energy total $\Lambda p$ cross sections
439: \cite{Alex,Sechi} and to experimental results for the missing mass
440: spectrum of the reaction $pp\to K^+X$ \cite{Siebert} separately, they derived
441: (spin averaged)
442: scattering lengths of $a=-1.81^{+0.18}_{-0.21}$ fm and $a=-2.57^{+0.20}_{-0.23}$ fm,
443: respectively. Taking into account the error bars this is roughly the difference we
444: would expect from the offset (of around $0.3$ fm) seen in our test calculations
445: and, therefore, one must consider the results as being practically consistent
446: with each other.
447: %
448: But the authors of Ref. \cite{Hinter} attempted to ``reconcile'' the results
449: even more by introducing a spin-dependence in the fitting procedure.
450: Indeed,
451: with the relative magnitude of singlet to triplet contribution in the production
452: reaction as free parameter (their relative strength in the elastic channel is
453: fixed at 1:3 by the spin weight!) a ``fully consistent'' description of the
454: combined data could be achieved \cite{Hinter} and apparently the spin-singlet as
455: well as spin-triplet $\Lambda p$ $S$-wave scattering lengths could be
456: determined from spin-averaged observables.
457: Our experience with the Jost-ERA approach reported above, however,
458: strongly suggests that the sensitivity to the spin seen in this analysis is most
459: likely just an artifact of the method applied.
460:
461: \begin{table}[h]
462: \caption{$S$-wave effective ranges $r_e$ (in fm) for various $YN$ \cite{NijV,Melni2}
463: and $NN$ \cite{Argonne} potentials.
464: The results for the original models are compared with those obtained by
465: applying the dispersion integral method (\ref{final}) and the approximation
466: Eq. (\ref{arform}) (Jost-ERA).
467: }
468: \vskip 0.2cm
469: \begin{tabular}[t]{|c|c|c|c|}
470: \hline
471: model& \ exact result \ & \ disp.int. \ & \ Jost-ERA \ \\
472: \hline
473: \hline
474: J\"ulich 01 singlet& 4.49&4.31 &2.48\\
475: \hline
476: Nijmegen 97a singlet& 6.01&4.78 &2.81\\
477: \hline
478: Nijmegen 97f singlet& 3.05&2.82 &1.60\\
479: \hline
480: J\"ulich 01 triplet& 2.57& 2.49&1.89\\
481: \hline
482: Nijmegen 97a triplet& 2.74&2.60 &1.70\\
483: \hline
484: Nijmegen 97f triplet& 3.34& 2.67&1.69\\
485: \hline
486: Argonne v14 singlet& 2.78 & 2.91&0.43 \\
487: \hline
488: \end{tabular}
489: \label{Table2}
490: \end{table}
491:
492: Let us now come to the effective range $r_e$.
493: Since the dispersion relations yield only an integral representation for the product
494: $a^2((2/3)a-r_e)$ but not for the effective range $r_e$ alone
495: \cite{Gasparyan2003} it follows that the attainable accuracy of $r_e$ is always
496: limited roughly by twice the relative error on $a$.
497: Still, it is interesting to see what values one gets for $r_e$ from the dispersion integrals.
498: Corresponding results are presented in Table \ref{Table2} and in Fig. \ref{fig0a}.
499: Evidently, the values extracted via the dispersion integral agree much better with the
500: original results as one might have expected. In fact, in practically all cases the
501: deviation is in the order of only 5 \% or even less. This suggests that one could use
502: the dispersion integrals also to extract the effective range $r_e$ from data. But
503: one should keep in mind that, unlike the case of the scattering length, now one cannot
504: rely on a solid and general estimate of the uncertainty.
505: As far as the Jost-ERA approach is concerned it is clear from Table \ref{Table2} that
506: it yields rather poor results. In case of the Migdal-Watson prescription (\ref{MW})
507: it turned out that the fit always prefers an effective range $r_e$ equal to
508: zero. This is due to the term proportional to $r^2_ep^4$ in the denominator of the
509: $A(m^2)$ that make the production cross section decrease too fast as compared
510: to the data (or to our calculations with realistic models).
511: Therefore we don't show any results of the Migdal-Watson fit for $r_e$.
512:
513: One should note here that the upper limit in the dispersion integrals was
514: always taken such that $p_{max}=205$ MeV/c (as in
515: \cite{Gasparyan2003}) that corresponds to $\epsilon_{max}\equiv
516: m_{max}-m_0\approx 40$ MeV for the $\Lambda N$ and $\Sigma N$ systems and
517: $\epsilon_{max}\approx 45$ MeV for the $NN$ system. In the latter case it
518: is interesting to see what happens if one varies the range of integration,
519: since the energy structure in the $NN$ interaction is much narrower due to
520: the large $NN$ scattering length. For $\epsilon_{max}=10$ MeV and
521: $\epsilon_{max}=20$ MeV as upper limits of the integration one gets the
522: scattering lengths $a=-22.62$ fm and $a=-23.17$ fm, respectively -- which
523: are in principle still close to the original value. For the effective range,
524: however, the calculation yields $r_e=4.78$ fm and $3.72$ fm, respectively.
525: The reason why the agreement for $\epsilon_{max}>40$ MeV is so good
526: is that the $NN$ $^1S_0$ phase shift becomes sufficiently small at such
527: energies, which implies a small uncertainty according to the error estimation
528: in Ref.~\cite{Gasparyan2003}.
529:
530:
531: \section{Dispersion relation in the presence of Coulomb repulsion}
532:
533: In the case when both baryons in the final state carry charges
534: (for example in the reaction $pp\to K^0 p\Sigma^+$) there
535: is a Coulomb interaction between them. Then the production amplitude
536: $A(m^2)$ acquires additional singularities at $p=0$, due to the
537: long-range nature of the Coulomb forces, and the formalism developed
538: in Sect. 2 is no longer applicable directly. In this section we want
539: to describe the modifications that are necessary in order to adapt the
540: dispersion-relation method to the situation when the Coulomb force
541: is present in the final-state interaction. We restrict ourselves to the
542: case of a repulsive Coulomb interaction so that no bound states are
543: present.
544:
545: In order to elucidate the principle idea we start
546: out from the case of elastic (two-body) scattering. Here the problem
547: can be most conveniently dealt with by applying the Gell-Mann--Goldberger
548: two-potential formalism \cite{Gell}. Let us assume that the total potential
549: $V = V_c + V_s$ is given by the sum of a short-ranged hadronic potential
550: $V_s$ and the Coulomb interaction $V_c$. Then the total reaction
551: amplitude $T$ can be written as $T = T_c + T_{cs}$, where $T_c$ is the
552: Coulomb amplitude and $T_{cs}$ is defined by
553:
554: \be
555: T_{cs}=(1+T_cG_0)t_{cs}(1+G_0T_c),
556: \ee
557: where $t_{cs}$ fulfils a Lippmann-Schwinger equation,
558: \be
559: t_{cs}=V_s+V_sG_ct_{cs},
560: \ee
561: with the short-range potential $V_s$ as driving term.
562: %
563: To obtain the physical on-shell amplitudes one needs to project
564: the corresponding $T$-operators on the so-called Coulombian asymptotic
565: states $\ket{p_\infty\pm}$ which are related to the Coulomb scattering states
566: (with fixed angular momentum -- in our case $l=0$) $\ket{p\pm}_c$ via
567: $ \ket{p\pm}_c=\ket{p_\infty\pm} + G_0^{\pm}T_c^{\pm}\ket{p_\infty\pm} $ \cite{vanHaeringen1976}.
568: Here $p$ denotes the center of mass momentum in the baryon-baryon system.
569: In this way one obtains in particular
570: \be
571: _c\bra{p-}t_{cs}\ket{p+}_c=-\frac{1}{\pi\mu}f_{cs}(p),
572: \ee
573: where $f_{cs}$ is the so-called Coulomb-modified nuclear scattering amplitude
574: and $\mu$ is the reduced mass.
575: Its relation to the phase shift $\delta_{cs}$ is the following
576: \be
577: f_{cs}=\frac{e^{2i\delta_c}(e^{2i\delta_{cs}}-1)}{2ip},
578: \ee
579: with $\delta_c$ denoting the pure Coulomb $S$-wave phase shift
580: given by \hbox{$\delta_{c}=\arg{(\Gamma(1+i\eta))}$} with
581: $\eta=\frac{\mu e^2}{p}$.
582:
583: It has been shown in Ref. \cite{Hamilton1973}
584: under rather general assumptions that
585: the modified amplitude $\tilde{f}(p)=e^{-2i\delta_c}f_{cs}(p)/C^2(p)$
586: is free of the Coulomb singularities on the physical sheet
587: and possesses only the singularities caused by dynamical cuts
588: (see also Refs. \cite{Cornille1962,vanHaeringen1977,Heller1966,Scotti1965}).
589: In addition below the inelastic cuts and above the two-baryon threshold
590: the modified unitarity relation reads
591: \be
592: \tilde{f}(s+i0)-\tilde{f}(s-i0)=2ip\tilde{f}(p)\tilde{f}^*(p)C^2(p)
593: \ee
594: with $ C^2(p)=\frac{2\pi\eta}{e^{2\pi\eta}-1}$ being the Coulomb
595: penetration factor.
596:
597: Furthermore an effective range function, modified for the presence of
598: the Coulomb interaction, can be defined as well. Is is given by
599: \be
600: S(p)\equiv pC^2(p)\cot\delta_{cs}(p)+Q(p)=-1/a_{cs}+r_ep^2/2+..,
601: \label{eff_range}
602: \ee
603: where $Q(p)\equiv\mu e^2[\psi(i\eta)+\psi(-i\eta)-2\ln\eta], \ \psi(z)=\Gamma'(z)/\Gamma(z)$.
604:
605: Coming back now to the production reaction it can be analogously
606: shown that also the modified production amplitude
607: \be
608: \tilde{A}(m^2)=e^{-i\delta_c}A(m^2)/C(p)
609: \ee
610: is free of the aforementioned singularities \cite{Hamilton1973}.
611: Therefore a dispersion relation similar to Eq.~\eqref{dispers} can be written down
612: \be
613: \tilde A(m^2)=\frac1\pi\int_{-\infty}^{\tilde m\, ^2} \frac{\tilde D(m' \, ^2)}{m' \, ^2-m^2}dm' \, ^2
614: +\frac1\pi\int_{m_0^2}^\infty \frac{\tilde D(m' \, ^2)}{m' \, ^2-m^2}dm' \, ^2.
615: \label{dispers1}
616: \ee
617: Unitarity implies that the discontinuity for the elastic cut is
618: \be
619: \tilde D(m^2) =
620: \tilde A(m^2)e^{-i\delta_{cs}}\sin{\delta_{cs}}.
621: \ee
622: The solution to Eq.~\eqref{dispers1} is found in complete analogy to
623: the case without the presence of the Coulomb interaction,
624: \be
625: \tilde A(m^2)=\exp\left[{\frac1\pi\int_{m_0^2}^{m_{max}^2}\frac{\delta_{cs}(m' \, ^2)}{m' \, ^2-m^2-i0}dm' \, ^2}\right]
626: \tilde \Psi(m^2) \label{C_intc} \ ,
627: \label{A_def1}
628: \ee
629: where $\tilde\Psi(m^2)$ is some function slowly varying with $m^2$.
630: If one neglects the the weak $m^2$ dependence present in $\tilde\Psi(m^2)$
631: the expression for the phase shift $\delta_{cs}$ in terms of the
632: differential cross section becomes
633: %\be
634: %\nonumber
635: %&&\frac{\delta_{cs}(m^2)}{\sqrt{m^2-m_0^2}}\approx \\
636: %& & -\frac1{2\pi}{\bf P}
637: %\int_{m_0^2}^{m_{max}^2}
638: %\frac{\log{\left[\displaystyle{\frac{1}{p'C^2(p')}\frac{d^2\sigma}{dm' \, ^2dt}}\right]}}
639: %{\sqrt{m' \, ^2-m_0^2} \ (m' \, ^2-m^2)}\sqrt{\frac{m_{max}^2-m^2}{m_{max}^2-m' \, ^2}}dm' \, ^2.
640: %\label{finalc}
641: %\ee
642: \be
643: \frac{\delta_{cs}(m^2)}{\sqrt{m^2-m_0^2}} =
644: -\frac1{2\pi}{\bf P}
645: \int_{m_0^2}^{m_{max}^2}
646: \frac{\log{\left[\displaystyle{\frac{1}{p'C^2(p')}\frac{d^2\sigma}{dm' \, ^2dt}}\right]}}
647: {\sqrt{m' \, ^2-m_0^2} \ (m' \, ^2-m^2)}\sqrt{\frac{m_{max}^2-m^2}{m_{max}^2-m' \, ^2}}dm' \, ^2.
648: \label{finalc}
649: \ee
650: Using the effective range expansion (\ref{eff_range}) one can then extract the
651: scattering length $a_{cs}$ from this dispersion integral.
652:
653: \section{Test of the method for the Coulomb case}
654:
655: One of the obvious reactions for applying the formalism with
656: Coulomb is $pp\to K^0\Sigma^+p$ where one could determine the
657: $\Sigma N$ scattering length for the isospin 3/2 state.
658: Note that the $\Sigma^+p$ channel does not couple to the $\Lambda N$
659: system and is therefore free of inelastic cuts (that start already on
660: the left-hand side) as
661: required for the applicability of the disperson integral method.
662: For this reaction one could perform a model calculation analogous to
663: the one for $pp\to K^+\Lambda p$ \cite{model} which we used for
664: testing the dispersion-integral method in the absence of the
665: Coulomb interaction \cite{Gasparyan2003}.
666: However, the implementation
667: of Coulomb effects into our momentum-space code is technically
668: complicated and requires also some approximations \cite{Hanhart}.
669: Thus, for the present test calculation we adopt a different
670: strategy. First, instead of the momentum-space $YN$ models
671: of Refs. \cite{Holz,Melni2,Haiden} we take the r-space Argonne
672: ($NN$) potential, however, with parameters modified in such a way
673: that the effective range parameters are similar to those
674: predicted by realistic $YN$ potentials \cite{Holz,NijV,Melni2,Haiden}
675: for the $\Sigma N$ $I=3/2$ $^1S_0$ partial wave.
676: In particular we prepared two models with Coulomb modified
677: scattering length of $a_{cs}=-3.24$ fm (model 1) and
678: $a_{cs}=-1.86$ fm (model 2), respectively. The corresponding scattering
679: lengths without
680: Coulomb interaction are $-4.11$ fm and $-2.01$ fm, respectively.
681: %
682: For the transition amplitude we use the scattering wave function
683: calculated from those potential models and evaluated at the origin.
684: This corresponds to the assumption that the production operator is
685: point-like, which is reasonable as long as we are interested only in the
686: mass dependence of the production amplitude. The corrections
687: stemming from a possible mass dependence of the production
688: operator were discussed in Ref.~\cite{Gasparyan2003}.
689:
690: The results of applying Eq.~\eqref{final} with $m_{max}-m_0=40$ MeV
691: ($p_{max}=205$ MeV/c)
692: are shown in Fig.~\ref{fig1}, where we plot the function
693: $1/S(p)$ which should coincide with the scattering length $-a_{cs}$
694: at $p=0$. Obviously, there is a strongly nonanalytic behavior of the
695: extracted inverse effective range function when approaching the
696: threshold -- which, however, can be easily understood.
697: It is clear from Eq.~\eqref{eff_range} that the threshold
698: behavior of the Coulomb modified phase shift is
699: $\delta_{cs}\approx -a_{cs}\,p\,C^2(p)$, i.e. $\delta_{cs}$
700: goes to zero very rapidly. To obtain such a
701: behavior on the left-hand side of Eq.~\eqref{finalc}
702: one needs to have a very precise cancellation in the integral
703: on the right-hand side of Eq.~\eqref{finalc} that is,
704: of course, impossible if one truncates the integral
705: at a finite momentum. But still one can expect
706: for Eq.~\eqref{finalc} to work for momenta not too close to the threshold,
707: namely above the typical Coulomb scale of $2\pi\alpha/\mu\approx 25$ MeV/c
708: where the factors $C^2(p)$ and $Q(p)$ that appear in the effective range
709: function (Eq. (\ref{eff_range})) become smoother.
710: This is indeed the case, as can be seen from
711: Fig.~\ref{fig1}. Thus, a natural and practical step here would be to
712: extrapolate the extracted $S(p)$ to the threshold from above.
713: Using a 4-th order polynomial of the type $-1/a_{cs}+r_ep^2/2-Pr_e^3p^4$
714: and fixing the coefficients
715: in the region $50-100$ MeV/c, i.e. well above the Coulomb structure,
716: one can then extrapolate $S(p)$ to the threshold,
717: cf. the dash-dotted lines in Fig.~\ref{fig1}.
718: In this case one gets a satisfactory agreement between the true and extracted
719: scattering lengths. In fact, the deviations are not worse
720: then in the case when we consider the same potentials without
721: Coulomb interaction and they are also within the theoretical error of
722: 0.3 fm estimated in Ref. \cite{Gasparyan2003}. The extracted values are
723: $a_{cs}=-3.10$ fm and $a_{cs}=-1.86$ fm for models 1 and 2,
724: respectively, with Coulomb, and $a_s=-4.05$ fm and $a_s=-2.06$ fm
725: when the Coulomb interaction is switched off. We also checked that
726: the sensitivity of the result to the region of interpolation of the
727: effecitve range function $S(p)$ is rather low. For instance if one shifts
728: the lower bound of this region to $70$ MeV/c the corresponding change
729: in the scattering length will be less then $0.05$ fm.
730:
731: If one wishes to consider a more realistic
732: situation, one needs to deal with mass distributions with
733: finite statistical errors, finite mass resolution,
734: and finite binning. To examine also this situation we have generated
735: two data sets, corresponding to the models 1 and 2
736: as shown on Fig.~\ref{fig2}. We have chosen the
737: binning as well as the mass resolution to be equal to $2$ MeV
738: (the same as in the experiment \cite{Siebert} that was analyzed in \cite{Gasparyan2003}),
739: and the statistics to be rather high in order to minimize the influence of
740: the statistical error bars on the results. The excess energy was set to $40$
741: MeV to simplify the simulation. In a realistic situation larger values are
742: preferable to minimize the influence of the meson-baryon interactions,
743: cf. the corresponding remarks in Ref. \cite{Gasparyan2003}.
744: In our test calculation such meson-baryon interactions are neglected
745: anyway.
746:
747: We start here with the procedure suggested in appendix A of \cite{Gasparyan2003},
748: namely by fitting the cross section with an exponential parameterization of the type
749: \be
750: \frac{d^2\sigma}{dm \, ^2dt}=C^2(p) \times \exp{\left[C_0+\frac{C_1^2}{(m^2-C_2^2)} \right]}
751: \times \text{phase space} \ .
752: \label{fit1}
753: \ee
754: This formula fits the generated cross section with
755: the $\chi^2$ per degree of freedom of $\chi^2_{dof}\sim 1$, cf. Fig.~\ref{fig2}.
756: A new problem that arises here is that the
757: production amplitude contains a very narrow structure
758: (of the size $2\pi\alpha/\mu$) close to threshold as can
759: be seen in Fig.~\ref{fig3} (solid lines).
760: Clearly, this structure can not be reproduced after the fitting
761: procedure as it gets smeared out by the mass resolution and
762: binning (their size is much larger than the scale of the structure).
763: The amplitudes coming from the fit are depicted in Fig.~\ref{fig3}
764: by dash-dotted lines.
765: However the fit can be improved if one notes that
766: the structure comes mostly from the part of the dispersion integral
767: \eqref{A_def1} containing the leading term in the
768: $\delta_{cs}$ expansion near the threshold, namely (in the nonrelativistic case)
769: \be
770: \exp\left[{\frac1\pi\int_{0}^\infty\frac{-a_{cs}C^2(p')}{p' \, ^2-p^2-i0}\left(\frac{p^2}{p'^2}\right)dp' \, ^2}\right]=\exp\left[-a_{cs}Q(p)\right],
771: \ee
772: where we made a subtraction at $p=0$ to render the integral convergent.
773: This does not change the energy dependence of the resulting exponent.
774: Indeed the structure disappears after dividing the production amplitude
775: by the factor $\exp[-a_{cs}Q(p)]$.
776:
777: Obviously the scattering length is unknown before its
778: extraction! But what can be done is to resort to an iterative procedure
779: including first the extraction of the unimproved scattering length,
780: then putting it into the fit function,
781: \be
782: \frac{d^2\sigma}{dm \, ^2dt}=C^2(p)\times\exp{\left[C_0+\frac{C_1^2}{(m^2-C_2^2)}
783: -2a_{cs}Q(p) \right]} \times \text{phase space} \ ,
784: \label{fit2}
785: \ee
786: and repeating this step until the procedure converges. Fortunately
787: the iterations converge already after three or four iterations,
788: and the resulting amplitudes are shown in Fig.~\ref{fig3}
789: by the dashed lines. The improvement of the fit is quite
790: obvious.
791: Finally we applied the combination of the extrapolation and iteration
792: procedures to obtain the scattering lengths from the pseudodata.
793: Since the data have a statistical uncertainty we generated a sample of
794: 1000 mass distributions and looked at the average value of the scattering
795: lengths. They turned out to be $-2.89\pm 0.06$ fm for model 1 and $-1.82\pm
796: 0.05$ for model 2. Note that the deviation from the correct values is now a
797: bit larger, but still reasonable ($~0.35$ fm in the worst case).
798: Fig.~\ref{fig4} shows how the extracted inverse effective range function
799: approaches to the correct one given by the model by the example of model 1.
800:
801: \section{Summary}
802:
803: In a recent publication \cite{Gasparyan2003} we have presented a formalism
804: based on dispersion relations
805: that allows one to relate spectra from large-momentum transfer reactions,
806: such as $pp\to K^+p\Lambda$ or $\gamma d\to K^+n\Lambda$, directly to
807: the scattering length of the interaction of the final-state particles.
808: An estimation of the systematic uncertainties of that method, relying on general
809: arguments, led to the conclusion that the theoretical error in the extracted
810: scattering length should be less than 0.3 fm. This finding was corroborated
811: in an application of the method to results of a microscopic model calculation
812: for $pp\to K^+p\Lambda$.
813:
814: In the present paper this dispersion theoretical method was generalized to
815: the case where a repulsive Coulomb force is present in the final-state interaction.
816: As an example let us mention the reaction $pp\to K^0p\Sigma^+$ which could
817: be used to extract the $p\Sigma^+$ scattering length.
818: %
819: Though the generalization of the formalism itself is
820: straight forward it turned out that there are some additional features
821: due to the Coulomb interaction which need to be taken into account
822: in concrete applications of the method to data.
823: %
824: These practical aspects were thoroughly discussed and it was shown how to
825: circumvent the difficulties. In a test calculation utilizing
826: potential models with effective range parameters similar to those
827: of realistic $YN$ interactions the extracted values for the scattering lengths
828: were found to agree within 0.3 fm with those predicted by the models.
829: Thus, the accuracy of the dispersion theoretical method for extracting the
830: scattering lengths from final-state interactions with Coulomb force is
831: comparable to the case where no Coulomb interaction is present.
832:
833: We presented also a more detailed examination of the accuracy of the
834: dispersion-integral method than in Ref. \cite{Gasparyan2003}. In particular we considered
835: final-state interactions of varying strengths, corresponding to a much larger range
836: of values of the scattering length. These investigations confirmed the reliability
837: of the general error estimate provided in Ref. \cite{Gasparyan2003}. Indeed in
838: most of the considered cases the deviation of the extracted scattering length
839: from the true value was significantly smaller than the uncertainty of 0.3 fm
840: derived in that paper.
841: %
842: In addition we studied the effective range $r_e$ which can be also extracted
843: by the proposed dispersion-integral method. For most of the interaction models
844: considered the extracted values of $r_e$ agreed remarkably good with the
845: true results. Thus, it might be sensible to use the proposed method to extract
846: the effective range from data - though one should always keep in mind that
847: for this quantity a generally valid error estimation is not possible
848: \cite{Gasparyan2003}.
849:
850: Finally, we compared the present method with the performance of other,
851: approximative treatments
852: of the final-state interaction that are commonly used in the literature to
853: extract information on the scattering length and also the effective range.
854: In particular we tested the Jost-function approach based on the effective-range
855: approximation (Jost-ERA) and an even simpler approach that relies simply on utilizing
856: the effective range approximation itself.
857: Thereby, we showed that the latter methods lead to systematic deviations
858: from the true values of the scattering lengths in the order of 0.3 fm (Jost-ERA)
859: and even 0.7 fm (direct effective-range approximation).
860: This suggests that one should be rather cautious in the interpretation of results
861: achieved with those methods.
862:
863: \acknowledgments{
864: A.G. would like to acknowledge finanical support by the
865: grant No. 436 RUS 17/75/04 of the Deutsche Forschungsgemeinschaft.
866: Furthermore he thanks the Institut f\"ur Kernphysik at the
867: Forschungszentrum J\"ulich for its hospitality during the period when
868: the present work was carried out.
869: }
870:
871: \begin{thebibliography}{10}
872: %%%%%%%%%%%%%%%% INTRO %%%%%%%%%%%%
873: \bibitem{Joach} See, e.g., C.J. Joachain, {\it Quantum Collision
874: Theory}, (North-Holland, Amsterdam 1975).
875:
876: \bibitem{Miller} G.A. Miller, B.M.K. Nefkens, and I. Slaus,
877: Phys. Rep. {\bf 194}, 1 (1990).
878:
879: \bibitem{Dover} C.B. Dover and H. Feshbach,
880: Ann. Phys. {\bf 198}, 321 (1990).
881:
882: \bibitem{xx} J.~Gasser, H.~Leutwyler and M.~E.~Sainio,
883: %``Sigma Term Update,''
884: Phys.\ Lett.\ B {\bf 253}, 252 (1991).
885:
886: % eta N bound states
887: \bibitem{Wycech}
888: S. Wycech, A.M. Green, and J.A. Niskanen, Phys. Rev.
889: C {\bf 52}, 544 (1995).
890: \bibitem{Belyaev1}
891: S.A. Rakityansky, S.A. Sofianos, W. Sandhas,
892: and V.B. Belyaev, Phys. Lett. B {\bf 359}, 33 (1995).
893: \bibitem{Belyaev2}
894: V.B. Belyaev, S.A. Rakityansky, S.A. Sofianos, M. Braun, and
895: W. Sandhas, Few. Body. Syst. Suppl. {\bf 8}, 309 (1995).
896: \bibitem{Rakityansky}
897: S.A. Rakityansky, S.A. Sofianos, M. Braun, V.B. Belyaev
898: and W. Sandhas, Phys. Rev. C {\bf 53}, R2043 (1996).
899: \bibitem{Fix1}
900: A. Fix and H. Arenh\"ovel, Phys. Rev. C {\bf 66}, 024002 (2002).
901: \bibitem{Niskanen} A. Sibirtsev, J. Haidenbauer, J.A. Niskanen, and Ulf-G. Mei{\ss}ner,
902: Phys. Rev. C {\bf 70}, 047001 (2004).
903: \bibitem{Niskanen1}
904: J.A. Niskanen, A. Sibirtsev, J. Haidenbauer, and C. Hanhart,
905: Int. J. Mod. Phys. A {\bf 20}, 634 (2005).
906: \bibitem{Haider2}
907: Q. Haider and L.C. Liu, Phys. Rev. C {\bf 66}, 045208 (2002).
908: \bibitem{rus}
909: V.~Antonelli, A.~Gall, J.~Gasser and A.~Rusetsky,
910: %``Effective Lagrangians in bound state calculations,''
911: Ann. Phys.\ {\bf 286}, 108 (2001).
912: %
913: \bibitem{Gotta}
914: D. Gotta, Prog. Part. Nucl. Phys. {\bf 52}, 133 (2004).
915:
916: \bibitem{Gibbs}
917: W.R. Gibbs, S.A. Coon, H.K. Han, and B.F. Gibson,
918: Phys. Rev. C {\bf 61}, 064003 (2000).
919:
920: \bibitem{Trotter}
921: D.E. Gonz\'alez Trotter et al.,
922: Phys. Rev. Lett. {\bf 83}, 3788 (1999).
923:
924: \bibitem{Huhn}
925: V. Huhn et al.,
926: Phys. Rev. C {\bf 63}, 014003 (2000).
927:
928: \bibitem{Deng}
929: J. Deng, A. Siepe, and W. von Witsch,
930: Phys. Rev. C {\bf 66}, 047001 (2002).
931:
932: \bibitem{xxx}
933: S.~R.~Beane, V.~Bernard, E.~Epelbaum, U.~G.~Meissner and D.~R.~Phillips,
934: %``The S-wave pion nucleon scattering lengths from pionic atoms using
935: %effective field theory,''
936: Nucl.\ Phys.\ A {\bf 720} (2003) 399
937: [arXiv:hep-ph/0206219];
938: A.~Gardestig and D.~R.~Phillips,
939: %``Using chiral perturbation theory to extract the neutron neutron scattering
940: %length from pi- d $\to$ n n gamma,''
941: arXiv:nucl-th/0501049;
942: \bibitem{Lensky}
943: V.~Lensky, V.~Baru, J.~Haidenbauer, C.~Hanhart, A.~E.~Kudryavtsev and U.~G.~Meissner,
944: %``Precision calculation of gamma d $\to$ pi+ n n within chiral perturbation
945: %theory,''
946: arXiv:nucl-th/0505039.
947:
948: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
949: \bibitem{Gasparyan2003}
950: A.~Gasparyan, J.~Haidenbauer, C.~Hanhart and J.~Speth,
951: %``How to extract the Lambda N scattering length from production reactions,''
952: Phys.\ Rev.\ C {\bf 69}, 034006 (2004)
953:
954: \bibitem{C11_1}
955: J.~T.~Balewski {\it et al.},
956: Phys. Lett. {\bf B 420}, 211 (1998);
957: S. Sewerin {\it et al.},
958: Phys. Rev. Lett. {\bf 83}, 682 (1999).
959:
960: \bibitem{Bilger}
961: R. Bilger {\it et al.},
962: Phys. Lett. {\bf B 420}, 217 (1998).
963:
964: \bibitem{jan}
965: J.~T.~Balewski {\it et al.},
966: %``Low-energy Lambda p scattering parameters from the p p $\to$ p K+ Lambda reaction,''
967: Eur.\ Phys.\ J.\ A {\bf 2}, 99 (1998).
968:
969: \bibitem{Renard}
970: F.M. Renard and Y. Renard,
971: Nucl. Phys. {\bf B1}, 389 (1967).
972:
973: \bibitem{Kerbikov}
974: B.O. Kerbikov, B.L.G. Bakker, and R. Daling,
975: Nucl. Phys. {\bf A480}, 585 (1988);
976: B.O. Kerbikov, Phys. Atom. Nucl. {\bf 64}, 1835 (2001).
977:
978: \bibitem{Mecking}
979: B. Mecking et al., CEBAF proposal PR-89-045 (1989).
980:
981: \bibitem{Adel}
982: R.A. Adelseck and L.E. Wright,
983: Phys. Rev. C {\bf 39}, 580 (1989).
984:
985: \bibitem{Li}
986: X. Li and L.E. Wright,
987: J. Phys. G {\bf 17}, 1127 (1991).
988:
989: \bibitem{Yama}
990: H. Yamamura, K. Miyagawa, T. Mart, C. Bennhold, H. Haberzettl, and
991: W. Gl\"ockle,
992: Phys. Rev. C {\bf 61}, 014001 (1999).
993:
994: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
995:
996: \bibitem{Howell}
997: C.R. Howell et al.,
998: Phys. Lett. B {\bf 444}, 252 (1998).
999:
1000: \bibitem{Psi} see, e.g., F. A. Harris,
1001: Int. J. Mod. Phys. A{\bf 20}, 445 (2005),
1002: and references therein.
1003:
1004: \bibitem{Belle} see, e.g., J. Schumann, hep-ex/0505098,
1005: and references therein.
1006: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1007: \bibitem{book}
1008: M. Goldberger and K.M. Watson, {\it Collision Theory} (Wiley, New York, 1964).
1009:
1010: \bibitem{migdal}
1011: K. Watson, Phys. Rev. {\bf 88}, 1163 (1952);
1012: A.B. Migdal, Sov. Phys. JETP {\bf 1}, 2 (1955).
1013:
1014: \bibitem{Geshkenbein1969}
1015: B.~V.~Geshkenbein,
1016: %``Determination Of The Electromagnetic Formfactor In The Space-Like Region From The Value Of Its Modulus In The Time-Like Region,''
1017: Yad.\ Fiz.\ {\bf 9}, 1232 (1969) [Sov. J. Nucl.
1018: Phys. {\bf 9}, 720 (1969)].
1019:
1020: \bibitem{Geshkenbein1998}
1021: B.~V.~Geshkenbein,
1022: %``Pion electromagnetic formfactor in the space-like region and P-phase delta(1)(1)(s) of pi pi scattering from the value of the modulus of
1023: %formfactor in the time-like region,''
1024: Phys.\ Rev.\ D {\bf 61}, 033009 (2000).
1025:
1026:
1027: \bibitem{Muskhelishvili1953}
1028: N.~I.~Muskhelishvili,
1029: {\it Singular Integral Equations},
1030: (P.~Noordhof N.~V., Groningen, 1953).
1031:
1032: \bibitem{Omnes1958}
1033: R.~Omnes,
1034: %``On The Solution Of Certain Singular Integral Equations Of Quantum Field Theory,''
1035: Nuovo Cim.\ {\bf 8}, 316 (1958).
1036:
1037: \bibitem{Frazer1959}
1038: W.~R.~Frazer and J.~R.~Fulco,
1039: %``Effect Of A Pion-Pion Scattering Resonance On Nucleon Structure,''
1040: Phys.\ Rev.\ Lett.\ {\bf 2}, 365 (1959).
1041:
1042: \bibitem{Bale} J.~T.~Balewski {\it et al.},
1043: Phys. Lett. {\bf B 420}, 211 (1998);
1044: S. Sewerin {\it et al.},
1045: Phys. Rev. Lett. {\bf 83}, 682 (1999).
1046:
1047: \bibitem{Hinter} F. Hinterberger and A. Sibirtsev,
1048: Eur. Phys. J. A {\bf 21}, 313 (2004).
1049:
1050: %%%%%%%%%%% YN and NN models %%%%%%%%%%%%%%%%%%%%%%%%%
1051: \bibitem{Holz}
1052: B. Holzenkamp, K. Holinde, and J. Speth,
1053: Nucl. Phys. {\bf A500}, 485 (1989).
1054:
1055: \bibitem{NijV}
1056: Th.A. Rijken, V.G. Stoks, and Y. Yamamoto,
1057: Phys. Rev. C {\bf 59}, 21 (1999).
1058:
1059: \bibitem{Melni2}
1060: J. Haidenbauer, W. Melnitchouk and J. Speth,
1061: AIP Conf. Proc. {\bf 603}, 421 (2001),
1062: [arXiv:nucl-th/0108062].
1063:
1064: \bibitem{Haiden}
1065: J. Haidenbauer and U.-G. Mei{\ss}ner,
1066: nucl-th/0506019.
1067:
1068: \bibitem{model}
1069: A.~Gasparian, J.~Haidenbauer, C.~Hanhart, L.~Kondratyuk and J.~Speth,
1070: %``The reactions p p $\to$ p Lambda K+ and p p $\to$ p Sigma0 K+ near their thresholds,''
1071: Phys.\ Lett.\ B {\bf 480}, 273 (2000).
1072:
1073: \bibitem{Argonne}
1074: R.B. Wiringa, R.A. Smith, and T.L. Ainsworth,
1075: Phys. Rev. C {\bf 29}, 1207 (1984).
1076:
1077: %%%%%%%%%%% YN etc. DATA %%%%%%%%%%%%%%%%%%%%%%%%%
1078: \bibitem{Alex} G.~Alexander et al.,
1079: Phys. Rev. {\bf 173}, 1452 (1968).
1080:
1081: \bibitem{Sechi} B.~Sechi-Zorn et al.,
1082: Phys. Rev. {\bf 175}, 1735 (1968).
1083:
1084: \bibitem{Siebert}
1085: R. Siebert et al.,
1086: Nucl.\ Phys.\ A {\bf 567}, 819 (1994).
1087:
1088: %%%%%%%%%%% COULOMB %%%%%%%%%%%%%%%%%%%%%%%%%
1089: \bibitem{Gell}
1090: M. Gell-Mann and M.L. Goldberger,
1091: Phys. Rev. {\bf 91}, 398 (1953).
1092:
1093: \bibitem{vanHaeringen1976}
1094: H.~van Haeringen,
1095: %``Coulombian Asymptotic States,''
1096: J.\ Math.\ Phys.\ {\bf 17}, 995 (1976).
1097:
1098: \bibitem{Hamilton1973}
1099: J.~Hamilton, I.~Oeverboe and B.~Tromborg,
1100: %``Coulomb Corrections In Non-Relativistic Scattering,''
1101: Nucl.\ Phys.\ B {\bf 60}, 443 (1973).
1102:
1103: \bibitem{Cornille1962}
1104: H.~Cornille and A.~Martine,
1105: Nuovo Cimento {\bf 26}, 298 (1962).
1106:
1107: \bibitem{vanHaeringen1977}
1108: H.~van Haeringen,
1109: %``T Matrix And Effective Range Function For Coulomb Plus Rational Separable
1110: %Potentials Especially For L = 1,''
1111: J.\ Math.\ Phys.\ {\bf 18}, 927 (1977).
1112:
1113: \bibitem{Heller1966}
1114: L.~Heller and M.~Rich,
1115: Phys.\ Rev.\ {\bf 144}, 1324 (1966)
1116:
1117: \bibitem{Scotti1965}
1118: A.~Scotti and D.~Y.~Wong,
1119: Phys.\ Rev.\ {\bf 138}, B145 (1965)
1120:
1121: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1122: \bibitem{Hanhart}
1123: C. Hanhart, J.~Haidenbauer, A. Reuber, C. Sch\"utz, and J.~Speth,
1124: Phys.\ Lett.\ B {\bf 358}, 21 (1995).
1125:
1126: \end{thebibliography}
1127:
1128: \newpage
1129: \begin{figure}[h]
1130: \begin{center}
1131: \epsfig{file=methodtest.eps, width=13.0cm}
1132: \end{center}
1133: \caption{(Color online) Comparison of different extraction methods for the
1134: scattering length $a$. Shown are the differences between results
1135: predicted by various $YN$ and $NN$ models and
1136: corresponding values extracted via the dispersion integral method
1137: (circles), the Jost-ERA approach (\ref{arform}) (squares)
1138: and the Migdal-Watson prescription (\ref{MW}) (triangles).
1139: The lines are drawn to guide the eye.
1140: }
1141: \label{fig0}
1142: \end{figure}
1143:
1144: \newpage
1145: \begin{figure}[h]
1146: \begin{center}
1147: \epsfig{file=methodtest_r.eps, width=13.0cm}
1148: \end{center}
1149: \caption{(Color online) Comparison of different extraction methods for the
1150: effective range $r_e$. Shown are the differences between results
1151: predicted by various $YN$ and $NN$ models and
1152: corresponding values extracted via the dispersion integral method
1153: (circles) and the Jost-ERA approach (\ref{arform}) (squares).
1154: The lines are drawn to guide the eye.
1155: }
1156: \label{fig0a}
1157: \end{figure}
1158:
1159: \newpage
1160: \begin{figure}[h]
1161: \begin{center}
1162: \epsfig{file=eff_range.eps, width=15.0cm}
1163: \end{center}
1164: \caption{The inverse of the effective range function $S(p)$
1165: for model 1 (upper curves) and model 2 (lower curves). The
1166: solid lines denote the phase shifts predicted by the corresponding
1167: model, whereas the dashed lines correspond to the phase shift extracted
1168: via Eq.~\eqref{finalc}. The dash-dotted lines show the result of
1169: the smooth extrapolation of the dashed lines, as explained in the text.}
1170: \label{fig1}
1171: \end{figure}
1172:
1173: \newpage
1174: \begin{figure}[h]
1175: \begin{center}
1176: \epsfig{file=XS.eps, width=15.0cm}
1177: \end{center}
1178: \caption{Pseudo data for the
1179: differential $\Sigma^+p$ cross section generated from
1180: model 1 (circles) and model 2 (squares) as a function of
1181: the $\Sigma p$ invariant mass $M_{\Sigma p}$ with corresponding
1182: fit by the exponential parameterization Eq. (\ref{fit1}).
1183: }
1184: \label{fig2}
1185: \end{figure}
1186:
1187: \newpage
1188: \begin{figure}[h]
1189: \begin{center}
1190: \epsfig{file=smearing.eps, width=15.0cm}
1191: \end{center}
1192: \caption{The production amplitude $A(m^2)$ divided by the factor $C(p)$
1193: for model 1 (upper curves) and model 2 (lower curves).
1194: The solid lines are the amplitudes as calculated from the models. The dash-dotted lines
1195: correspond to the fitted amplitudes. The dashed lines denote the fitted
1196: amplitudes improved by the iterative procedure as discussed in the text.}
1197: \label{fig3}
1198: \end{figure}
1199:
1200:
1201: \newpage
1202: \begin{figure}[h]
1203: \begin{center}
1204: \epsfig{file=iterations.eps, width=15.0cm}
1205: \end{center}
1206: \caption{The inverse of the effective range function $S(p)$ calculated by means of
1207: the iterative procedure as discussed in the text. The dotted line corresponds
1208: to the zeroth iteration, the dashed line corresponds to the first
1209: iteration, and the dash-dotted line corresponds to the second iteration. The
1210: solid curve denotes the exact result. Shown are results for the
1211: model 1.
1212: }
1213: \label{fig4}
1214: \end{figure}
1215:
1216: \end{document}
1217:
1218: