1: \documentstyle[prc,aps,graphicx,floats,amsmath,amsfonts]{revtex}
2: %\documentstyle[prc,aps,epsf,epsfig,floats,twocolumn,amsmath,amsfonts]{revtex}
3:
4: \setlength{\textwidth}{7in}
5: \setlength{\textheight}{9.5in}
6:
7: \newcommand{\bra}{\left \langle}
8:
9: \newcommand{\ket}{\right \rangle}
10:
11: \newcommand{\Ref}[1]{(\ref{#1})}
12:
13: \newcommand{\Half}{\textstyle{\frac{1}{2}}}
14:
15: \newcommand{\I}{\imath }
16:
17: \newcommand{\D}{\textstyle{\rm d}}
18:
19: \newcommand{\E}{\textstyle{\rm e}}
20:
21: \begin{document}
22:
23: \title{\bf Determination of $\pi N$ scattering lengths from
24: pionic hydrogen and pionic deuterium data}
25: \author{ A.~Deloff }
26:
27: \address {\normalsize Soltan Institute for Nuclear Studies,
28: Hoza~69, 00-681~Warsaw, Poland }
29: \maketitle
30:
31: \begin{abstract}
32: The $\pi$N s-wave scattering lengths have been inferred from a joint
33: analysis of the pionic hydrogen and the pionic deuterium x-ray data using a
34: non-relativistic approach in which the $\pi$N interaction is simulated
35: by a short-ranged potential. This potential is assumed to be isospin
36: invariant and its range, the same for isospin I=3/2 and I=1/2, is
37: regarded as a free parameter. The proposed model admits an exact
38: solution of the pionic hydrogen bound state problem,
39: i.e. the $\pi$N scattering
40: lengths can be expressed analytically in terms of the range parameter
41: and the shift ($\epsilon$) and width ($\Gamma$)
42: of the 1s level of the pionic hydrogen. We
43: demonstrate that for small shifts and short ranges from the exact
44: expression one retrieves the standard
45: range independent Deser-Trueman formula. The $\pi$d
46: scattering length has been calculated exactly
47: by solving the Faddeev equations
48: and also by using a static approximation. It has been shown that the same
49: very accurate static formula for $\pi$d scattering length can be
50: derived (i) from a set of boundary conditions; (ii) by a reduction
51: of Faddeev equations; and (iii) through a summation of Feynman
52: diagrams. By imposing the requirement that the $\pi$d scattering
53: length, resulting from Faddeev-type calculation, be in agreement
54: with pionic deuterium data, we obtain bounds on the $\pi$N
55: scattering lengths.
56: The dominant source of uncertainty on the deduced values
57: of the $\pi$N scattering lengths are the experimental errors in the
58: pionic hydrogen data.
59:
60: \medskip\noindent
61:
62: {PACS numbers: 11.80.Jy, 13.75.Gx, 25.80.Dj, 25.80.Hp, 36.10.Gv }
63:
64: \end{abstract}
65:
66: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67: \section{Introduction} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68: \label{se:one}
69: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70:
71: The determination of low-energy pion-nucleon ($\pi$N) parameters
72: has been the focus of much theoretical and experimental efforts.
73: The s-wave $\pi$N scattering lengths are of particular importance
74: serving as testing ground for various theoretical considerations.
75: In addition to that, their isovector combination provides input
76: in the Goldberger-Miyazawa-Oehme \cite{GMO} sum rule to be used to extract
77: the $\pi$NN coupling constant. In recent years major advances have been
78: made in the experimental and theoretical investigation of the
79: $\pi$N system. With the advent of meson factories (LAMPF, PSI and
80: TRIUMF) and the corresponding influx of the new high accuracy
81: $\pi$N scattering data considerable progress has been achieved in the
82: $\pi$N phase shift analyses \cite{SAID,GIB98,Gashi}
83: providing means to examine even such
84: subtleties as isospin symmetry breaking effects \cite{GIB98,Li,MAT97}
85: Recently, the $\pi$N
86: scattering experiments have been complemented by high quality pionic
87: x-ray measurements performed, both on pionic hydrogen \cite{Sigg,SCH99}
88: and on pionic deuterium \cite{HAU98}.
89: The measurements of the shifts and widths in the 1s levels
90: in these atomic systems, resulting from strong $\pi$N interaction,
91: allows to extract directly the corresponding scattering lengths,
92: i.e. $a_{\pi p}$ and $a_{\pi d}$, respectively. Therefore, the new x-ray
93: data constitute an independent source of information on the low-energy
94: $\pi$N scattering parameters. On the theoretical side, the physical
95: quantities bearing on the low-energy $\pi$N interaction have now
96: become accessible to calculations \cite{Nadia} conducted within quantum
97: chromodynamics (QCD). Since QCD is known to be highly non-perturbative
98: at low energies, its low-energy implementation has been based instead
99: on a chiral perturbation theory in which the effective Lagrangian
100: is expanded in increasing powers of derivatives in meson fields and
101: quark masses. This approach in practice involves a Taylor expansion
102: in the meson four-momenta and therefore it may be expected that the lower
103: the energy is, the more accurate are the predictions. In this context,
104: the precise knowledge of the experimental values of the low-energy
105: $\pi$N scattering parameters is essential for further development of
106: the theory.
107: \par
108: The purpose of this work is to extract the s-wave $\pi$N scattering
109: lengths using exclusively the pionic hydrogen and pionic deuterium
110: x-ray data.
111: The key reason for proceeding along this route is that
112: the low-energy regime can be thereby investigated without recourse to
113: scattering data and
114: there is no danger that the low-energy parameters have been
115: largely determined by the data at high energies. Our treatment
116: is purely phenomenological based on an isospin invariant potential
117: model and we wish to clarify at the onset that this approach
118: relinquishes any pretense of being a theory in favour of practicable
119: calculational scheme.
120: The investigation has two parts. In part one we take as our input
121: the values of the $\pi$N scattering lengths determined previously
122: from pionic hydrogen data and use them in a microscopic calculation
123: of the $\pi$d scattering length. The latter has not been measured
124: directly in a scattering experiment but may be
125: extracted from the pionic deuterium x-ray data by applying the
126: Deser-Trueman formula \cite{Deser}.
127: It is an empirical fact that the $\pi$N scattering lengths are small as
128: compared with the deuteron size and it has been a common practice
129: \cite{BARU}
130: to use the multiple scattering expansion for calculating the $\pi$d
131: scattering length. Since this series rapidly converges, what has
132: been confirmed by early Faddeev calculations \cite{PETROV,AFN74,MIZ77},
133: in the past with the poorly known $\pi$d scattering length
134: there was little incentive to go beyond the second order
135: (for a review, cf. \cite{Judah,THO80,ERI88}).
136: At present, the experimental error on $\pi$d scattering length is
137: at the level of 2\% and the adequacy of the
138: second order formula might be
139: questionated. Strictly speaking, a truncation
140: of the multiple scattering series
141: can really only receive its justification when we
142: actually quantify the magnitude of the higher-order terms
143: to establish whether they are truly negligible.
144: This question is examined in detail in this paper and
145: the $\pi$d zero-energy scattering
146: problem is solved exactly within a three-body
147: formalism by introducing a zero-range
148: model to simulate the $\pi$N s-wave interaction. One advantageous
149: feature of this model is that it allows to obtain an analytic solution of
150: the three-body problem in the static approximation. We demonstrate
151: that the static solution can be obtained either by reduction of the
152: Faddeev equations, or by imposing a suitable set of boundary
153: conditions, or finally by performing a summation of Feynman diagrams.
154: All three methods converge to the same analytic formula expressing
155: the $\pi$d scattering length in terms of the $\pi$N scattering lengths.
156: Static solution in coordinate space is very appealing and helps to
157: develop an intuitive picture of how the individual $\pi$N
158: amplitudes contribute to build up the $\pi$d scattering length.
159: By solving numerically the Faddeev equations we show that the accuracy
160: of the static approximation is comparable with the present experimental
161: uncertainty on $a_{\pi d}$. In order to find out what the pionic
162: deuterium data can teach us about the $\pi$N scattering lengths,
163: the $\pi$d scattering lengths obtained as a solution of the Faddeev
164: equations is compared with experiment. It turns out that the three-body
165: calculation is in agreement with experiment only when the input $\pi$N
166: scattering lengths belong to a relatively small subset of values
167: that are consistent with pionic hydrogen data. The $\pi$N
168: scattering lengths that belong to this subset simultaneously satisfy
169: the constraints imposed by the pionic hydrogen and pionic deuterium
170: data.
171: \par
172: In part two of the present work
173: we introduce explicitly a range parameter
174: in order to examine the validity of the
175: zero-range model.
176: To achieve this goal it is essential to devise a simple and transparent
177: representation of the $\pi$N interaction in which the two-body
178: scattering problem with and without Coulomb interaction admits an analytic
179: solution and we show that a two-channel isospin invariant separable
180: potential lends itself to that end.
181: Moreover, within this representation the exact
182: bound state condition appropriate for the pionic hydrogen problem
183: takes also an analytic form. The latter being a single complex
184: constraint, is equivalent to two real equations that can be explicitly
185: solved and as a result the $\pi$N scattering lengths are obtained
186: as functions of the range parameter together with the 1s level
187: shift and width in the pionic hydrogen. In particular, when the level
188: shift is small as compared with the Coulomb energy and the range
189: of the interaction is small in comparison with the Bohr radius,
190: from the exact bound state condition we retrieve the Deser-Trueman
191: formula (independent of the range parameter).
192: Regarding the range as a free parameter we are able to extend the
193: zero-range model and by varying this parameter in physically reasonable
194: limits we find the results to be insensitive to the value of the range.
195: The uncertainty on the $\pi$N scattering length caused by the lack
196: of knowledge of the range is much smaller than that resulting from
197: the experimental errors on the pionic hydrogen level shift and width.
198: \par
199: The organization of this paper is as follows. In Sec. \ref{se:two} we
200: develop a zero-range model and review various derivations leading
201: to the static solution of the $\pi$d scattering problem.
202: The accuracy of the static solution is examined by comparing it with the
203: solution of the Faddeev equations. We infer isoscalar and isovector
204: $\pi$N scattering lengths that are consistent with both, pionic hydrogen
205: and pionic deuterium data.
206: In Sec. \ref{se:three} we lift the zero-range
207: limitation by introducing a finite range into our formalism.
208: We present an exact treatment of the pionic hydrogen and we
209: derive Deser-Trueman formula for that particular case.
210: The $\pi$d scattering length obtained from the solution of the Faddeev
211: equation is compared with experiment.
212: Finally, the results are summarized in Sec. \ref{se:four}.
213:
214: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
215: \section{Zero-range model} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
216: \label{se:two}
217: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
218:
219: The central issue we wish to address in this section is how to
220: construct a theoretical framework in which we can use the pionic
221: deuterium data to gain information on the $\pi$N scattering lengths.
222: The measurement of the shift and the width of the 1s level in pionic
223: deuterium presents us with the value of $\pi$d scattering length
224: $a_{\pi d}$.
225: The latter quantity is defined as the elastic $\pi$d scattering
226: amplitude evaluated at zero kinetic energy of the incident pion.
227: This amplitude is necessarily complex because absorption reaction channels
228: are open even at the very threshold. The most important of them is the
229: $\pi^{-}d\to nn$ reaction, and to a lesser extend the radiative absorption
230: $\pi^{-}d\to \gamma nn$ channel. In principle, there would be also
231: the charge-exchange break-up
232: channel $\pi^{-}d\to\pi^{0} nn$ that is open at threshold
233: but this process is strongly suppressed by the centrifugal barrier.
234: Indeed, with s-wave $\pi$N interaction there is no spin-flip possible so
235: that for the two neutrons the $\mbox{}^{1}S_{0}$ state is not available,
236: whereas the $\mbox{}^{3}S_{1}$ state is forbidden
237: and they have to be produced in higher partial waves.
238: On the whole, however, the absorptive effects are not large at
239: threshold, judging from the magnitude of the imaginary part of
240: the $\pi$d scattering length which empirically
241: constitutes only about a quarter of
242: the real part of $a_{\pi d}$. Strictly speaking, the absorptive
243: processes contribute to both, the real and the imaginary part
244: of $a_{\pi d}$ but in the following we are going to ignore the
245: absorptive corrections to the real part of $a_{\pi d}$.
246: Disregarding the absorptive processes, we shall concentrate our
247: attention on a microscopic calculation of $a_{\pi d}$ and in
248: order to be able to solve the ensuing three-body problem
249: we introduce a potential description of the $\pi$N interaction to be
250: used in the appropriate Faddeev equations.
251: \par
252: In order to facilitate the discussion of the Faddeev approach,
253: it is instructive to take the static model as our point of
254: departure. The attractive feature of the static model is that it
255: is much easier to develop and to compute since the final result
256: for pion-deuteron scattering length takes the form of a single analytic
257: formula that does not require off-shell information. Moreover,
258: in our case the latter model also happens to be extremely good
259: approximation to the full solution of the three-body problem.
260: The earliest version of a static model, due to Brueckner \cite{BRUCK},
261: was based on the fixed scatterer concept and ignored
262: all isospin complications. Here, we wish to make it
263: somewhat more realistic introducing as our dynamical framework
264: a set of appropriate boundary conditions, but
265: on the other hand, we are prepared
266: to content ourselves with a theory that has isospin invariant
267: point like interactions.
268: Labeling the pion as $1$ and the nucleons as $2$ and $3$,
269: the boundary conditions representing the
270: zero-range $\pi$-N interaction
271: taking place on nucleon $i$ where $i=2,3$, may be written as
272: \begin{equation}
273: \lim_{x_{1} \to x_{i}}
274: \overline{ |\bbox{x}_{1}-\bbox{x}_{i}| \Psi (\bbox{x}_{1},\bbox{x}_{2},
275: \bbox{x}_{3})}=
276: (\mu/m)(b_{0}+b_{1}\,\bbox{I}\cdot \bbox{\tau}_{i})
277: \lim_{x_{1} \to x_{i}} \dfrac{\D}{\D x_{1}}\;
278: \overline{ |\bbox{x}_{1}-\bbox{x}_{i}| \Psi (\bbox{x}_{1},\bbox{x}_{2},
279: \bbox{x}_{3}) }
280: \label{a1}
281: \end{equation}
282: where the bar denotes an average over directions
283: $\bbox{x}_{1}-\bbox{x}_{i}$ what is equivalent to
284: projecting out the s-wave component of the wave function $\Psi$,
285: and the boundary condition \Ref{a1} is to be imposed for each of the
286: two nucleons. The vectors $\bbox{I}$ and $\bbox{\tau}$ are, respectively,
287: the pion and the nucleon isospin operators, whereas $b_{0}$ and $b_{1}$
288: denote the isoscalar and isovector $\pi$-N scattering lengths,
289: $\mu$ is the $\pi$-N reduced mass and $m$ is the pion mass.
290: In the following we choose the c.m. of the two nucleons as the origin
291: of the coordinate system, i.e. we set $\bbox{x}_{2}=\Half\bbox{r}$
292: and $\bbox{x}_{3}=-\Half\bbox{r}$ with $\bbox{r}$ being the
293: nucleon-nucleon separation vector. The pion vector in this
294: Jacobi coordinate system will be denoted as $\bbox{\rho}$.
295: When the wave function $\Psi(\bbox{r},\bbox{\rho})$ describing
296: the $\pi$NN system for the case of $\pi^{-}$ scattered off the
297: deuteron is known, the amplitude leading to the final state with
298: asymptotic wave function $\Phi_{f}$ is $-\bra\Phi_{f}|V|\Psi\ket$
299: where $V$ denotes the potentials that have been taken out
300: in the derivation of $\Phi_{f}$. For elastic scattering
301: $\Phi_{f} ( \bbox{\rho} , \bbox{r} )=
302: \exp{ (\imath \bbox{p}' \cdot \bbox{\rho} ) } \, \psi_{d}(\bbox{r}) $
303: where $\bbox{p}'$ is the momentum of the
304: outgoing pion, $\psi_{d}$ is the deuteron wave function and
305: $V$ is the sum of the two $\pi$N potentials as asymptotically there is
306: no $\pi$-deuteron interaction. Although in our formalism we never
307: needed $\pi$N potentials and the $\pi$N interaction is represented by
308: the boundary condition \Ref{a1}, it is in fact possible to give a
309: formal expression for such potential (cf. \cite{HUANG})
310: and for the operator $V$ we take
311: \begin{equation}
312: V\Psi(\bbox{\rho},\bbox{r})=-\frac{2\pi}{\mu}
313: \left \{
314: (b_{0}+b_{1}\,\bbox{I}\cdot\bbox{\tau}_{2})
315: \,\delta(\bbox{\rho}-\Half\bbox{r}) \dfrac{\D}{\D\rho}
316: |\bbox{\rho}-\Half\bbox{r}|+
317: (b_{0}+b_{1}\,\bbox{I}\cdot\bbox{\tau}_{3})
318: \,\delta(\bbox{\rho}+\Half\bbox{r}) \dfrac{\D}{\D\rho}
319: |\bbox{\rho}+\Half\bbox{r}|\right\}\Psi(\bbox{\rho},\bbox{r}).
320: \label{a2}
321: \end{equation}
322: Denoting the incident pion momentum as
323: $\bbox{p}$ and
324: making use of the boundary conditions \Ref{a1} in \Ref{a2},
325: the $\pi$-d elastic scattering amplitude
326: $f(\bbox{p}',\bbox{p})$ takes the form
327: \begin{equation}
328: f(\bbox{p}^{\prime},\bbox{p})
329: =\frac{\nu}{m} \int \E^{-\imath\,\bbox{p}' \cdot\bbox{\rho}}\;
330: \psi_{d}^{\dagger}(\bbox{r}) \; \left \{
331: \delta(\bbox{\rho}-\Half\bbox{r})
332: |\bbox{\rho}-\Half\bbox{r}|+
333: \delta(\bbox{\rho}+\Half\bbox{r})
334: |\bbox{\rho}+\Half\bbox{r}|\right\}\Psi(\bbox{\rho},\bbox{r})
335: \D^{3}\rho\;\D^{3}r,
336: \label{a3}
337: \end{equation}
338: where $\nu$ is $\pi$-d reduced mass.
339: Given the elastic $\pi$-d scattering amplitude \Ref{a3},
340: the $\pi$-d scattering length follows immediately from
341: \begin{equation}
342: a_{\pi d}=f(0,0).
343: \label{a4}
344: \end{equation}
345: \par
346: With the $\pi$-N interaction assumed to be isospin invariant, it will
347: be convenient for us to adopt an isospin notation.
348: For the initial $\pi^{-}$-d
349: system, the isotopic spin wave function has the form
350: \begin{equation}
351: \chi_{a}=\pi^{-}\tfrac{1}{\sqrt{2}}(p_{2}n_{3}-n_{2}p_{3}),
352: \label{a5}
353: \end{equation}
354: where the symbols $p, n, \pi^{-}$ in \Ref{a5}
355: stand for the isospin wave functions of
356: the corresponding particles. The wave function \Ref{a5} is
357: antisymmetric in the nucleon labels, as appropriate for the
358: state where the isospin of the two-nucleon subsystem
359: $I_{23}$ equals zero.
360: As a result of the interaction, the two nucleons
361: can undergo a transition to
362: a symmetric configuration corresponding to $I_{23}=1$ and we shall
363: need also a function that is symmetric under two-nucleon permutation
364: \begin{equation}
365: \chi_{s}= \tfrac{1}{2} \pi^{-} (p_{2}n_{3}+n_{2}p_{3})-
366: \tfrac{1}{\sqrt{2}} \pi^{0}n_{2}n_{3}.
367: \label{a6}
368: \end{equation}
369: Since our interest here is confined to s-wave interactions,
370: no spin flip is possible and therefore
371: the spin part of the wave function does not change.
372: Regarding the nucleons as fixed scattering centers,
373: we may anticipate that the wave function
374: $\Psi(\bbox{\rho},\bbox{r})$ for the full system of the target
375: nucleons and the meson will take the approximate form
376: %
377: \begin{eqnarray}
378: \Psi(\bbox{\rho},\bbox{r})=
379: e^{\imath\bbox{p}\cdot\bbox{\rho}}\,
380: u_{d}(r)\, \chi_{a}
381: + A(\bbox{r}) \left [ \dfrac{
382: \exp{(\imath p |\bbox{\rho}-\Half\bbox{r}|)}}
383: {|\bbox{\rho}-\Half\bbox{r}|}
384: + \dfrac{
385: \exp{(\imath p |\bbox{\rho}+\Half\bbox{r}|)}}
386: {|\bbox{\rho}+\Half\bbox{r}|} \right ] \chi_{a} +
387: % \hspace{1cm}
388: \nonumber \\
389: \mbox{} + X(\bbox{r}) \left [
390: \dfrac{\exp{(\imath p |\bbox{\rho}-\Half\bbox{r}|)}}
391: {|\bbox{\rho}-\Half\bbox{r}|}
392: - \dfrac{\exp{(\imath p |\bbox{\rho}+\Half\bbox{r}|)}}
393: {|\bbox{\rho}+\Half\bbox{r}|}
394: \right ] \chi_{s},
395: \label{a7}
396: \hspace{1cm}
397: \end{eqnarray}
398: %
399: where $u_{d}$ is the spatial part of deuteron wave function that
400: includes also the deuteron spin and in particular
401: may contain also the D-component. The projectile enters with
402: momentum $\bbox{p}$ and in the initial asymptotic region the pion and
403: the target have separate wave functions (a plane wave and $u_{d}(r)$,
404: respectively) and the propagation from one scattering center to
405: another is described by a superposition of spherical waves.
406: The hitherto unknown amplitudes
407: denoted in \Ref{a7}, respectively, as $A(\bbox{r})$ and $X(\bbox{r})$
408: multiplying these outgoing waves emitted by the two centers
409: account for the multiple scattering phenomena. They will
410: be determined from the boundary conditions \Ref{a1}.
411: To satisfy Pauli principle the wave function \Ref{a7}
412: must be antisymmetric in the two nucleon variables.
413: This implies that we have to stipulate that
414: the coefficients $A(\bbox{r})$ and $X(\bbox{r})$ are even
415: under permutation of the nucleons, i.e. they
416: must be invariant under the reflections $\bbox{r}\to -\bbox{r}$.
417: For zero-energy scattering considered in this work,
418: however, this is always the case because
419: $A(r)$ and $X(r)$ depend then only upon the magnitude of
420: $\bbox{r}$. It is worth noting that the wave function \Ref{a7}
421: includes explicitly virtual charge exchange amplitude $X(r)$.
422: Since our interest here is confined to zero-energy scattering,
423: in the following we take $p=0$ in \Ref{a7}.
424: Equations for the functions $A(r)$ and $X(r)$ may be obtained
425: by substituting \Ref{a7} in \Ref{a1} for $i=2$ and equating the
426: coefficients multiplying the same isospin functions. With
427: two different isospin functions we obtain two equations
428: and this procedure determines uniquely $A(r)$ and $X(r)$.
429: Owing to the proper antisymmetrization of our wave function the
430: boundary condition for $i=3$ will be then automatically satisfied.
431: The equations obtained from \Ref{a1} are
432: \begin{subequations}
433: \label{a8}
434: \begin{eqnarray}
435: A(r)&=&\tilde{b}_{0}u_{d}(r) +(\tilde{b}_{0}/r)\, A(r)+
436: \sqrt{2}(\tilde{b}_{1}/r)\, X(r),
437: \label{a8:a}
438: \\
439: -X(r)&= &\sqrt{2}\tilde{b}_{1}u_{d}(r) +\sqrt{2}
440: (\tilde{b}_{1}/r)\,A(r)
441: +(\tilde{b}_{0}+\tilde{b}_{1})/r \,X(r).
442: \label{a8:b}
443: \end{eqnarray}
444: \end{subequations}
445: In \Ref{a8} we introduced the abbreviation $\tilde{b}_{j}=(1+m/M)b_{j}$
446: where $M$ is the nucleon mass.
447: The $\pi$-d scattering length is given by the overlap integral
448: \begin{equation}
449: a_{\pi d}=(2\nu/m)\,\int u_{d}(r)^{\dagger}A(r) \, \D^{3}r
450: \label{a9}
451: \end{equation}
452: where $A(r)$ is the solution of \Ref{a8}
453: \begin{equation}
454: A(r)= \frac{ \tilde{b}_{0}+ (\tilde{b}_{0}+\tilde{b}_{1})
455: (\tilde{b}_{0}-2\tilde{b}_{1})/r }
456: {1-\tilde{b}_{1}/r - (\tilde{b}_{0}+\tilde{b}_{1})
457: (\tilde{b}_{0}-2\tilde{b}_{1})/r^{2} } \; u_{d}(r).
458: \label{a10}
459: \end{equation}
460: Using \Ref{a10} in \Ref{a9} and
461: expanding $A(r)$ in powers of the $\pi$N scattering lengths,
462: we retrieve the well known
463: second order formula for the $\pi$-d scattering length
464: (cf. \cite{ERI88})
465: \begin{equation}
466: a_{\pi d}^{(2)}=\frac{2\nu}{m}\left [
467: \tilde{b}_{0}+(\tilde{b}_{0}^{2}-2\,\tilde{b}_{1}^{2})
468: \bra \frac{1}{r} \ket \right ],
469: \label{a12}
470: \end{equation}
471: where the expectation value is taken with respect to the
472: deuteron wave function. As advertised at the beginning of this section,
473: formula \Ref{a10} provides a complete solution of the problem.
474: To examine the accuracy of the static formula we have to compare it
475: with the exact solution of the three-body problem. The latter will be
476: obtained by solving the Faddeev equations on which we now embark.
477: \par
478: To solve the Faddeev equations it will be convenient for us to work in
479: momentum space. Introducing the Faddeev partitions, we
480: write the three-body wave function as
481: \begin{equation}
482: \Psi= \psi^{(1)}(\bbox{q}_{1},\bbox{k}_{1})
483: + \psi^{(2)}(\bbox{q}_{2},\bbox{k}_{2})
484: + \psi^{(3)}(\bbox{q}_{3},\bbox{k}_{3}),
485: \label{a15}
486: \end{equation}
487: where $\bbox{q}_{1}$ denotes the relative momentum of the (23) pair
488: whereas $\bbox{k}_{1}$ is the c.m. momentum of particle 1
489: and cyclic permutations are implied. To obtain Faddeev equations for
490: the amplitudes, the different partitions are written as (cf.
491: \cite{PETROV})
492: \begin{subequations}
493: \label{a16}
494: \begin{eqnarray}
495: \psi^{(1)}(\bbox{q},\bbox{k})&=&(2\pi)^{3}\phi(\bbox{q})
496: \delta(\bbox{k}-\bbox{p}) \chi_{a}
497: +[F(\bbox{q},\bbox{k}) \chi_{a}+G(\bbox{q},\bbox{k})\chi_{s}]
498: /(E-q^{2}/M-k^{2}/2\nu);
499: \label{a16:a}
500: \\
501: \psi^{(2)}(\bbox{q},\bbox{k})&=&
502: [A(-\bbox{q},\bbox{k})\chi_{a} - X(-\bbox{q},\bbox{k})\chi_{s}]
503: /(E-q^{2}/2\mu-k^{2}/2\nu_{N});
504: \label{a16:b}
505: \\
506: \psi^{(3)}(\bbox{q},\bbox{k})&=&
507: [A(\bbox{q},\bbox{k})\chi_{a} + X(\bbox{q},\bbox{k})\chi_{s}]
508: /(E-q^{2}/2\mu-k^{2}/2\nu_{N});
509: \label{a16:c}
510: \end{eqnarray}
511: \end{subequations}
512: where $\nu_{N}$ is the reduced mass of the nucleon and that of
513: the $\pi$N pair, $E$ is the c.m. three particle kinetic
514: energy and $\phi(\bbox{q})$ is the
515: deuteron wave function in the momentum space.
516: In \Ref{a16} we have introduced four scattering amplitudes
517: $F(\bbox{q},\bbox{k}), G(\bbox{q},\bbox{k}), A(\bbox{q},\bbox{k})$ and
518: $X(\bbox{q},\bbox{k})$. However, the amplitude $G(\bbox{q},\bbox{k})$
519: to be non-zero requires at least p-wave NN interaction and
520: therefore will be excluded from our considerations, while
521: the three remaining amplitudes
522: will be determined from the Faddeev equations.
523: It is evident from \Ref{a16} that under the $P_{23}$ permutation
524: $\psi^{(1)} \to -\psi^{(1)}$ and
525: $\psi^{(2)} \leftrightarrow -\psi^{(3)}$, so that
526: the total wave function is,
527: as required, antisymmetric in the nucleon labels.
528: Assuming exact isospin conservation, we can write the Faddeev
529: equations
530: %%%% Faddeev %%%%
531: \begin{subequations}
532: \label{a17}
533: \begin{eqnarray}
534: \lefteqn{}
535: F( \bbox{q}, \bbox{k} ) =
536: \int \frac{\D^{3}\,k^{\prime}}{(2\pi)^{3}}
537: \,\frac{
538: \bra \bbox{q}| t(E-k^{2}/2\nu) | \Half\bbox{k}+\bbox{k}^{\prime}\ket+
539: \bra \bbox{q}| t(E-k^{2}/2\nu) | -\Half\bbox{k}-\bbox{k}^{\prime}\ket}
540: {E-(\bbox{k}+\mu\bbox{k}^{\prime}/M)^{2}/2\mu-k^{\prime 2}/2\nu_{N}}
541: % \times \nonumber \\ && \hspace{-5cm} \times
542: \; A(\bbox{k}+\bbox{k}^{\prime}\frac{\mu}{M},\bbox{k}^{\prime});
543: \label{a17:a}
544: \end{eqnarray}
545: %%%
546: \begin{eqnarray}
547: \lefteqn{}
548: &&
549: A(\bbox{q},\bbox{k})=
550: \bra \bbox{q}
551: \left | t_{0}(E-\frac{k^{2}}{2\nu_{N}}) \right |\frac{\mu}{M} \bbox{k}
552: +\bbox{p} \ket \, \phi(\bbox{k}+\Half \bbox{p})+
553: \nonumber
554: \\
555: && \hspace{1cm}
556: + \int \frac{\D^{3}\,k^{\prime}}{ (2\pi)^{3} }
557: \,\frac{\bra \bbox{q}| t_{0}(E-k^{2}/2\nu_{N}) |\mu
558: \bbox{k}/M+\bbox{k}^{\prime}\ket }
559: { E-(\bbox{k}+\Half\bbox{k}^{\prime})^{2}/M-k^{\prime 2}/2\nu }
560: \; F(-\bbox{k}-\Half\bbox{k^{\prime}},\bbox{k^{\prime}})+
561: \nonumber
562: \\
563: && \hspace{1cm}
564: + \int \frac{\D^{3}\,k^{\prime}}{ (2\pi)^{3} }
565: \,\frac{ \bra \bbox{q}| t_{0}(E-k^{2}/2\nu_{N})
566: |-\mu\bbox{k}/m-\bbox{k^{\prime}}\ket}
567: { E-(\bbox{k}+\mu\bbox{k^{\prime}}/m)^{2}/2\mu-k^{\prime 2}/2\nu_{N} }
568: \; A(-\bbox{k}-\frac{\mu}{m}\bbox{k^{\prime}},\bbox{k^{\prime}})+
569: \nonumber
570: \\
571: && \hspace{1cm}
572: +\sqrt{2} \int \frac{\D^{3}\,k^{\prime}}{(2\pi)^{3}}
573: \,\frac{ \bra \bbox{q}| t_{1}(E-k^{2}/2\nu_{N})
574: |-\mu\bbox{k}/m-\bbox{k^{\prime}}\ket }
575: { E-(\bbox{k}+\mu\bbox{k^{\prime}}/m)^{2}/2\mu-k^{\prime 2}/2\nu_{N} }
576: \;X(-\bbox{k}-\frac{\mu}{m}\bbox{k^{\prime}},\bbox{k^{\prime}});
577: \label{a17:b}
578: \end{eqnarray}
579: %%%
580: \begin{eqnarray}
581: \lefteqn{}
582: &&
583: -X(\bbox{k},\bbox{k})=\sqrt{2}
584: \bra \bbox{k}
585: \left |t_{1}(E-\frac{k^{2}}{2\nu_{N}}) \right |\frac{\mu}{M}\bbox{k}
586: +\bbox{p} \ket \phi(\bbox{k}+\Half\bbox{p})+
587: \nonumber
588: \\
589: && % \hspace{1cm}
590: +\sqrt{2} \int \frac{\D^{3}\,k^{\prime}}{(2\pi)^{3}}
591: \,\frac{\bra \bbox{k}| t_{1}(E-k^{2}/2\nu_{N})
592: |\mu\bbox{k}/M+\bbox{k^{\prime}}\ket }
593: {E-(\bbox{k}+\Half\bbox{k^{\prime}})^{2}/M-k^{\prime 2}/2\nu}
594: \; F(-\bbox{k}-\Half\bbox{k^{\prime}},\bbox{k^{\prime}})+
595: \nonumber
596: \\
597: && %\hspace{1cm}
598: + \int \frac{\D^{3}\,k^{\prime}}{(2\pi)^{3}}
599: \,\frac{\bra \bbox{k}| [t_{0}(E-k^{2}/2\nu_{N})
600: - t_{1}(E-k^{2}/2\nu_{N})] |-\mu\bbox{k}/m-\bbox{k^{\prime}}\ket}
601: { E-(\bbox{k}+\mu\bbox{k^{\prime}}/m)^{2}/2\mu-k^{\prime 2}/2\nu_{N} }
602: \; X(-\bbox{k}-\frac{\mu}{m}\bbox{k^{\prime}},\bbox{k^{\prime}})+
603: \nonumber
604: \\
605: && %\hspace{1cm}
606: +\sqrt{2} \int \frac{\D^{3}\,k^{\prime}}{(2\pi)^{3}}
607: \,\frac{ \bra \bbox{k}| t_{1}(E-k^{2}/2\nu_{N})
608: |-\mu\bbox{k}/m-\bbox{k^{\prime}}\ket }
609: { E-(\bbox{k}+\mu\bbox{k^{\prime}}/m)^{2}/2\mu-k^{\prime 2}/2\nu_{N} }
610: \; A(-\bbox{k}-\frac{\mu}{m}\bbox{k^{\prime}},\bbox{k^{\prime}});
611: \label{a17:c}
612: \end{eqnarray}
613: \end{subequations}
614: where in \Ref{a17} $\bra \bbox{q}'|t(E)| \bbox{q} \ket $
615: is the NN scattering t-matrix for zero isospin and
616: $\bra \bbox{q}'|t_{j}(E)| \bbox{q} \ket $ are, respectively,
617: the isoscalar $(j=0)$ and isovector $(j=1)$ $\pi$N scattering t-matrices.
618: The elastic scattering amplitude is given by the expression
619: \begin{equation}
620: f(\bbox{p}',\bbox{p})=\lim_{p'\to p}\frac{{p'}^{2}-p^{2}}{4\pi}
621: \int \phi(\bbox{q})^{\dagger}\frac{F(\bbox{q},\bbox{p'})}
622: {E-q^{2}/M-{p'}^{2}/2\nu}\,\frac{\D^{3}q}{(2\pi)^{3}}
623: \label{a20}
624: \end{equation}
625: and the scattering length is obtained from \Ref{a4}. We can use
626: \Ref{a17:a} to eliminate $F(\bbox{q},\bbox{k})$ in \Ref{a20} in favour of
627: the amplitude $A(\bbox{q},\bbox{k})$. In the NN scattering matrices
628: occurring in \Ref{a17}, as a result of the limiting
629: procedure, only the deuteron pole contributes and
630: scattering length is given as an overlap integral
631: \begin{equation}
632: a_{\pi d}= -\frac{\nu}{\pi} \int \phi(k)^{\dagger}
633: A(\bbox{k}\,\frac{\mu}{M}, \bbox{k})\,\frac{\D^{3}k}{(2\pi)^{3}}.
634: \label{a21}
635: \end{equation}
636: The above formula is analogous to \Ref{a9}, and, in fact, the static
637: approximation results \Ref{a9}-\Ref{a10} could have been derived from
638: the Faddeev formalism.
639: In order to demonstrate that \Ref{a9}-\Ref{a10}
640: follow from \Ref{a17}
641: we note that when the nucleons are static
642: they are not supposed to scatter ($t \to 0 $) and the amplitude
643: $F(\bbox{q},\bbox{k})$ drops out in \Ref{a17:b} and \Ref{a17:c}
644: so that we are left with only two coupled integral equations.
645: When the underlying forces are of zero range,
646: the off-shell $\pi$N scattering amplitudes can be simplified,
647: and in that case
648: $$
649: \bra \bbox{q}'|t_{j}(E)|\bbox{q} \ket=
650: -(2\pi/\mu)\,b_{j}/(1+\kappa b_{j}),\qquad j=0,1;
651: $$
652: where $ \kappa^{2}=2\mu B $ and $B$ is the binding energy
653: of the deuteron. The important consequence of the zero-range
654: assumption, apparent from the above formula, is that
655: the t-matrices become
656: independent upon the off-shell momenta.
657: Therefore, the amplitudes $A(\bbox{q},\bbox{k})$ and
658: $X(\bbox{q},\bbox{k})$ will be functions of one variable only and it
659: will be convenient for us to introduce a notation that emphasizes that
660: fact, setting $A(\bbox{q},\bbox{k})=-(m/2\pi){\cal A}(k)$ and
661: $X(\bbox{q},\bbox{k})=-(m/2\pi){\cal X}(k)$, where ${\cal A}(k)$ and ${\cal
662: X}(k)$ are two, hitherto unknown amplitudes. With static nucleons,
663: the energy denominators in
664: \Ref{a17:b} and \Ref{a17:c} become all equal to
665: $-B-(\bbox{k}'+\bbox{k})^{2}/2m$
666: and we end up with the following set of integral equations
667: for the amplitudes ${\cal A}(k)$ and ${\cal}X(k)$:
668: \begin{subequations}
669: \label{a23}
670: \begin{equation}
671: {\cal A}(k)= \hat{b}_{0} \phi(k) +
672: 4\pi \;\hat{b}_{0}
673: \int \frac{\D^{3} k'}
674: {(2\pi)^{3}} \frac{{\cal A}(k')}{\kappa^{2}+(\bbox{k}'+\bbox{k})^{2}}
675: +\sqrt{2} \; 4\pi \hat{b}_{1}
676: \int \frac{\D^{3} k'}
677: {(2\pi)^{3}} \frac{{\cal X}(k')}{\kappa^{2}+(\bbox{k}'+\bbox{k})^{2}};
678: \label{a23:a}
679: \end{equation}
680: \begin{eqnarray}
681: -{\cal X}(k)= \sqrt{2}\hat{b}_{1} \phi(k)
682: +4\pi
683: ( \hat{b}_{0}-\hat{b}_{1} )
684: && \int \frac{\D^{3} k'} {(2\pi)^{3}}
685: \frac{{\cal X}(k')}{\kappa^{2}+(\bbox{k}'+\bbox{k})^{2}}+
686: %\hspace{4cm} \nonumber \\ \mbox{}
687: + \sqrt{2}\;4\pi \hat{b}_{1}
688: % &&
689: \int \frac{\D^{3} k'} {(2\pi)^{3}}
690: \frac{{\cal A}(k')}{\kappa^{2}+(\bbox{k}'+\bbox{k})^{2}},
691: \label{a23:b}
692: \end{eqnarray}
693: \end{subequations}
694: where
695: \begin{equation}
696: \hat{b}_{j}=b_{j}\,(1+m/M)/(1+\kappa b_{j});\qquad j=0,1.
697: \label{a24}
698: \end{equation}
699:
700: The above set of integral equations can be immediately
701: solved by introducing the Fourier transform
702: \begin{equation}
703: A(r)=
704: \int e^{\imath \bbox{k}\bbox{r}} {\cal A}(k) \D^{3}k
705: \label{a25}
706: \end{equation}
707: together with a similar relationship for ${\cal X}(k)$ and $\phi (k)$
708: and using the well known formula
709: $$
710: \frac{4\pi}{\kappa^{2}+(\bbox{k}+\bbox{k}')^{2}}=\int
711: e^{-\imath (\bbox{k}+\bbox{k}')\bbox{r} }\;
712: \frac{e^{ -\kappa r}}{r} \D^{3} r.
713: $$
714: In order to solve \Ref{a23} we multiply the latter
715: equations by $e^{\imath \bbox{k}\bbox{r}}$ and subsequently integrate
716: them over $\bbox{k}$.
717: As a result, we obtain a set of two algebraic equations
718: for $A(r)$ and $X(r)$ that differ from
719: \Ref{a8} only by $\exp{(-\kappa r)}/r$ replacing $1/r$
720: and $\hat{b}_{j}$
721: replacing $\tilde{b}_{j}$.
722: Since \Ref{a21} goes over into \Ref{a9}, we are led
723: to the extension of the static formula \Ref{a10}
724: \begin{equation}
725: A(r)= \frac{ \hat{b}_{0}+ (\hat{b}_{0}+\hat{b}_{1})
726: (\hat{b}_{0}-2\hat{b}_{1})e^{-\kappa r} /r }
727: {1-\hat{b}_{1} e^{-\kappa r}/r - (\hat{b}_{0}+\hat{b}_{1})
728: (\hat{b}_{0}-2\hat{b}_{1})e^{-2\kappa r}/r^{2} } \; u_{d}(r).
729: \label{a26}
730: \end{equation}
731: This formula is to be used in \Ref{a9} but now accounts for
732: the binding energy correction.
733: %-----------------------
734: \par
735: Concluding our discussion of the static model
736: we wish to recall that a closed form expression for $\pi$d
737: scattering length has been also obtained by effecting
738: an explicit summation of Feynman diagrams and the
739: most complete treatment can be found in Ref. \cite{Victor}.
740: The ultimate static formula for $a_{\pi d}$,
741: that takes into account isospin degree
742: of freedom, given in \cite{Victor}
743: is rather complicated and at first sight
744: appears to be different from \Ref{a26}. However, a closer inspection
745: reveals that the authors of Ref. \cite{Victor}
746: apparently did not realize that their
747: fractional formula for $a_{\pi d}$ could have been significantly
748: simplified because a common factor equal
749: $$
750: 1+\tilde{b}_{1}e^{-\kappa r}/r-(\tilde{b}_{0}+\tilde{b}_{1})
751: (\tilde{b}_{0}-2\tilde{b}_{1})e^{-2\kappa r}/r^{2}
752: $$
753: may be pulled out both, from the numerator, as well as from the
754: denominator and eventually drops out. Indeed,
755: when the redundant factor has been cancelled, the resulting expression
756: is identical with \Ref{a26}.
757: Therefore, when binding corrections are disregarded, this
758: approach reproduces the static model result \Ref{a10} and it is
759: reassuring that in this case all three methods give the same answer.
760: \par
761: To improve upon the static model one needs a numerical solution of
762: the Faddeev equations and in the following,
763: similarly as in the previous calculations \cite{PETROV,AFN74,MIZ77},
764: in order to reduce the computational effort,
765: all the pairwise interactions invoked will be represented
766: by rank-one separable potentials. The $\pi$N s-wave interaction
767: is taken in the form of a standard Yamaguchi potential with the same
768: form factor in both isospin states. Since the inverse range parameter
769: $\beta$ that enters that form factor is not known, similarly as
770: before, we consider
771: the zero-range limit, i.e. $\beta \to \infty$.
772: For an assigned value of $\beta$, the strength parameter of the
773: potential may be eliminated in favour of the scattering length and
774: the appropriate s-wave t-matrices, are
775: \begin{equation}
776: \bra k|t_{j}(E)|k' \ket=-\frac{2\pi}{\mu}
777: \frac{1}{1+k^{2}/\beta^{2}}\;
778: \frac {b_{j}}
779: {1-\imath p b_{j} \, (1-2\imath p/\beta)(1-\imath p/\beta)^{-2}}
780: \; \frac{1}{1+{k'}^{2}/\beta^{2}}
781: \label{a27}
782: \end{equation}
783: where $p=\sqrt{2\mu E}$ and $j=0,1$ and it is evident from \Ref{a27}
784: that the zero-range limit can be effected.
785: When the nucleon motion is taken into account, the p-wave $\pi$N
786: interaction gives contribution to the $\pi$d scattering amplitude
787: even at threshold. Therefore,
788: in addition to s-wave, we are going to include
789: also the p-wave interaction, limiting ourselves only to the P33 wave
790: as in that case both the strength and the statistical weight are dominant
791: rendering the remaining p-waves negligible.
792: The corresponding p-wave form factor of the form
793: $$
794: g_{\Delta}(k)=k/(k^{2}+\beta_{\Delta}^{2})
795: $$
796: has been adopted from \cite{AFN74}
797: with $\beta_{\Delta}=5.33\, fm^{-1}$
798: where the depth of the separable potential can be adjusted to the
799: experimentally known value of the P33 scattering volume
800: taken to be $0.64\,fm^{3}$. It is well known that with the above
801: form, the shape of the delta resonance cannot be well reproduced
802: but this is less important here, the essential thing is to have the P33
803: amplitude at threshold correctly reproduced.
804: Besides, the p-wave constitutes
805: only a small correction and using a more complicated model
806: does not seem to be currently justified.
807: For the NN interaction we use two separable models: a simple
808: Hulthen-Yamaguchi potential with inverse range parameter equal
809: $\beta_{N}=6.01162\, \sqrt{MB} $ whose strength is fixed by the
810: deuteron binding energy, and the PEST potential
811: constructed in Ref. \cite{PEST}
812: with a more sophisticated form factor of the form
813: \begin{equation}
814: g(k)=\sum_{i=1}^{6}\frac{C_{i}}{k^{2}+\beta_{i}^{2}},
815: \label{a28}
816: \end{equation}
817: where the parameters $C_{i}$ and $\beta_{i}$ have been tabulated in
818: Ref. \cite{PEST}.
819: This potential has been devised in such a way that the
820: corresponding NN half-off-shell T-matrix has the same behaviour as
821: that of the Paris potential \cite{Paris}. This separable replica
822: of the Paris potential takes into account the
823: short range repulsion that is absent in the Yamaguchi potential
824: yet retaining the simplicity of the latter.
825: \par
826: Using standard partial wave projections
827: the Faddeev equations \Ref{a17} can be reduced to
828: a system of four coupled inhomogeneous integral equations
829: in a single variable that
830: are amenable for numerical treatment.
831: In the actual practice, in order to cross-check our
832: numerical procedures, we used two independent methods of solving
833: these equations. The direct method introduces an integration mesh
834: what allows us to replace integrals by sums so that the integral
835: equations take the form of a system of linear algebraic equations
836: easily solvable by standard methods. The second method solves
837: the system of integral equations by successive iterations.
838: The iterative procedure
839: is equivalent to a power expansion in $\pi$N scattering lengths what
840: allows tracing down the contribution from the different orders.
841: Since the scattering lengths are rather small, as compared with
842: the deuteron size,
843: the iterative sequence proves to be rapidly convergent.
844: \par
845: The experimental $\pi$-d scattering length
846: has been extracted form the 1s level
847: shift in pionic deuterium by using the Deser-Trueman \cite{Deser} formula.
848: Therefore, the extracted quantity is in fact the Coulomb corrected
849: scattering length, denoted hereafter as $a_{\pi d}^{c}$,
850: and before confronting the calculated pion-deuteron scattering length
851: with experiment one needs the experimental value of $a_{\pi d}$, i.e.
852: the purely nuclear scattering length.
853: Of course, Coulomb correction could be anticipated to be very small
854: but since the experimental errors are also small, it is of interest
855: to give some quantitative estimate of the Coulomb correction.
856: In principle, for calculating the latter one needs to know
857: the pion-deuteron nuclear potential responsible for the level shift.
858: This potential is not known but
859: with the zero-range potential simulating the
860: $\pi$N interaction in the first approximation it is reasonable to
861: expect that the effective potential is proportional to the nuclear density,
862: so that the shape of the nuclear potential is given by the square of
863: the deuteron wave function $u_{d}(2r)^{2}$.
864: Still, the depth is not known, but on the nuclear scale
865: this potential must be rather weak because the experimental value of
866: $a_{\pi d}^{c}$ is quite small.
867: Therefore, to quantify the value of
868: the ratio of $a_{\pi d}^{c}/a_{\pi d}$ it is sufficient to
869: keep only the first order terms in the nuclear potential.
870: Since in this case the potential depth drops out,
871: we are led to the formula
872: \begin{equation}
873: \frac{a_{\pi d}^{c}}{a_{\pi d}}=
874: \frac{\int_{0}^{\infty} u_{d}(2r)^{2}\,\phi_{0}(0,r)^{2}\, \D r}
875: {\int_{0}^{\infty} u_{d}(2r)^{2}\,r^{2} \, \D r},
876: \label{a13}
877: \end{equation}
878: where $\phi_{\ell}(k,r)$ denotes the regular Coulomb wave function
879: that for zero-momentum (k=0) and zero
880: orbital momentum ($\ell$=0), simplifies to the form
881: \begin{equation}
882: \phi_{0}(0,r) = r\;J_{1}(2\sqrt{2\nu \alpha r})/\sqrt{2\nu\alpha r}
883: \label{a14}
884: \end{equation}
885: where $\alpha$ is the fine structure
886: constant and $J_{1}(x)$ denotes the Bessel function.
887: Expanding \Ref{a13} in powers of $\alpha$,
888: we obtain quite adequate first
889: order formula $a_{\pi d}^{c}/a_{\pi d}=1-\alpha\nu \bra r \ket$
890: where the expectation value is with respect to the deuteron wave
891: function.
892: We have checked that for a variety of deuteron wave functions
893: the calculated ratio \Ref{a13} has ben very stable and its
894: numerical value is 0.985.
895: Using this number together with the experimental value of
896: $a_{\pi d}^{c}$
897: $$
898: a_{\pi d}^{c}= [-(2.61 \pm 0.05)+\imath\,(0.63 \pm
899: 0.07)]\times 10^{-2}/m_{\pi}
900: $$
901: taken from Ref. \cite{HAU98} where $m_{\pi}$ is the mass of
902: the charged pion, we deduce the value of
903: the purely nuclear $\pi$d scattering length
904: \begin{equation}
905: a_{\pi d}=(-2.65\pm 0.05)\times 10^{-2}/m_{\pi},
906: \label{a29}
907: \end{equation}
908: and hereafter the above number
909: will be referred to as the ''experimental'' $\pi$d scattering length
910: in which all absorptive effects have been neglected.
911: \par
912: Adopting the zero-range model of the $\pi$N interaction,
913: for calculating the $\pi$d scattering length one needs as input just
914: the isoscalar and the isovector $\pi$N scattering scattering lengths.
915: The values of $b_{0}$ and $b_{1}$ that have been extracted from
916: the pionic hydrogen data in Ref. \cite{SCH99}, are
917: \begin{equation}
918: b_{0}=-(0.22 \pm 0.43)\times 10^{-2}/m_{\pi};\qquad
919: b_{1}=-(9.05 \pm 0.42)\times 10^{-2}/m_{\pi},
920: \label{a30}
921: \end{equation}
922: where the quoted uncertainty comprises the experimental errors
923: together with the uncertainty introduced by applying a specific procedure
924: that allows to deduce $b_{0}$ and $b_{1}$ from the
925: measured x-ray spectra. The theoretical uncertainty is quoted to be about
926: twice as large as the experimental error.
927: Besides, the errors on $b_{0}$ and on $b_{1}$ are strongly correlated.
928: \par
929: Using \Ref{a30} as our input, we have calculated the $\pi$d scattering
930: length and the results are presented in Table \ref{table1}.
931: All entries are doubled because
932: we employ two models of NN interaction: the numbers without brackets
933: have been obtained using PEST wave function
934: and, respectively, the bracketed quantities
935: correspond to the Yamaguchi potential.
936: For each set of input values of $(b_{0}, b_{1})$ we computed $a_{\pi d}$
937: using five different methods discussed before,
938: beginning from the simplest second order
939: formula \Ref{a12}, through the static model \Ref{a10} and \Ref{a26},
940: up to the full Faddeev calculation without, and, with $\Delta$,
941: respectively. The results of the Faddeev calculation with s-wave
942: interaction only (without $\Delta$)
943: constitute a benchmark for the various approximations.
944: Contrary to what has been often claimed in the literature, the second
945: order formula is insufficient as the error incurred is roughly
946: four times bigger than the present experimental uncertainty on $a_{\pi d}$.
947: It is apparent from Table \ref{table1} that the closest to
948: Faddeev result is in all cases the static model \Ref{a10}.
949: The accuracy of the latter is very good, the error being
950: always below 2\%. By contrast,
951: the performance of the implementation \Ref{a26} of static model
952: is rather disappointing, especially that
953: from formula \Ref{a26} containing binding energy correction,
954: one might expect further improvement. Nevertheless, the numbers
955: show just the opposite, that in fact the included corrections
956: go in the wrong direction worsening the results so much
957: that even the second order formula proves to be more accurate.
958: Of course, it is not just the binding energy correction
959: that is responsible
960: for the difference between the static model and the Faddeev result,
961: as only the latter properly accounts for the nucleon recoil.
962: However, the lions share of the recoil correction seems to be
963: cancelled with the binding energy correction and
964: this cancellation explains the success of
965: the static formula \Ref{a10}
966: containing neither of these corrections. An explicit demonstration
967: that, at least to the second order,
968: such mechanism is at work can be found in Ref. \cite{FAL77}.
969: \par
970: Since the static model \Ref{a10} proves to be so accurate for
971: Yamaguchi and PEST models of the NN interaction, we took advantage
972: of this fact, using it to examine more realistic NN potentials containing
973: also the D-wave part. The results of our computations are displayed
974: in table \ref{table2} where we compare the two separable models
975: (Hulthen-Yamaguchi and PEST), used in Faddeev
976: calculations, with two popular local potentials
977: (Paris \cite{Paris} and Bonn \cite{Bonn}). As expected,
978: the PEST wave function results are indeed very
979: close to those obtained with Paris wave function despite the lack of
980: the D-component in the PEST wave function. Therefore,
981: neglecting the D-wave in the Faddeev calculation does not
982: appear to be a serious omission.
983: It is also gratifying that PEST, Paris and Bonn models
984: give very similar results.
985: \par
986: In table \ref{table3} we present the values of $\pi$d scattering
987: length obtained in result of iterative solution of the Faddeev
988: equations. Since for zero-range $\pi$N interaction there is no
989: additional suppression due to the $\pi$N form factor,
990: the rate of convergence is somewhat slower but the converged
991: result is obtained in less than 10 iterations.
992: We give $a_{\pi d}$
993: values calculated with and without p-wave $\pi$N interaction what
994: allows to evaluate the p-wave contribution in each order.
995: For Yamaguchi NN interaction the p-wave correction in the first order
996: is quite large and contributes $0.47 \times 10^{-2}/m_{\pi}$.
997: The p-wave contribution to the second order
998: (called sp-term in Ref. \cite{BARU})
999: has opposite sign and equals $-0.35\times 10^{-2}/m_{\pi}$.
1000: In general, the net effect of the p-wave interaction
1001: on the converged result is reduced owing to the
1002: destructive interference between repulsive s-waves and attractive
1003: p-waves, amounting in total only $0.29 \times 10^{-2}/m_{\pi}$.
1004: Similar features are observed for the PEST model but
1005: since the convergence rate is faster, the higher order corrections
1006: are suppressed and the
1007: interference effects seem to be smaller, i.e. the first order
1008: p-wave correction is $0.45\times 10^{-2}/m_{\pi}$ while the corresponding
1009: correction to the converged result is $0.39\times 10^{-2}/m_{\pi}$.
1010: \par
1011: It is apparent form table \ref{table1} that
1012: the calculated $\pi$d scattering length values are rather sensitive
1013: to the input values of $(b_{0}, b_{1})$ and therefore it is not
1014: so easy to see when the calculation agrees with experiment.
1015: To facilitate the comparison with experiment
1016: the values of $a_{\pi d}$
1017: resulting from Faddeev calculation (PEST with $\Delta$)
1018: and displayed in table \ref{table1} have been
1019: represented analytically using bilinear interpolation on a
1020: grid in the $(b_{0}, b_{1})$ plane.
1021: Then, given the interpolating polynomial, we equated it to
1022: the experimental value of $a_{\pi d}$, adding
1023: or subtracting the experimental error. This
1024: procedure gave us two constraints
1025: of algebraic form in the $(b_{0}, b_{1})$ variables,
1026: readily solvable with respect to one of these variables.
1027: The two functions obtained this way may be plotted
1028: in the $(b_{0}, b_{1})$ plane
1029: where, as shown in Fig. \ref{fig1}
1030: they set the boundary of the
1031: tilted band representing the one
1032: standard deviation constraint imposed by the $\pi d$
1033: scattering length deduced form pionic deuterium data.
1034: The rectangle in Fig. \ref{fig1} represents the experimental values of
1035: $(b_{0}, b_{1})$ to within one standard deviation inferred from
1036: pionic hydrogen data.
1037: The ultimate $(b_{0},b_{1})$ values that are consistent with both the
1038: pionic hydrogen and the pionic deuterium data fill the area of the
1039: black strip.
1040:
1041:
1042: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1043: \section{Finite range approach}%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1044: \label{se:three}
1045: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1046:
1047: Thus far our treatment of the pion-deuteron scattering problem
1048: has been carried out exclusively within the zero-range model.
1049: Although this model has served us well, it is based on certain
1050: idealization whose validity and consequences need to be examined.
1051: We therefore turn now to the question of formulating a finite range
1052: version of the approach presented in the preceding section.
1053: Relaxing the zero-range limitation has of course its
1054: {\it quid pro quo} in that we have to worry now about the off-shell
1055: extension of the $\pi$N scattering amplitude and this means that
1056: the pionic hydrogen problem has to be considered
1057: {\it ab intitio} in order to provide the
1058: necessary input for the $\pi$d calculation. Anticipating the
1059: application in the Faddeev type calculation, it will be
1060: convenient for us to work with separable potentials. To get
1061: insight into the pionic hydrogen problem,
1062: let us consider a two-channel situation, where
1063: the upper channel labeled as 1 corresponds to the neutral $\pi^{0}n$ system
1064: and the lower channel labeled as 2, respectively, to the
1065: $\pi^{-}p$ system. We assume that the two-channel interaction
1066: respects isospin invariance and the isospin symmetry is broken only by the
1067: Coulomb potential operative
1068: in channel 2 and by the mass splitting within
1069: isospin multiplets. Since we wish to consider an atomic system
1070: it is essential to treat the Coulomb interaction exactly.
1071: To meet this requirement, we choose the two-channel
1072: Lippmann-Schwinger equation as our dynamical framework that in
1073: coordinate representation takes the form
1074: \begin{subequations}
1075: \label{b0}
1076: \begin{equation}
1077: u_{1}(r)=\int_{0}^{\infty}
1078: \bra r \left | G_{1}^{+}(W) \right | r'\ket \,
1079: \left [ V_{11}(r',r'')\,u_{1}(r'') + V_{12}(r',r'')\,u_{2}(r'')\right ]
1080: \,\D r'\;\D r''
1081: \label{b0:a}
1082: \end{equation}
1083: \begin{equation}
1084: u_{2}(r)=\int_{0}^{\infty}
1085: \bra r \left | G_{2}^{+}(W) \right | r'\ket \,
1086: \left [ V_{21}(r',r'')\,u_{1}(r'') + V_{22}(r',r'')\,u_{2}(r'')\right ]
1087: \,\D r'\;\D r''
1088: \label{b0:b}
1089: \end{equation}
1090: \end{subequations}
1091: where we have assumed spherical symmetry of the problem and
1092: $u_{j}(r)$ denotes zero orbital momentum
1093: radial wave function in channel $j$ .
1094: The strong $\pi N$ interaction is adopted here in the form of
1095: a non-local potential matrix $V_{ij}$. In \Ref{b0} we have
1096: introduced the Green matrix whose only non-vanishing diagonal elements are
1097: \begin{equation}
1098: \bra r \left | G_{1}^{+}(W) \right | r'\ket=-(2\mu_{1} /p_{1})
1099: \exp{(\imath p_{1}r_{>})}\;\sin{( p_{1}r_{<})},
1100: \label{b2}
1101: \end{equation}
1102: for the neutral channel, while in the charged channel we have
1103: to take into account the Coulomb interaction and the
1104: exact Green's function in this case reads
1105: \begin{equation}
1106: \bra r \left | G_{2}^{+}(W) \right | r'\ket=-(2\mu_{2} /p_{2})
1107: \left [G_{0}(\eta,p_{2}\, r_{>})+\imath
1108: \, F_{0}(\eta,p_{2}\, r_{>})\right ] \; F_{0}(\eta, p_{2}\,r_{<}),
1109: \label{b3}
1110: \end{equation}
1111: where $r_{<}=\min(r,r'),\;r_{>}=\max(r,r')$.
1112: In \Ref{b0}-\Ref{b3} $W$ denotes the total c.m. energy
1113: (including the rest mass), $\mu_{j}$ are
1114: the reduced masses in the two channels and $p_{j}$ are the
1115: channel momenta: $ p_{j}=\pm \sqrt{ 2\mu_{j} (W-E_{j}) } $ with
1116: $E_{j}$ being the threshold energies and the sign ambiguity will
1117: be resolved in a moment. All masses here are assumed to take
1118: their physical values. In \Ref{b3} $\eta = -\alpha \mu_{2}/p_{2}$
1119: and
1120: $G_{0}, F_{0}$ denote the standard Coulomb wave functions for
1121: orbital momentum $\ell=0$, defined in \cite{Abramowitz}.
1122: Finally, it should be noted that there is no
1123: ingoing wave in \Ref{b0}, as appropriate for a bound state problem.
1124: \par
1125: As mentioned above,
1126: to simplify matters, we assume that the interaction is separable, i.e.
1127: that the potential matrix is
1128: \begin{equation}
1129: V_{ij}(r,r') = -v(r)\,s_{ij}\,v(r'),
1130: \label{b4}
1131: \end{equation}
1132: where the function $v(r)$ represents the shape of the potential and
1133: the dimensionless parameters $s_{ij}$ are the measure of
1134: the strength of the
1135: potential. Time reversal implies $s_{ij}=s_{ji}$. With separable
1136: potentials, the system of integral equations \Ref{b0}
1137: can be solved analytically. To this end it is
1138: sufficient to multiply each of the equations by $v(r)$ and
1139: integrate over $r$. This gives a system of two homogeneous algebraic
1140: equations for the two unknown quantities
1141: $$
1142: X_{j} = \int_{0}^{\infty}v(r)\,u_{j}(r)\, \D r, \quad j=1,2
1143: $$
1144: and the latter will have a non-trivial solution if, and only if,
1145: the determinant of the system $D(W)$ vanishes. Expanding the determinant,
1146: we are led to the explicit bound state condition
1147: %\begin{eqnarray}
1148: \begin{equation}
1149: D(W) = \left (1+s_{11}\bra v |G_{1}^{+}(W)|v \ket \right )
1150: \left (1+s_{22}\bra v |G_{2}^{+}(W)|v \ket \right )
1151: % \nonumber \\
1152: - s_{12}^{2}\; \bra v |G_{1}^{+}(W)|v \ket \;
1153: \bra v |G_{2}^{+}(W)|v \ket=0
1154: \label{b5}
1155: \end{equation}
1156: %\end{eqnarray}
1157: where we have introduced the abbreviation
1158: $$
1159: \bra v |G_{j}^{+}(W)|v \ket=\int_{0}^{\infty}
1160: v(r)\, \bra r \left | G_{j}^{+}(W) \right | r'\ket
1161: \, v(r') \;\D r\; \D r'.
1162: $$
1163: The determinant can vanish only at some
1164: particular value of the energy $W=E_{B}$ that will be interpreted as the
1165: bound state energy.
1166: Normally, knowing the underlying interaction, by solving \Ref{b5}
1167: one obtains the binding energy. However, in the problem at issue we
1168: have a reversed situation: we know the binding energy from
1169: experiment and it is the interaction that we are after. In the case of
1170: the pionic hydrogen atom we have an unstable bound state in the
1171: charged channel and the binding energy will be a complex number.
1172: We set
1173: \begin{equation}
1174: \label{b50}
1175: E_{B} = E_{2} + E_{1s} - (\epsilon + \imath \Half \gamma)
1176: \end{equation}
1177: where $E_{1s}=-\mu_{2}\alpha^{2}/2$
1178: is the purely Coulombic 1s state binding energy. Since in our
1179: formalism there is no room for the radiative decay of the pionic hydrogen
1180: the partial width $\gamma$ is a fraction of
1181: the total width $\Gamma$ given by the formula $\gamma=\Gamma/(1+P^{-1})$
1182: where $P$ is the Panofsky ratio. It has been shown in ref.
1183: \cite{GIB86}
1184: that the effect of the ($\pi^{-}$,$\gamma$) reaction on the
1185: accounted for hadronic channels is negligible.
1186: The experimental values for
1187: $\epsilon, \Gamma$ (cf. \cite{Sigg})
1188: and P (cf. \cite{Panofsky}) adopted in this work, are
1189: \begin{eqnarray*}
1190: \epsilon &=&7.108 \pm 0.013\text{(stat)}\pm 0.034\text{(syst)}\,eV,\\
1191: \Gamma &=& 0.868 \pm 0.040\text{(stat)} \pm 0.038\text{(syst)}\,eV,\\
1192: P &=& 1.546\pm 0.009,
1193: \end{eqnarray*}
1194: and in the following we shall take
1195: $\epsilon=7.108\pm 0.047\,eV$ and $\gamma=0.527\pm 0.047\,eV$
1196: as the input values.
1197: It must be immediately explained here that in this work
1198: we have defined $\epsilon$ in accordance with a different convention,
1199: so that our $\epsilon$ has opposite sign than that used in ref.
1200: \cite{Sigg}.
1201: In our approach we have tacitly assumed that
1202: under perturbative treatment
1203: all electromagnetic corrections contribute the same amount to the
1204: purely Coulombic level and to the level shifted by strong
1205: interaction. More precisely, we are going to ignore the small
1206: effects caused by the distortion of the wave function. Accordingly,
1207: the electromagnetic corrections need not concern us here and they
1208: have been left out altogether but, of course, they would be
1209: indispensable for calculating the total displacement of the level from
1210: its Coulombic position.
1211: \par
1212: The pole of the T-matrix that corresponds to the solution of \Ref{b5}
1213: can be located on one of the four Riemann sheets as appropriate for
1214: a two-channel problem. This is also apparent from the
1215: mentioned above sign ambiguity
1216: in the definition of the channel momenta in \Ref{b3}.
1217: The right choice of the Riemann sheet is
1218: essential and this can be accomplished by proper adjustment of
1219: the signs of the imaginary parts of the channel momenta $p_{j}$.
1220: We are using here the standard enumeration of the Riemann sheets
1221: introduced in ref. \cite{HENDRY}, i.e.
1222: \begin{eqnarray*}
1223: \text{sheet I:} \quad & Im\, p_{1}>0;& \; Im\, p_{2}>0\\
1224: \text{sheet II:} \quad & Im\, p_{1}<0;& \; Im\, p_{2}>0\\
1225: \text{sheet III:} \quad & Im\, p_{1}<0;& \; Im\, p_{2} <0\\
1226: \text{sheet IV:} \quad & Im\, p_{1}>0;& \; Im\, p_{2}<0.
1227: \end{eqnarray*}
1228: In the pionic hydrogen case, with an unstable bound state in channel 2,
1229: we have to enforce the pole to be located on the second sheet.
1230: \par
1231: To proceed further we need some concrete shape factor $v(r)$ and our
1232: choice here is the exponential shape, i.e. we set
1233: \begin{equation}
1234: v(r)=\sqrt{\beta^{3}/\mu}\; \exp{(-\beta\,r)}
1235: \label{b6}
1236: \end{equation}
1237: where $\mu$ is the reduced pion-nucleon mass in the case of exact isospin
1238: symmetry (we take average mass for each isospin multiplet) and
1239: $\beta$ is the inverse range parameter.
1240: With the exponential form \Ref{b6}, the potential \Ref{b4} is
1241: identical with the familiar Yamaguchi potential
1242: and the Green's function matrix elements can be obtained in
1243: an analytic form. The final result is
1244: \begin{equation}
1245: \bra v |G_{1}^{+}(W)|v \ket=-\frac{\mu_{1}}{\mu}
1246: \; \frac{1}{(1-\imath p_{1}/\beta)^{2}}
1247: \label{b7}
1248: \end{equation}
1249: for the neutral channel, while the corresponding formula for the
1250: charged channel reads
1251: \begin{equation}
1252: \bra v |G_{2}^{+}(W)|v \ket=-\frac{\mu_{2}}{\mu}
1253: \; \frac{1}{(1-\imath p_{2}/\beta)^{2}}
1254: \; \frac{\mbox{}_{2}F_{1}(1,\,\imath \eta;\,\imath \eta+2;\,z^{2})}
1255: {\imath \eta+1}
1256: \label{b8}
1257: \end{equation}
1258: with $z=(\beta+\imath p_{2})/(\beta-\imath p_{2})$.
1259: The last fraction in \Ref{b8} accounts for the Coulomb interaction and
1260: the symbol
1261: $\mbox{}_{2}F_{1}(a,b;c;z)$ denotes the hypergeometric function
1262: defined in \cite{Abramowitz}.
1263: The computation of the hypergeometric function entering
1264: \Ref{b8} is greatly simplified
1265: owing to the fact that the first parameter is equal to unity
1266: in which case the
1267: continued fraction representation of $\mbox{}_{2}F_{1}(1,b;c;z)$
1268: discovered by Gauss \cite{Gauss} proves to be useful.
1269: The continued fraction summation converges
1270: in the whole of the complex $z$ plane away form the branch cut
1271: on the real axis running from one to infinity.
1272: \par
1273: With exact isospin symmetry the three strength parameters $s_{11},
1274: s_{12}, s_{22}$ are not independent and can be expressed in terms of
1275: isospin 1/2 and isospin 3/2 strengths denoted hereafter
1276: as $s_{1}$ and $s_{3}$, respectively.
1277: In the bound state condition \Ref{b5} both the real and the imaginary
1278: part of $D(E_{B})$ have to vanish simultaneously and that gives us two
1279: real equations. Since the bound state energy is known
1280: (cf. \Ref{b50}), we put
1281: $ s_{11}= (s_{1}+2s_{3})/3\; ; s_{22}= (2s_{1}+s_{3})/3\; ;
1282: s_{12}= \sqrt{2}(s_{3}-s_{1})/3 $
1283: in \Ref{b5}, and regarding $s_{1}$ and $s_{3}$
1284: as our two unknowns,
1285: we arrive at two algebraic equations that can be solved analytically
1286: \begin{subequations}
1287: \label{b55}
1288: \begin{equation}
1289: s_{1}^{2}\,\text{Im}\,ac^{*}+s_{1}\,\text{Im}(ab^{*}-c)-\text{Im}\,b=0;
1290: \label{b55:a}
1291: \end{equation}
1292: \begin{equation}
1293: s_{3}=-(1+s_{1}\,\text{Re}\,a)/(\text{Re}\,b+s_{1}\,\text{Re}\,c),
1294: \label{b55:b}
1295: \end{equation}
1296: \end{subequations}
1297: where $a=(\bra v|G_{1}^{+}(E_{B})|v\ket
1298: +2\bra v|G_{2}^{+}(E_{B})|v\ket)/3;
1299: b=(2\bra v|G_{1}^{+}(E_{B})|v\ket +\bra v|G_{2}^{+}(E_{B})|v\ket)/3$ and
1300: $c=\bra v|G_{1}^{+}(E_{B})|v\ket \bra v|G_{2}^{+}(E_{B})|v\ket$.
1301: With $s_{1}$ and $s_{3}$ in hand, the corresponding
1302: scattering lengths ($a_{2I}$ with I=1/2 and 3/2) are obtained from
1303: \begin{equation}
1304: a_{2I}=\frac{2}{\beta} \frac{s_{2I}}{1-s_{2I}}.
1305: \label{b10}
1306: \end{equation}
1307: \par
1308: For local potentials the method outlined above could be
1309: also applied but in such case it would be more
1310: convenient to use instead of \Ref{b0} an equivalent
1311: set of two coupled Schr\"odinger equations. For fixed energy
1312: and the proper choice of the Riemann sheet,
1313: these differential equations can be integrated numerically and the
1314: bound state equation is obtained from the requirement of vanishing
1315: of the Wronskian determinant. The latter is again a function of the
1316: isospin 1/2 and isospin 3/2 strength parameters, or
1317: if one prefers, the corresponding potential
1318: depths. Although the bound state condition is
1319: defined then only numerically but from it one can get
1320: two real equations that can be solved numerically using standard procedures.
1321: With a local potential, however, the solution of the three-body
1322: problem becomes much more complicated and this is the main reason
1323: why we preferred to work with a separable potential.
1324: \par
1325: Our calculational scheme is now complete
1326: and we shall present our results. Using as our input
1327: the experimental values of the pionic hydrogen level shift and width,
1328: the bound state equation has been solved analytically
1329: by adopting a number of ''reasonable'' values for $\beta$ and
1330: in our computations we have used the
1331: values from $2\,fm^{-1}$ to $10\,fm^{-1}$. Although,
1332: we do not know the precise value of the range but
1333: there is no physical mechanism known that might generate
1334: long range forces in the $\pi N$ system,
1335: the longest range is unlikely to be bigger than $0.5\,fm$ and
1336: this sets the lower limit of acceptable $\beta$ values. In principle,
1337: there is no upper limit for $\beta$ but for $\beta>10\,fm^{-1}$ we have
1338: in practice reached the limit of the zero-range forces
1339: and things change very little above that limit.
1340: The exact solutions of the bound state equation are presented in
1341: Table \ref{table4}
1342: where the errors reflect only the experimental uncertainty of
1343: our input, i.e. $\epsilon, \Gamma$ and the Panofsky ratio.
1344: Our isoscalar and isovector scattering lengths are in
1345: good agreement with the values extracted in \cite{Sigg}.
1346: This has been illustrated in Fig. \ref{fig3} where we have compared
1347: a representative sample of
1348: our computations with the values obtained by
1349: Sigg {\it et al.} \cite{Sigg}.
1350: The solutions corresponding to $\beta$ spanning the range
1351: $2-10$ fm$^{-1}$
1352: are located very close to each other in the (b$_{1}$, b$_{0}$) plane
1353: and putting more than three points on the plot might have obscured
1354: the picture. The error bars reflect only the experimental
1355: uncertainty of our input.
1356: As mentioned above, the bound state equation \Ref{b5}
1357: yields a second order equation for $s_{1}$ and $s_{3}$ and therefore
1358: we have always two solutions (cf. \Ref{b55}).
1359: Only one of them is presented in Table \ref{table4} whereas
1360: the second solution leads to both $b_{0}$ and $b_{1}$ positive and
1361: has had to be rejected.
1362: When the two strength parameters $s_{1}$ and $s_{3}$ are known
1363: we can calculate not only the scattering lengths
1364: but also the effective ranges
1365: in each of the two isospin states and these values are presented
1366: in Table \ref{table4}. Instead of the effective range we
1367: use the parameter $B_{2I}$ that is defined from the
1368: expansion of the real part of the s-wave scattering amplitude
1369: in powers of the c.m. momentum $k$, i.e. close to threshold, we have
1370: $Re f_{2I}(k)=a_{2I}+B_{2I}\,k^{2}+\cdots$.
1371: For comparison, at the bottom of Table
1372: \ref{table4} we give the values of all parameters inferred from
1373: a recent phase shift analysis \cite{Gashi}.
1374: The calculated scattering lengths, listed in Table \ref{table4},
1375: are almost independent upon $\beta$, in contrast with the slope
1376: parameters $B_{2I}$ which change quite a bit when $\beta$
1377: is varied in the interval 2-10 fm$^{-1}$. In addition to that,
1378: our $B_{3}$ values
1379: turn out to be always positive and therefore have opposite sign
1380: than those deduced from phase shift analysis \cite{GIB98,Gashi}.
1381: Actually, the pionic hydrogen data
1382: provide a strong constraint only for the scattering lengths and
1383: sticking to a simple $\pi$N Yamaguchi potential
1384: it is not possible to get $B_{3}$ negative just
1385: by varying $\beta$. Indeed, for fixed a$_{2I}$
1386: the slope parameter B$_{2I}$
1387: is given by the exact formula
1388: $$
1389: B_{2I}=-a_{2I}^{3}\;[1+\frac{1}{2\beta a_{2I}}(3+\frac{4}{\beta
1390: a_{2I}})]
1391: $$
1392: and since the expression in the square bracket is necessarily positive
1393: the sign of B$_{2I}$ is bound to be opposite to that of a$_{2I}$.
1394: To obtain a negative B$_{3}$ a more sophisticated
1395: potential involving both repulsion and attraction would have been required
1396: \cite{GIB98}.
1397: There is no need for such extension, however, because our model
1398: has been devised for describing only the near threshold phenomena
1399: and is quite adequate at that. Expanding the phase shift
1400: close to threshold in powers of $k$, we have
1401: $\delta_{2I}=a_{2I}k+O(k^{3})$ and it is apparent that a model
1402: providing merely the scattering length reproduces satisfactorily
1403: the phase shift in the neighbourhood of zero where $\delta_{2I}$
1404: exhibits a linear behaviour.
1405: In our case this is all that counts as we never
1406: deal with higher energies. This means that
1407: the determination of the slope parameters is out of reach
1408: within our model since the appropriate
1409: energy scale has been set by the Coulomb energy in the
1410: pionic hydrogen, in which case terms proportional to $B_{2I}$
1411: make negligible contribution. For an assigned value of $\beta$
1412: the slope parameters may be calculated but they are of no
1413: physical significance and comparing them with those resulting
1414: from phase shift analysis does not make much sense.
1415: \par
1416: As noted in \cite{Sigg,GIB86},
1417: at the energy value close to the unstable bound state
1418: in channel 2, the scattering
1419: amplitude in the open channel 1, shows a strong resonant behaviour.
1420: For a separable potential, the s-wave scattering amplitude $f(W)$
1421: in channel 1 may be easily
1422: calculated analytically and takes a simple form
1423: \begin{equation}
1424: f(W)= \text{e}^{ \I \delta } \sin \, \delta /p_{1}=
1425: \frac{\mu_{1} } {\mu} \frac{2}{\beta} \quad
1426: \frac{ s_{11}+ (s_{11}s_{22}-s_{12}^{2})
1427: \bra v |G_{2}^{+}(W)|v \ket } {(1+p_{1}^{2}/\beta^{2} )^{2}\;D(W) },
1428: \label{b16}
1429: \end{equation}
1430: where $\delta$ is the corresponding phase shift that for real $W$
1431: below the $\pi^{-}$p threshold is a real number.
1432: The resonance is not of a Breit-Wigner shape but its position $E_{r}$
1433: may be easily established from \Ref{b16} as the energy at which
1434: the phase shift is equal to $\Half\pi$. Close to the
1435: resonant energy, i.e. for $W\approx E_{r}$ we have
1436: $\cot \delta \approx (W-E_{r})/(\Half\Gamma_{r})$ and this allows
1437: us to infer the value of the width $\Gamma_{r}$ of the resonance.
1438: In ref. \cite{Sigg} the values of $(\epsilon,\gamma)$ have
1439: been calculated by identifying them with $(E_{2}+E_{1s}-E_{r},\Gamma_{r})$.
1440: In principle, the values of $(\epsilon,\gamma)$
1441: obtained that way do not have to be identical with those determined
1442: from the position of the bound state pole. To check that point,
1443: we have repeated the procedure of ref. \cite{Sigg}
1444: but using our separable potentials
1445: whose depths have been adjusted to reproduce the values
1446: of $(\epsilon,\gamma)$ obtained in \cite{Sigg}.
1447: We found that the two methods give nearly identical results and
1448: the differences in $(\epsilon,\gamma)$ did not exceed 1 meV.
1449: For illustration, in Fig.\ref{fig3} we show the behaviour of $\sin \delta$
1450: close to the resonance for the case of $\beta=3\,fm^{-1}$ where
1451: the strengths parameters inferred from the pole location
1452: were $s_{1} =0.271820$ and $s_{3}=-0.245868$.
1453: \par
1454: Before concluding
1455: our discussion of the pionic hydrogen we wish to mention
1456: one last thing, namely we are going to show
1457: how from \Ref{b5} one can retrieve the Deser-Trueman formula
1458: (cf. \cite{Deser}). This task will be accomplished by obtaining an
1459: approximate solution of \Ref{b5} and to this end
1460: \Ref{b5} is cast to the form
1461: \begin{equation}
1462: 1+s_{\text{eff}}(W)\; \bra v |G_{2}^{+}(W)|v \ket =0,
1463: \label{b11}
1464: \end{equation}
1465: where we have introduced an effective energy dependent complex strength
1466: parameter $s_{\text{eff}}$,
1467: defined as
1468: \begin{equation}
1469: s_{\text{eff}}(W)=s_{22}-s_{12}^{2} \bra v|G_{1}^{+}(W)|v \ket
1470: /\left (1+s_{11} \bra v|G_{1}^{+}(W)|v \ket \right ).
1471: \label{b12}
1472: \end{equation}
1473: The complex $\pi^{-} p$ scattering length $a_{\pi p}$ can be expressed
1474: in terms of $s_{\text{eff}}(W)$ evaluated at threshold
1475: \begin{equation}
1476: a_{\pi p}= \frac{\mu_{2}\,2}{\mu\,\beta}
1477: \frac{s_{\text{eff}}(E_{2})} {1-s_{\text{eff}}(E_{2})},
1478: \label{b13}
1479: \end{equation}
1480: and the Coulomb corrected $\pi^{-}$p scattering length, denoted as
1481: $a_{\pi p}^{c}$, can be obtained from the exact formula derived in
1482: \cite{vanH}
1483: \begin{equation}
1484: 1/a_{\pi p}^{c}=\text{e}^{\xi}/a_{\pi p}+
1485: 2\mu_{2}\alpha \; \text{Ei}(\xi),
1486: \label{b13a}
1487: \end{equation}
1488: where $\xi = 4 \alpha\mu_{2}/\beta$ and
1489: Ei($\xi$) is the exponential integral function defined in
1490: \cite{Abramowitz}. It should be noted here that the zero-range limit
1491: ($\beta\to \infty$) does not exist in \Ref{b13a} because
1492: the function Ei($\xi$) for $\xi$=0
1493: has a logarithmic singularity. For the case of
1494: $\beta=3\,fm^{-1}$ just considered, we obtain
1495: \begin{eqnarray*}
1496: a_{\pi p} &=& 0.12081 + \imath\, 0.004441\,fm; \\
1497: a_{\pi p}^{c}&=& 0.12068 + \imath\, 0.004458\,fm;
1498: \end{eqnarray*}
1499: so that the Coulomb corrections do not exceed a fraction of a
1500: percent. However, in general,
1501: the Coulomb correction is model dependent, and, in particular, it is rather
1502: sensitive to the range of the nuclear potential what can be seen
1503: when the above result is juxtaposed with the $\pi$d case where
1504: the range of the potential was comparable with the size of the deuteron
1505: and, accordingly, the Coulomb correction to $\pi$d
1506: scattering length was much bigger (1.5\%).
1507: \par
1508: Since we wish to obtain an approximate solution of \Ref{b11} that
1509: is located not far from the Coulomb bound state, we set
1510: $W=E_{2}+E_{1s}+\delta E$ where $\delta E$ is a small displacement.
1511: To calculate $\delta E$ and
1512: derive the Deser-Trueman formula
1513: from \Ref{b11}, we have to assume that ({\it i}) the
1514: complex energy shift $\delta E = -\epsilon -\imath \Half \gamma $ is small
1515: in comparison with Coulomb energy ($|\delta E/E_{1s}|<<1 $), and,
1516: ({\it ii}) that the range of the strong interaction is small as compared
1517: with the Bohr radius ($\beta>>\mu_{2}\alpha $).
1518: Introducing a complex momentum
1519: $p_{c}=\sqrt{2\mu_{2}E_{1s}}=\imath \mu_{2}\alpha $
1520: corresponding to the Coulomb bound state, we can see that
1521: when $p_{2} \to p_{c}$ then $\imath \eta \to -1 $ and
1522: the Green's function \Ref{b8}
1523: occurring in \Ref{b11} becomes singular.
1524: This singularity is of paramount importance since it induces a zero in
1525: the nuclear S-matrix that is necessary to cancel the bound pole in the
1526: Coulomb S-matrix. As a result of this cancellation, the full S-matrix
1527: in the charged channel, which is a product of the Coulomb S-matrix
1528: and the nuclear S-matrix, remains finite at $p_{2}=p_{c}$.
1529: In compliance with the small shift assumption, we set
1530: $p_{2}=p_{c}+\delta p $ where $\delta p $ is supposed to be a small
1531: correction and since the most rapid variation in \Ref{b8}
1532: arises on account of the pole term, we approximate
1533: $1+\imath \eta$ by $\delta p/p_{c} $. Apart from
1534: that, elsewhere we replace $p_{2}$ by $p_{c}$.
1535: The hypergeometric function
1536: for $\imath\eta = -1$ reduces to a polynomial $1-z^{2}$
1537: and neglecting small terms
1538: of the order of $p_{c}/\beta $, from \Ref{b11} we obtain
1539: \begin{equation}
1540: \delta p \approx -4\imath\,(p_{c}^{2}/\beta)\,(\mu_{2}/\mu)
1541: s_{\text{eff}}(E_{2}) \approx -2\imath\, p_{c}^{2}\,a_{\pi p}
1542: \label{b14}
1543: \end{equation}
1544: where we have used \Ref{b13}
1545: retaining only linear term in $a_{\pi p} $. The above result
1546: gives the Deser-Trueman formula \cite{Deser} in its standard form
1547: \begin{equation}
1548: \delta E \approx p_{c}\, \delta p/\mu_{2} \approx
1549: -2\mu_{2}^{2}\,\alpha^{3}\, a_{\pi p},
1550: \label{b15}
1551: \end{equation}
1552: where, in view of the above discussion, it does not really matter whether
1553: we take $a_{\pi p}$ or $a^{c}_{\pi p}$.
1554: It is perhaps in order to recall that although the Deser-Trueman
1555: formula \Ref{b15} has been derived here for a specific choice of
1556: the underlying interaction, but its validity is quite general.
1557: To examine the accuracy of Deser-Trueman formula we turn again to our
1558: previous example when $\beta=3\,fm^{-1}$ and by computing $a_{\pi p}$
1559: from \Ref{b13} and inserting in
1560: \Ref{b15}, we obtain
1561: $(\epsilon, \gamma)$ = (7.024, 0.516) eV to be compared with our input
1562: values equal $(\epsilon, \gamma)$ = (7.108, 0.527) eV that ought
1563: to have been reproduced if formula \Ref{b15} had been exact.
1564: It is a remarkable property of the Deser-Trueman formula that it is
1565: independent of the range of the underlying interaction and therefore
1566: the error in this formula must be of the same size as the uncertainty
1567: in the exact result caused by varying $\beta$. If one is prepared
1568: to tolerate such uncertainty formula \Ref{b15} could be used to
1569: infer $a_{1}$ and $a_{3}$. Introducing a two-channel K-matrix,
1570: isospin invariance can be invoked to pin down its
1571: elements at the single unsplit threshold
1572: $$
1573: K=\left (\;
1574: \begin{matrix}
1575: & \tfrac{1}{3}a_{1}+\tfrac{2}{3}a_{3}
1576: & \tfrac{\sqrt{2}}{3}(a_{3}-a_{1})\\
1577: & \tfrac{\sqrt{2}}{3}(a_{3}-a_{1})
1578: & \tfrac{2}{3} a_{1}+\tfrac{1}{3} a_{3}\\
1579: \end{matrix}
1580: \;\right )
1581: $$
1582: and the complex $\pi^{-}p$ scattering length takes the form
1583: \begin{equation}
1584: a_{\pi p}= K_{22}+\imath p_{t}K_{12}^{2}/
1585: (1-\imath p_{t}K_{11}),
1586: \label{b30}
1587: \end{equation}
1588: where $p_{t}$ is the momentum in the $\pi^{0}n$ channel evaluated at
1589: the $\pi^{-}p$ threshold. The scattering length \Ref{b30},
1590: unlike \Ref{b13}, does not depend upon the range.
1591: Inserting \Ref{b30} in \Ref{b15}
1592: and separating the real and the imaginary part, we end up with
1593: two real equations for the two unknowns $a_{1}$ and $a_{3}$.
1594: To more than sufficient accuracy, the explicit solutions, are
1595: \begin{subequations}
1596: \label{b31}
1597: \begin{eqnarray}
1598: a_{1}&=&\left [x \pm y(1-2p_{t}y)/\sqrt{2p_{t}y}\right ]/(1-p_{t}y);
1599: \label{b31:a}\\
1600: a_{3}&=&\left [ x \mp y(2-p_{t}y)/\sqrt{2p_{t}y}\right ]/(1-p_{t}y),
1601: \label{b31:b}
1602: \end{eqnarray}
1603: \end{subequations}
1604: where $x=\epsilon/2\mu_{2}^{2}\alpha^{3},
1605: y=\Half\gamma/2\mu_{2}^{2}\alpha^{3}$ and the double sign in \Ref{b31}
1606: stems the fact that eq. \Ref{b30} is quadratic in $a_{2I}$.
1607: If ($\epsilon$,$\gamma$)
1608: have been obtained in a model independent way then the results \Ref{b31}
1609: are also model independent.
1610: As seen from Table \ref{table4} the uncertainty on $a_{1}$ and
1611: $a_{3}$ (3\% and 9\%, respectively) induced by experimental errors
1612: on $(\epsilon, \gamma)$ is much bigger than the uncertainty caused by
1613: varying $\beta$ (about 1\%).
1614: Under these circumstances it is perfectly justified
1615: to infer the $\pi$N scattering lengths
1616: via Deser-Trueman formula and their numerical values
1617: obtained from \Ref{b31} are displayed in Table \ref{table4} whereas
1618: the corresponding $b_{0}$ and $b_{1}$ are presented in Fig. \ref{fig3}.
1619: \par
1620: It is apparent from Table \ref{table4} that to improve upon
1621: Deser-Trueman formula we need
1622: some additional clue concerning $\beta$ and it becomes something
1623: of a challenge to find ways to ferret out more precisely what the
1624: value of $\beta$ might be.
1625: So far in our considerations we have not mentioned yet the pionic deuterium
1626: data and at this stage it is logical to ask whether this additional
1627: information might not help to pin down the
1628: range parameter of the $\pi$N potential. Therefore,
1629: in the next step, we use the values given in Table \ref{table4}
1630: as input for a
1631: three-body calculation, i.e. we use the separable potential \Ref{b6}
1632: in the Faddeev equations
1633: for calculating the $\pi d$ scattering length. The results of our
1634: computations are presented in Fig. \ref{fig4} where we have plotted the
1635: $\pi d$ scattering length versus $\beta$. The full circles represent
1636: the results obtained by the including the p-wave interaction
1637: (more precisely, just the P33 wave), while
1638: the open circles correspond to a situation where the delta has been
1639: left out. For reasons of clarity of the presentation these two sets
1640: of points have been given at
1641: different $\beta$ values.
1642: The indicated error bars reflect the uncertainty in the
1643: input values (cf. Table \ref{table4} ).
1644: For comparison, the experimental value of
1645: $\pi d$ scattering length to within one standard deviation
1646: is given in Fig \ref{fig4}
1647: as the area between the two horizontal lines. The striking
1648: feature apparent from Fig \ref{fig4}
1649: is that the results are almost independent
1650: of the range parameter $\beta$. Furthermore, the calculated
1651: scattering lengths are consistent with experiment for all $\beta$
1652: no matter whether the delta has been included or not.
1653: This result may come as a disappointment since the deuteron data
1654: give no illumination how to bracket the value of $\beta$.
1655: \par
1656: In order to understand how the above result comes about we shall
1657: invoke again the static model, taking advantage of the fact that
1658: with the Yamaguchi potential representing the $\pi$N
1659: interaction the static solution
1660: of the Faddeev equations may be readily obtained (cf. ref.\cite{SE} ).
1661: Thus, introducing
1662: the Yamaguchi form factors and going to the static limit we can repeat
1663: the procedure outlined in the preceding section. The static solution
1664: of the Faddeev equations may be then sought in the form
1665: \begin{eqnarray*}
1666: A(\bbox{q},\bbox{k})&=&
1667: -\frac{m}{2\pi}\;\frac{\beta^{2}}{q^{2}+\beta^{2}}\;{\cal A}(k);\\
1668: X(\bbox{q},\bbox{k})&=&
1669: -\frac{m}{2\pi}\;\frac{\beta^{2}}{q^{2}+\beta^{2}}\;{\cal X}(k),
1670: \end{eqnarray*}
1671: and the above ansatz used in the Faddeev equations yields a set of
1672: two integral equations that differ from \Ref{a23} in that the
1673: appropriate kernels contain now an extra factor
1674: $1/[1+(\bbox{k}+\bbox{k'})^{2}/\beta^{2}]^{2}$. Despite this
1675: additional complication, the
1676: Fourier transform of this extended kernel still can be effected
1677: and leads to a simple analytic expression
1678: $$
1679: \frac{4\pi}{\kappa^{2}+(\bbox{k}+\bbox{k'})^{2}}
1680: \frac{\beta^{4}}{[\beta^{2}+(\bbox{k}+\bbox{k}')^{2}]^{2}}=
1681: (1-\frac{\kappa^{2}}{\beta^{2}})^{-2} \int
1682: e^{-\imath(\bbox{k}+\bbox{k}')\bbox{r}} \frac{\D^{3}\,r}{r}
1683: \left \{e^{-\kappa r}-e^{-\beta r}[1+\frac{\beta r}{2}
1684: (1-\frac{\kappa^{2}}{\beta^{2}})] \right \}.
1685: $$
1686: Using the above formula,
1687: similarly as before, we end up with a system of two algebraic
1688: equations for $A(r)$ and $X(r)$. Neglecting the binding energy correction
1689: ($\kappa \to 0$), the resulting equations differ from \Ref{a8} in that
1690: the zero-range pion propagator $1/r$ has to be multiplied by
1691: the function $g(r)$ given by the formula
1692: \begin{equation}
1693: \label{b17}
1694: g(r)= 1- e^{-\beta r}(1+\Half \beta r).
1695: \end{equation}
1696: Therefore, the sought for solution for $A(r)$ follows from \Ref{a10}
1697: after replacing $1/r$ by $g(r)/r$. Formula \Ref{b17} proves to be quite
1698: useful for estimating the size of the $\beta$ dependent
1699: correction and to this end we need to evaluate $g(r)$ at some
1700: average value of $r$ and a plausible candidate for such average value
1701: is the deuteron radius
1702: $r_{d}=\Half\sqrt{<r^{2}>}\approx 2$ fm.
1703: Indeed, with this choice the second order formula \Ref{a12} that provides
1704: for a major contribution to $a_{\pi d}$ will be little affected since
1705: by setting $r=r_{d}$,
1706: we get $\bra 1/r \ket$ = 0.5 fm$^{-1}$, not far from the values listed
1707: in Table \ref{table2}. When $\beta$ is varied in
1708: the range $2-10$ fm$^{-1}$, we have r$_{d}\beta >$4
1709: in the exponential damping factor in \Ref{b17}, so that the
1710: $\beta$ dependent terms make a contribution to $g(r)$ at the level of
1711: a few percent and the resulting $\pi$d scattering length is almost
1712: independent upon $\beta$. This feature, sustained in the full
1713: Faddeev solution, is a consequence of the fact that the adopted range
1714: of the $\pi$N forces was small as compared with the deuteron radius.
1715: \par
1716: In conclusion, we have seen that
1717: the uncertainty in the calculated $a_{1}$ and $a_{3}$,
1718: as well as in $a_{\pi d}$, connected with the lack of knowledge of the
1719: range parameter constitutes only a small fraction of the uncertainty
1720: resulting from the experimental errors on the pionic hydrogen data.
1721: The above results may be viewed as an {\it a posteriori} justification
1722: of our zero-range model developed in Sec. \ref{se:two}: introducing
1723: a finite range would be merely a fine tuning
1724: which is not yet affordable in the current state of affairs.
1725:
1726: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1727: \section{Discussion} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1728: \label{se:four} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1729: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1730:
1731: Assuming that the underlying $\pi$N interaction is isospin invariant,
1732: we have analysed the recent pionic hydrogen and pionic deuterium data
1733: with the purpose to extract from them $\pi$N s-wave
1734: scattering lengths $a_{2I}$ for I=1/2 and I=3/2.
1735: It is an empirical fact that the complex energy shift
1736: in either of these two atomic systems is small
1737: when compared with the corresponding Coulomb energy and
1738: with the appropriate Bohr radii setting the length scale,
1739: the $\pi$-p and $\pi$-d interactions are of a short range.
1740: Under these circumstances Deser-Trueman formula
1741: provides an extremely good approximation,
1742: relating in a model independent way the 1s level shifts and widths
1743: in the pionic hydrogen and pionic deuterium to the complex scattering
1744: lengths $a_{\pi p}$ and $a_{\pi d}$, respectively. However,
1745: to infer $a_{2I}$ from the latter quantities is a non-trivial
1746: dynamical problem and to be able to solve it we
1747: introduced a simple and transparent potential representation of
1748: the $\pi$N interaction.
1749: Within this model we obtain explicit solution
1750: of the $\pi^{-}p$ bound state problem and also of the related
1751: three-body $\pi$d scattering problem at zero energy.
1752: \par
1753: We have assumed throughout this work
1754: that the $\pi$N forces are of a very
1755: short range and this supposition follows from a particle
1756: exchange picture: there is no sufficiently light particle presently known
1757: that might be capable of generating forces whose range would
1758: exceed 0.3-0.4 fm (which roughly corresponds to a vector meson exchange).
1759: In this situation it was logical to take the zero-range limit
1760: as our point of departure.
1761: In order to find out what the deuteron data
1762: can teach us about $\pi$N scattering lengths,
1763: we calculated the $\pi$d scattering length
1764: by solving the appropriate three-body $\pi$NN problem.
1765: This task was accomplished, both within the static
1766: approximation, and also by using the Faddeev formalism.
1767: We demonstrated that the same static formula for $a_{\pi d}$ can
1768: be derived from: {\it (i)} a set of boundary conditions; {\it (ii)}
1769: a static solution of Faddeev equations, and {\it (iii)}
1770: a summation of Feynman diagrams. The static formula
1771: expressing $a_{\pi d}$ in terms of $\pi$N scattering
1772: lengths was found to be surprisingly accurate:
1773: the error, estimated by comparing the static result with
1774: the full Faddeev solution, was at the level of 2\%,
1775: i.e. of the same size as the experimental error on $a_{\pi d}$.
1776: The standard second order formula was shown
1777: to be insufficient: the incurred error was
1778: three times bigger than the present experimental
1779: uncertainty on $a_{\pi d}$.
1780: Using as input the $\pi$N scattering lengths, that had been
1781: inferred earlier \cite{Sigg} from pionic hydrogen data,
1782: we obtained $a_{\pi d}$ by solving
1783: the Faddeev equations for zero-range $\pi$N forces.
1784: The requirement that the calculated
1785: $a_{\pi d}$ be in agreement with experiment
1786: to within one standard deviation, imposes bounds
1787: on the isoscalar and isovector $\pi$N scattering lengths.
1788: The values of the $\pi$N scattering lengths
1789: determined that way, consistent
1790: with both the pionic hydrogen and the pionic deuterium data,
1791: are presented in Fig. \ref{fig1}.
1792: \par
1793: In the next stage of this investigation we
1794: lifted the zero-range limitation introducing a range parameter.
1795: The pionic hydrogen bound state problem was solved afresh for
1796: a variety of range values.
1797: We derived the appropriate bound state condition and
1798: taking the 1s level shift and width of the pionic
1799: hydrogen as input, we used this condition to
1800: determine the s-wave $\pi$N potentials. This was possible
1801: since a complex condition is equivalent to two real equations, which
1802: for an assigned range, can be exactly solved
1803: for the I=1/2 and I=3/2 depth
1804: parameters entering the $\pi$N potentials.
1805: Knowing the potentials, it was a trivial matter to calculate
1806: the corresponding s-wave scattering amplitudes.
1807: As can be seen from Table \ref{table4},
1808: the resulting $\pi$N scattering lengths are rather insensitive to the
1809: adopted value of the range parameter.
1810: \par
1811: The analysis of the pionic hydrogen presented in this work parallels that
1812: given in \cite{Sigg}. We differ, however, in the adopted dynamical
1813: frameworks: in \cite{Sigg} Klein-Gordon equation together with a local
1814: $\pi$N potential has been used, whereas we consider a non-relativistic
1815: Lippmann-Schwinger equation (equivalent to a Schr\"odinger equation)
1816: with a separable $\pi$N potential.
1817: As may be seen from Fig. \ref{fig3}, the $\pi$N scattering lengths
1818: inferred in this paper are in good agreement with those deduced
1819: in \cite{Sigg}. This is a direct consequence of
1820: the fact that Deser-Trueman formula provides such a
1821: good approximation that we can make considerable progress in deducing
1822: the $\pi$N scattering lengths without committing ourselves in great
1823: deal to the nature of the $\pi$N dynamics.
1824: Since Deser-Trueman formula depends neither upon the shape of the $\pi$N
1825: potential nor upon its range, the small changes in the
1826: $\pi$N scattering lengths
1827: caused by varying the range parameter, must be attributed to the
1828: differences between the approximate Deser-Trueman formula and the
1829: exact range dependent solutions
1830: of the bound state equation. Thus, Fig. \ref{fig3} illustrates the
1831: accuracy of Deser-Trueman formula.
1832: \par
1833: For an assigned range value, the pionic hydrogen data specify completely the
1834: $\pi$N potentials, so that they may be used in
1835: the Faddeev equations in order to obtain the $\pi$d scattering length.
1836: The latter quantity was shown to be almost independent upon the range
1837: parameter (cf. Fig. \ref{fig4}) but was rather sensitive to the values of the
1838: $\pi$N scattering lengths used as input in the Faddeev equations.
1839: The above finding, supporting the zero-range approach,
1840: could be explained by the fact that
1841: the range of the $\pi$N interaction that was considered physically justified
1842: was small in comparison with the deuteron size.
1843: \par
1844: We conclude that the
1845: lack of knowledge of the range of the $\pi$N interaction is
1846: responsible for some
1847: uncertainty in the deduced $\pi$N scattering lengths
1848: but this uncertainty is rather small, at the level of 1\%.
1849: The main source of error is still the experimental
1850: uncertainty in the pionic hydrogen data.
1851: \par
1852: It is rather obvious that the presented model
1853: contains several omissions but
1854: we think that they are not too severe, especially that the investigation
1855: has been confined to near threshold phenomena.
1856: As in all non-relativistic models based on static potentials virtual particle
1857: production, crossing symmetry, retardation and relativistic effects
1858: have not been even touched upon.
1859: Besides that, a separable potential is not considered
1860: to have a strong theoretical basis and has been adopted here merely
1861: for convenience as it simplifies considerably the solution of
1862: the Faddeev equations. There are also limitations on the completeness
1863: of the Faddeev approach where by restriction to three-body channels
1864: we were forced to leave out a wealth of inelastic features. The
1865: absorption channels leading to two-nucleon states are not easily
1866: incorporated in a Faddeev theory and require considerable
1867: enlargement of the present model which does not seem to be currently
1868: justified. While cognizant of the above deficiencies,
1869: we wish to believe that they are outweighted by the model merits.
1870:
1871:
1872: \begin{thebibliography}{300}
1873: %--------------------------------------------------------------
1874: \bibitem
1875: {GMO} M. L. Goldberger, H. Miyazawa and R. Oehme,
1876: Phys. Rev. {\bf 99}, 986 (1955).
1877: %--------------------------------------------------------------
1878: \bibitem {SAID}
1879: R. A. Arndt, M. M. Pavan,
1880: R. L. Workman, and I. I. Strakovsky,
1881: Scattering Interactive
1882: Dial-Up (SAID), VPI, Blacksburg,
1883: The VPI/GWU $\pi $N solution SM99 (1999);
1884: http://said.phys.vt.edu/analysis/pin\_analysis.html.
1885: %--------------------------------------------------------------
1886: \bibitem {GIB98}
1887: W. R. Gibbs, Li Ai and W. B. Kaufmann, Phys. Rev.
1888: C {\bf 57}, 784 (1998).
1889: %--------------------------------------------------------------
1890: \bibitem{Gashi}
1891: A. Gashi, E. Matsinos, G. C. Oades, G. Rasche and W. S. Woolcock,
1892: Nucl. Phys A (to be published), hep-ph/0009081.
1893: %--------------------------------------------------------------
1894: \bibitem {Li}
1895: W.R. Gibbs, Li Ai and W.B. Kaufmann, {\it Phys. Rev.
1896: Lett.} {\bf 74}, 3740 (1995);
1897: $\pi$N Newsletter No 11, Vol II (1995), p 84.
1898: %--------------------------------------------------------------
1899: \bibitem {MAT97}
1900: E. Matsinos, Phys. Rev. C {\bf 56}, 3014 (1997).
1901: %--------------------------------------------------------------
1902: \bibitem {Sigg}
1903: D. Sigg, A. Badertscher, P. F. A. Goudsmit, H. J. L. Leisi
1904: and G. C. Oades, Nucl. Phys. A {\bf 609}, 310 (1996).
1905: %--------------------------------------------------------------
1906: \bibitem{SCH99}
1907: H.~C.~Schr\"oder, A. Badertscher, P.F.A. Goudsmit,
1908: M. Janousch, H.J. Leisi, E. Matsinos, D. Sigg,
1909: Z.G. Zhao, D. Chatellard, J.P. Egger {\it et al.},
1910: Phys.\ Lett.\ B {\bf 469}, 25 (1999).
1911: %--------------------------------------------------------------
1912: \bibitem {HAU98}
1913: P. Hauser, K. Kirch, L. M. Simons, G. Borchert, D.
1914: Gotta, T. Siems, P. El-Khoury, P. Indelicato, M. Augsburger,
1915: D.~Chatellard {\it et al.}, Phys. Rev. C {\bf 58}, R1869 (1998).
1916: %------------------------------------------------------------------------
1917: \bibitem{Nadia}
1918: N. Fettes, Ulf-G. Meissner and S. Steininger,
1919: Nucl. Phys. A {\bf 640}, 199 (1998);\\
1920: V.E. Lyubitskij and A. Russetsky, Phys. Lett. B {\bf 494}, 9 (2000).
1921: %--------------------------------------------------------------
1922: \bibitem{Deser}
1923: S.~Deser, M.L.~Goldberger, K.~Baumann, and W.~Thirring, Phys.~Rev.
1924: {\bf 96}, 774 (1954);
1925: T.~L.~Trueman, Nucl.~Phys.~{\bf 26}, 57 (1961).
1926: %--------------------------------------------------------------
1927: \bibitem {BARU}
1928: V. V. Baru and A. E. Kudryavtsev, Phys. Atom. Nucl. {\bf60}, 1475
1929: (1997);\\
1930: T.E.O. Ericson, B. Loiseau and A.W. Thomas, hep-ph/0009312.
1931: %--------------------------------------------------------------
1932: \bibitem{PETROV}
1933: V. V. Peresypkin and N. M. Petrov, Nucl. Phys. {\bf A 220}, 277
1934: (1974);\\
1935: N. M. Petrov and V. V. Peresypkin, J. Nucl. Phys. (USSR) {\bf 18}, 791 (1973).
1936: %--------------------------------------------------------------
1937: \bibitem {AFN74}
1938: I. R. Afnan and A. W. Thomas, Phys. Rev. C {\bf 10}, 109 (1974).
1939: %--------------------------------------------------------------
1940: \bibitem {MIZ77}
1941: T. Mizutani and D. Koltun, Ann. Phys. (NY) {\bf 109}, 1 (1977).
1942: %--------------------------------------------------------------
1943: \bibitem {Judah}
1944: J.M. Eisenberg and D.S. Koltun, {\it Theory of Meson Interactions with
1945: Nuclei}, (Wiley, New York 1980).
1946: %--------------------------------------------------------------
1947: \bibitem{THO80}
1948: A. W. Thomas and R. H. Landau, Phys. Rep. {\bf 58}, 121 (1980).
1949: %--------------------------------------------------------------
1950: \bibitem {ERI88}
1951: T. E. O. Ericson and W. Weise, {\it Pions and Nuclei},
1952: (Clarendon, Oxford, 1988).
1953: %--------------------------------------------------------------
1954: \bibitem{BRUCK}
1955: K. A. Brueckner, Phys. Rev. {\bf 89}, 834 (1953).
1956: %--------------------------------------------------------------
1957: \bibitem{HUANG}
1958: Kerson Huang, {\em Statistical Mechanics}, (Wiley, New York 1963), Chap.
1959: XIII.
1960: %--------------------------------------------------------------
1961: \bibitem {Victor}
1962: V. M. Kolybasov and A. E. Kudryavtsev, Sov. Phys. JETP
1963: {\bf 36}, 18 (1973).
1964: %--------------------------------------------------------------
1965: \bibitem{PEST}
1966: H. Zankel, W. Plessas and J. Haidenbauer, Phys. Rev. C {\bf 28}, 538 (1983).
1967: %--------------------------------------------------------------
1968: \bibitem{Paris}
1969: M. Lacombe, B. Loiseau, R. Vinh Mau, J. C\^ot\'e, P.
1970: Pir\'es, and R. de Tourreil, Phys. Lett. B {\bf 101}, 139 (1981).
1971: %--------------------------------------------------------------
1972: \bibitem {FAL77}
1973: G. F\"{a}ldt, Physica Scripta {\bf 16}, 81 (1977).
1974: %--------------------------------------------------------------
1975: \bibitem{Bonn}
1976: R. Machleidt, K. Holinde and Ch. Elster,
1977: Phys. Rep. {\bf 149}, 1 (1987), tables 11 and 13.
1978: %--------------------------------------------------------------
1979: \bibitem{Abramowitz}
1980: {\em Handbook of Mathematical Functions}, edited by M. Abramowitz and
1981: I.A. Stegun (Dover, New York, 1965).
1982: %--------------------------------------------------------------
1983: \bibitem{GIB86}
1984: P.B. Siegel and W. R. Gibbs, Phys. Rev. C {\bf 33}, 1407 (1986).
1985: %--------------------------------------------------------------
1986: \bibitem{Panofsky}
1987: J. Spuller, {\it et al.}, Pys. Lett. B {\bf 67}, 479 (1977).
1988: %--------------------------------------------------------------
1989: \bibitem{HENDRY}
1990: W. R. Frazer and A. W. Hendry, Phys. Rev. {\bf 134 B}, 1307 (1964).
1991: %--------------------------------------------------------------
1992: \bibitem{Gauss}
1993: W.~B.~Jones and W.~J.~Thorn, {\em Continued fractions: Analytic Theory
1994: and Applications}, (Addison-Wesley, Reading, Mass. 1980).
1995: %------------------------------------------------------------------------
1996: \bibitem{vanH}
1997: H. van Haeringen, Nucl. Phys. A {\bf 253}, 355, (1975).
1998: %--------------------------------------------------------------
1999: \bibitem{SE}
2000: L.L. Foldy and J.D. Walecka, Ann. Phys. (N.Y.) {\bf 54}, 447 (1969); \\
2001: J.H. Koch and J.D. Walecka, Nucl. Phys. B {\bf 72}, 283 (1974); \\
2002: F.A. Gareev, M.Ch. Gizzatzulov and J. Revai, Nucl. Phys. A {\bf 286},
2003: 512 (1977).
2004: %--------------------------------------------------------------
2005: \end{thebibliography}
2006:
2007: \newpage
2008:
2009: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% tables %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2010: \begin{table}
2011: \caption{
2012: $\pi$d scattering length
2013: obtained from the static model and from a Faddeev calculation
2014: in the zero-range model for different $b_{0}$ and $b_{1}$.
2015: For the NN forces we used PEST and Yamaguchi potentials,
2016: the results for the latter case are presented here in brackets.
2017: All entries are in $10^{-2}/m_{\pi}$ units.}
2018: \label{table1}
2019: \vspace*{2mm}
2020: \begin{tabular}{cccccc}
2021: \multicolumn{6}{c}{}\\
2022: & & & & $b_{1}$ & \\
2023: & & & & & \\
2024: \hline
2025: & & & & & \\
2026: & $b_{0}$ & model & -9.47 & -9.05 & -8.63\\
2027: & & & & & \\
2028: \hline
2029: & & 2-nd order & -4.22 (-4.87) &-3.97 (-4.57) &-3.74 (-4.28)\\
2030: & &static \Ref{a10} & -3.89 (-4.21) &-3.69 (-3.98) &-3.49 (-3.77)\\
2031: &-0.65&static \Ref{a26} & -3.44 (-3.77) &-3.29 (-3.58) &-3.10 (-3.39)\\
2032: & & Faddeev & -3.97 (-4.27) &-3.76 (-4.04) &-3.55 (-3.81)\\
2033: & &{\em ditto} with $\Delta$ & -3.59 (-3.97) &-3.37 (-3.73) &-3.16 (-3.50)\\
2034: \hline
2035: & & 2-nd order & -3.30 (-3.96) &-3.06 (-3.66) &-2.82 (-3.37)\\
2036: & &static \Ref{a10} & -2.99 (-3.32) &-2.78 (-3.09) &-2.58 (-2.87)\\
2037: &-0.22&static \Ref{a26} & -2.53 (-2.87) &-2.36 (-2.68) &-2.19 (-2.49)\\
2038: & & Faddeev & -3.07 (-3.37) &-2.85 (-3.14) &-2.65 (-2.92)\\
2039: & &{\em ditto} with $\Delta$ & -2.68 (-3.08) &-2.46 (-2.85) &-2.25 (-2.62)\\
2040: \hline
2041: & & 2-nd order & -2.38 (-3.04) &-2.14 (-2.74) &-1.90 (-2.45)\\
2042: & &static \Ref{a10} & -2.08 (-2.42) &-1.87 (-2.19) &-1.68 (-1.97)\\
2043: & 0.21 &static \Ref{a26} & -2.62 (-1.97) &-1.45 (-1.77) &-1.28 (-1.59)\\
2044: & & Faddeev & -2.16 (-2.47) &-1.95 (-2.24) &-1.74 (-2.02)\\
2045: & &{\em ditto} with $\Delta$ & -1.76 (-2.20) &-1.54 (-1.96) &-1.34 (-1.73)\\
2046: \multicolumn{6}{c}{}\\
2047: \end{tabular}
2048: \end{table}
2049: %%%%%%%%%%%%%%% end %%%%%%%%%%%%%%%%%%%%%%%%%%%%
2050:
2051: %%%%%%%%%%% next table
2052: \begin{table}
2053: \caption{
2054: The expectation values of r, 1/r
2055: and the values of $\pi$d scattering length
2056: calculated for different NN wavefunctions.
2057: For $\pi$N scattering lengths we have adopted their central values, i.e.
2058: $b_{0}=-0.22$ and $b_{1}=-9.05$.
2059: All scattering lengths are given in $10^{-2}/m_{\pi}$ units.
2060: }
2061: \label{table2}
2062: \vspace*{2mm}
2063: \begin{tabular}{cccccc}
2064: \multicolumn{6}{c}{}\\
2065: & & & NN wavefunction & & \\
2066: &&&&&\\
2067: \hline
2068: & & Hulthen & PEST & Paris & Bonn \\
2069: \hline
2070: &$\bra r \ket fm$ & 3.1345 & 3.2309 & 3.2685 & 3.2536 \\
2071: &$\bra 1/r \ket fm^{-1}$ & 0.55501 & 0.45507 & 0.44864 & 0.46314 \\
2072: & 2-nd order $a_{\pi d}$ & -3.66 & -3.06 & -3.04 & -3.13 \\
2073: & static $a_{\pi d}$ & -3.09 & -2.78 & -2.78 & -2.82 \\
2074: \multicolumn{6}{c}{}\\
2075: \end{tabular}
2076: \end{table}
2077: %%%%%%%%%%%%%%% end %%%%%%%%%%%%%%%%%%%%%%%%%%%%
2078:
2079: %%%%%%%%%%% next table
2080: \begin{table}
2081: \caption{
2082: $\pi$d scattering lengths calculated from
2083: consecutive iterations of the Faddeev equations.
2084: All entries are in $10^{-2}/m_{\pi}$ units.
2085: }
2086: \label{table3}
2087: \vspace*{2mm}
2088: \begin{tabular}{cccccc}
2089: \multicolumn{6}{c}{}\\
2090: & & PEST&PEST & Yamaguchi&Yamaguchi\\
2091: & order & no $\Delta$ & with $\Delta$ & no $\Delta$ & with $\Delta $\\
2092: &&&&&\\
2093: \hline
2094: & 1 & -1.66 & -1.21 & -1.70 & -1.23\\
2095: & 2 & -2.98 & -2.66 & -3.42 & -3.30\\
2096: & 3 & -2.89 & -2.44 & -3.20 & -2.77\\
2097: & 4 & -2.85 & -2.48 & -3.11 & -2.91\\
2098: & 5 & -2.85 & -2.45 & -3.14 & -2.82\\
2099: & 6 & & -2.46 & -3.15 & -2.87\\
2100: & 7 & & -2.46 & -3.14 & -2.84\\
2101: & 8 & & & -3.14 & -2.86\\
2102: & 9 & & & & -2.85\\
2103: & 10 & & & & -2.85\\
2104: \multicolumn{6}{c}{}\\
2105: \end{tabular}
2106: \end{table}
2107: %%%%%%%%%%%%%%% end %%%%%%%%%%%%%%%%%%%%%%%%%%%%
2108:
2109: %%%%%%%%%%% next table
2110: \begin{table}
2111: \caption{ $\pi$N scattering lengths inferred from pionic
2112: hydrogen data ($B_{2I}$ are slope parameters defined in the text).
2113: }
2114: \label{table4}
2115: \vspace*{2mm}
2116: \begin{tabular}{cccccc}
2117: \multicolumn{6}{c}{}\\
2118: & $\beta$ & a $_{1}$ & a$_{3}$ & B$_{1}$ & B$_{3}$ \\
2119: & [fm$^{-1}$]&[m$_{\pi}^{-1}$]&[10$^{-1}\,$m$_{\pi}^{-1}$]&
2120: [10$^{-2}\,$m$_{\pi}^{-3}]$& [10$^{-2}\,$m$_{\pi}^{-3}$]\\
2121: \hline
2122: & 2.0& 0.1767$\pm$ 0.0046& -0.9377$\pm$ 0.0852
2123: & -6.63 & 1.96\\
2124: & 3.0& 0.1760$\pm$ 0.0046& -0.9306$\pm$ 0.0846
2125: & -3.60 & 0.81\\
2126: & 4.0& 0.1757$\pm$ 0.0046& -0.9263$\pm$ 0.0841
2127: & -2.46 & 0.43\\
2128: & 5.0& 0.1756$\pm$ 0.0046& -0.9228$\pm$ 0.0837
2129: & -1.90 & 0.27\\
2130: & 6.0& 0.1756$\pm$ 0.0046& -0.9197$\pm$ 0.0834
2131: & -1.57 & 0.18\\
2132: & 7.0& 0.1756$\pm$ 0.0046& -0.9167$\pm$ 0.0830
2133: & -1.37 & 0.14\\
2134: & 8.0& 0.1756$\pm$ 0.0046& -0.9138$\pm$ 0.0827
2135: & -1.23 & 0.11\\
2136: & 9.0& 0.1756$\pm$ 0.0047& -0.9110$\pm$ 0.0823
2137: & -1.12 & 0.09\\
2138: & 10.0& 0.1757$\pm$ 0.0047& -0.9082$\pm$ 0.0820
2139: & -1.05 & 0.08\\
2140: \hline
2141: & Deser & 0.1760 $\pm$ 0.0046&-0.9258 $\pm$0.0857& & \\
2142: \hline
2143: & ref. \cite{Gashi}& 0.1679 $\pm$ 0.0059 &-0.785 $\pm$ 0.034 &
2144: -7.24 $\pm$ 3.06 & -4.08 $\pm$ 1.46\\
2145: \multicolumn{6}{c}{}\\
2146: \end{tabular}
2147: \end{table}
2148: %%%%%%%%%%%%%%% end tables %%%%%%%%%%%%%%%%%%%%%%%%%%%%
2149:
2150: \newpage
2151:
2152: \begin{figure}[ht]
2153: %\setlength{\belowcaptionskip}{20pt}
2154: \centering
2155: \caption{
2156: Constrains on the isoscalar and isovector scattering lengths imposed by
2157: pionic deuterium data. The black strip corresponds the one standard
2158: deviation region. The rectangle corresponds to the values
2159: obtained from pionic hydrogen data in \protect\cite{Sigg}.
2160: }
2161: \label{fig1}
2162: \vspace*{8mm}
2163:
2164: %\includegraphics[width=0.75\textwidth,totalheight=0.5\textheight]{f2.eps}
2165: \includegraphics[width=0.5\textwidth,totalheight=0.5\textheight]{f2.eps}
2166: \end{figure}
2167: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2168: \newpage
2169: \begin{figure}[ht]
2170: %\setlength{\belowcaptionskip}{20pt}
2171: \centering
2172: \caption{$\sin \delta$ vs energy close to the resonance
2173: calculated from \Ref{b16} for $\beta=3\,fm^{-1}$.
2174: }
2175: \label{fig2}
2176: \vspace*{8mm}
2177:
2178: %\includegraphics[width=0.75\textwidth,totalheight=0.5\textheight]{sin.eps}
2179: \includegraphics[width=0.5\textwidth,totalheight=0.5\textheight]{sin.eps}
2180: \end{figure}
2181: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2182: \newpage
2183: \begin{figure}[ht]
2184: %\setlength{\belowcaptionskip}{20pt}
2185: \centering
2186: \caption{
2187: The values of isoscalar and isovector scattering lengths obtained
2188: by solving the bound state equation \Ref{b5}
2189: for $\beta$ equal, respectvely, 2.0 fm$^{-1}$, 6.0 fm$^{-1}$,
2190: and 10.0 fm$^{-1}$ (indicated on the plot).
2191: The point marked as Deser has been obtained from \Ref{b31}.
2192: The rectangle corresponds to the values obtained in
2193: \protect\cite{Sigg}.
2194: }
2195: \label{fig3}
2196: \vspace*{8mm}
2197:
2198: %\includegraphics[width=0.75\textwidth ,totalheight=0.5\textheight ]{f3.eps}
2199: \includegraphics[width=0.4\textwidth ,totalheight=0.4\textheight ]{f3.eps}
2200: \end{figure}
2201: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2202: \newpage
2203: \begin{figure}[ht]
2204: %\setlength{\belowcaptionskip}{20pt}
2205: \centering
2206: \caption{ $\pi$d scattering length vs. the inverse range parameter
2207: $\beta$ of the $\pi$N potential. Full (open) circles correspond
2208: to a Faddeev calculation with (without) p-wave $\pi$N interaction.
2209: }
2210: \label{fig4}
2211: \vspace*{8mm}
2212:
2213: %\includegraphics[width=0.75\textwidth,totalheight=0.5\textheight]{apid.eps}
2214: \includegraphics[width=0.4\textwidth,totalheight=0.4\textheight]{apid.eps}
2215: \end{figure}
2216:
2217:
2218: %%%%%%%%%%%%%%%%%%%%%%% END %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2219:
2220: \end{document}
2221:
2222: