physics0102022/fm.tex
1: 
2: \documentstyle[prc,aps,psfig,amsmath]{revtex}
3: %,preprint,tighten,
4: \begin{document}
5: 
6: 
7: 
8: \draft
9: 
10: \twocolumn [ \hsize\textwidth\columnwidth\hsize\csname
11: @twocolumnfalse\endcsname
12: 
13: \title{Three-potential formalism for the three-body 
14: scattering problem with attractive Coulomb interactions}
15: 
16: \author{Z.\ Papp${}^{1,2}$, C-.Y.\ Hu${}^{1}$,  Z.\ T.\ Hlousek${}^{1}$, 
17: B.~K\'onya${}^{2}$ and S.\ L.\ Yakovlev${}^{3}$ }
18: \address{${}^{1}$ Department of Physics and Astronomy, 
19: California State University, Long Beach, CA 90840, USA \\
20: ${}^{2}$ Institute of Nuclear Research of the
21: Hungarian Academy of Sciences, Debrecen, Hungary \\
22: ${}^{3}$ Department of Mathematical and Computational Physics, 
23: St.\ Petersburg State University, St.\ Petersburg,  Russia }
24: \date{\today}
25: \maketitle
26: 
27: \begin{abstract}
28: \noindent
29: A three-body scattering process in the presence of Coulomb interaction
30: can be decomposed formally into a two-body single channel,
31: a two-body multichannel and a genuine three-body scattering.
32: The corresponding integral equations are coupled Lippmann-Schwinger and 
33: Faddeev-Merkuriev integral equations. We solve them by applying 
34: the Coulomb-Sturmian separable expansion method.  We present elastic 
35: scattering and reaction cross sections of the $e^++H$ system both below and
36: above the $H(n=2)$ threshold. We found excellent agreements with previous
37: calculations in most cases.
38: \end{abstract}
39: 
40: \vspace{0.5cm} 
41: \pacs{PACS number(s): 31.15.-p, 34.10.+x, 34.85.+x, 21.45.+v, 03.65.Nk, 
42: 02.30.Rz, 02.60.Nm}
43: ]
44: 
45: 
46: The three-body Coulomb scattering
47: problem is one of the most challenging long-standing problems
48: of non-relativistic quantum mechanics. The source of the difficulties
49: is related to the long-range character of the Coulomb potential.
50: In the standard scattering theory it is
51: supposed that the particles move freely  asymptotically. That is
52: not the case if Coulombic interactions are involved.
53: As a result the fundamental equations of the three-body
54: problems, the Faddeev-equations, become ill-behaved if they are applied
55: for Coulomb potentials in a straightforward manner.
56: 
57: The first, and formally exact, approach was proposed by Noble \cite{noble}. 
58: His formulation was designed for solving the nuclear three-body
59: Coulomb problem, where all Coulomb interactions are repulsive.
60: The interactions were split into short-range and long-range Coulomb-like
61: parts and the long-range parts were formally included in the 
62: "free" Green's operator.
63: Therefore the corresponding Faddeev-Noble equations become mathematically
64: well-behaved and in the absence of Coulomb interaction 
65: they fall back to the standard equations. However, the associated 
66: Green's operator is not known. This formalism, as presented
67: at that time, was not suitable for practical calculations.
68: 
69: In Noble's approach the separation of the Coulomb-like potential into
70: short-range and long-range parts were carried out in the 
71: two-body configuration space.
72: Merkuriev extended the idea of Noble by performing the 
73: splitting in the three-body configuration space.
74: This was a crucial development since it made possible to treat attractive
75: Coulomb interactions on an equal footing as repulsive ones. 
76: This theory has been developed using integral equations with
77: connected (compact) kernels and transformed 
78: into  configuration-space differential
79: equations with asymptotic boundary conditions \cite{fm-book}. 
80: In practical calculations, so far 
81: only the latter version of the theory has been considered. The primary
82: reason is that the more complicated structure of the Green's operators 
83: in the kernels of the
84: Faddeev-Merkuriev integral equations has not yet allowed any direct solution.
85: However, use of integral equations is a very appealing approach since
86: no boundary conditions are required.
87: 
88: Recently, one of us has developed a novel method 
89: for treating the three-body problem with repulsive
90: Coulomb interactions in three-potential picture
91: \cite{pzsc}. In this approach a three-body Coulomb scattering
92: process can be decomposed formally into a two-body single channel,
93: a two-body multichannel and a genuine three-body scattering.
94: The corresponding integral equations are coupled Lippmann-Schwinger
95: and  Faddeev-Noble integral equations, which were solved by  using
96: the Coulomb-Sturmian separable expansion method.
97: The approach was tested first
98: for bound-state problems \cite{pzwp} with repulsive Coulomb plus nuclear
99: potential. Then it was extended to calculate $p-d$
100: scattering at energies below the breakup threshold \cite{pzsc} and
101: more recently we have used the method  to calculate 
102: resonances  of three-$\alpha$ systems \cite{zis}. 
103: Also atomic  bound-state problems with attractive Coulomb
104: interactions have been considered \cite{pzatom}.
105: These calculations showed an excellent agreement with the results of other 
106: well established methods. The efficiency and the accuracy 
107: of the method was demonstrated. 
108: 
109: The aim of this paper is to generalize this method for solving the three-body
110: Coulomb problem with repulsive and attractive Coulomb interactions.
111: We combine the concept of three-potential formalism with the Merkuriev's
112: splitting of the interactions and solve the resulting set of 
113: Lippmann-Schwinger and Faddeev-Merkuriev integral equations by applying the 
114: Coulomb-Sturmian separable expansion method. In this paper we restrict 
115: ourselves to energies below the three-body breakup threshold. 
116: 
117: 
118: \section{Integral equations of the three-potential picture}
119: 
120: 
121: We consider a three-body system with  Hamiltonian 
122: \begin{equation}
123: H=H^0 + v_\alpha^C + v_\beta^C + v_\gamma^C,
124: \label{H}
125: \end{equation}
126: where $H^0$ is the three-body kinetic energy 
127: operator and $v_\alpha^C$ denotes the Coulomb-like
128: interaction in subsystem $\alpha$. The potential
129: $v_\alpha^C$ may have repulsive
130: or attractive Coulomb tail and any short-range component.
131: We use the usual
132: configuration-space Jacobi coordinates
133:  $x_\alpha$ and $y_\alpha$; $x_\alpha$ is the coordinate
134: between the pair $(\beta,\gamma)$ and $y_\alpha$ is the
135: coordinate between the particle $\alpha$ and the center of mass
136: of the pair $(\beta,\gamma)$.
137: Thus the potential $v_\alpha^C$, the interaction of the
138: pair $(\beta,\gamma)$, appears as $v_\alpha^C (x_\alpha)$.
139: We also use the notation $X=\{x_\alpha,y_\alpha \}\in{\bf R}^6$.
140: 
141: 
142: \subsection{Merkuriev's cut of the Coulomb potential}
143: 
144: 
145: The  Hamiltonian (\ref{H}) is defined in the three-body 
146: Hilbert space. The two-body potential operators are formally
147: embedded in the three-body Hilbert space
148: \begin{equation}
149: v^C = v^C (x) {\bf 1}_{y}.
150: \label{pot0}
151: \end{equation}
152: Merkuriev introduced a separation of the three-body
153: configuration space into different
154: asymptotic regions. The two-body asymptotic region $\Omega_\alpha$ is
155: defined as a part of the three-body configuration space where
156: the conditions
157: \begin{equation}
158: |x_\alpha| <  x^0_\alpha ( 1  +
159: |y_\alpha|/ y^0_\alpha)^{1/\nu},
160: \label{oma}
161: \end{equation}
162: with $x^0_\alpha, y^0_\alpha >0$ and $\nu > 2$, are satisfied.
163: He proposed to split the Coulomb interaction  in 
164: the three-body configuration space into
165: short-range and long-range terms 
166: \begin{equation}
167: v_\alpha^C =v_\alpha^{(s)} +v_\alpha^{(l)} ,
168: \label{pot}
169: \end{equation}
170: where the superscripts
171: $s$ and $l$ indicate the short- and long-range
172: attributes, respectively. 
173: The splitting is carried out with the help of a splitting function $\zeta$,
174: \begin{eqnarray}
175: v^{(s)} (x,y) & = & v^C(x) \zeta (x,y),
176: \\
177: v^{(l)} (x,y) & = & v^C(x) [1- \zeta (x,y) ].
178: \label{potl}
179: \end{eqnarray}
180: The function $\zeta$ is defined such that 
181: \begin{equation}
182: \zeta(x,y) \xrightarrow{X \to \infty}
183: \left\{ 
184: \begin{array}{ll}
185: 1, &  X \in \Omega_\alpha \\
186: 0  & \mbox{otherwise.}
187: \end{array}
188: \right.
189: \end{equation}
190: In practice, in the configuration-space differential equation
191: approaches, usually the functional form
192: \begin{equation}
193: \zeta (x,y) =  2/\left\{1+ \left[ {(x/x^0)^\nu}/{(1+y/y^0)} \right] \right\},
194: \label{oma1}
195: \end{equation}
196: was used.
197: 
198: The long-range Hamiltonian is defined as
199: \begin{equation}
200: H^{(l)} = H^0 + v_\alpha^{(l)}+ v_\beta^{(l)}+ v_\gamma^{(l)},
201: \label{hl}
202: \end{equation}
203: and its resolvent operator is
204: \begin{equation}
205: G^{(l)}(z)=(z-H^{(l)})^{-1}.
206: \end{equation}
207: Then, the three-body Hamiltonian takes the form
208: \begin{equation}
209: H = H^{(l)} + v_\alpha^{(s)}+ 
210:  v_\beta^{(s)}+ v_\gamma^{(s)}.
211: \label{hll}
212: \end{equation}
213: 
214: In the conventional Faddeev theory the wave function 
215: components are defined by
216: \begin{equation}
217: | \psi_\alpha \rangle  = (z-H^0)^{-1} v_\alpha | \Psi \rangle,
218: \label{fagyi}
219: \end{equation}
220: where $v_\alpha$ is a short-range potential 
221: and $| \psi_\alpha \rangle$ is the Faddeev
222: component of the total wave function $| \Psi \rangle$. 
223: While the total wave function $| \Psi \rangle$,
224: in general, has three different kind of two-body asymptotic channels,
225: $| \psi_\alpha \rangle$ possesses only 
226: $\alpha$-type two-body asymptotic channel.
227: The other channels are suppressed by the short-range potential
228: $v_\alpha$. This procedure is called asymptotic filtering and it guarantees
229: the asymptotic orthogonality of the Faddeev components  \cite{vanzani}.
230: 
231: The aim of the Merkuriev procedure was to formally obtain a three-body 
232: Hamiltonian with short-range potentials $v^{(s)}$ and long-range
233: Hamiltonian $H^{(l)}$ in order that we can repeat the procedure of 
234: the conventional Faddeev theory.
235: The total wave function 
236: $|\Psi  \rangle $ is split into three components,
237: \begin{equation}
238: |\Psi  \rangle = |\psi_{\alpha} \rangle + |\psi_{\beta} \rangle +
239: |\psi_{\gamma} \rangle,
240: \end{equation}
241: with components defined by 
242: \begin{equation}
243: |\psi_{\alpha} \rangle= G^{(l)} v_\alpha^{(s)} |\Psi  \rangle.
244: \label{fdec}
245: \end{equation}
246: This procedure is an example of asymptotic filtering. The short-range potential
247:  $v_\alpha^{(s)}$ acting on $|\Psi  \rangle$ 
248: suppresses the possible $\beta$ and $\gamma$ asymptotic two-body
249: channels, provided $G^{(l)}$ itself does not introduce any new two-body
250: asymptotic channels.  With the Merkuriev splitting this is avoided because
251: $H^{(l)}$ does not have two-body asymptotic channels even if some 
252: of the long-range potentials have attractive Coulomb tail. 
253: In the attractive case  $v^{(l)}$ appears as
254: a valley along the $y=x^\nu$ parabola-like
255: curve with Coulomb-like asymptotic behavior in
256: $x$ at any finite $y$. (See Figs.\ \ref{vs} and \ref{vl} for the short- 
257: and long-range parts, respectively). 
258: However, as $y \to \infty$ the depth of the
259: valley goes to zero, consequently the two-body bound states are pushed up, and 
260: finally the system does not have any two-body asymptotic channels.
261: We note that the Merkuriev formalism contains the Noble's  in the limit 
262: $y^0\to\infty$. 
263: 
264: 
265: \subsection{The three-potential picture }
266: 
267: In Ref.\ \cite{pzsc} the three body scattering problem with repulsive 
268: Coulomb interactions were considered
269: in the three-potential picture. In this picture the 
270: scattering process  can be decomposed formally
271: into three consecutive scattering processes:
272: a two-body single channel, a two-body multichannel and a genuine three-body
273: scattering. This formalism also provides the integral equations and the 
274: method of constructing the $S$-matrix.
275: Below we adapt this formalism to attractive Coulomb interactions along the
276: Merkuriev approach.
277: 
278: The asymptotic Hamiltonian is defined as 
279: \begin{equation}
280: H_\alpha=H^0 +  v_\alpha^C,
281: \end{equation}
282: and the asymptotic states are the eigenstates of $H_\alpha$ 
283: \begin{equation}
284: H_\alpha | \Phi_{\alpha} \rangle = E | \Phi_{\alpha} \rangle,
285: \end{equation}
286: where $\langle x_\alpha y_\alpha |
287:  \Phi_{\alpha} \rangle = \langle
288: y_\alpha| \chi_{\alpha} \rangle \langle 
289: x_\alpha | \phi_{\alpha} \rangle$ 
290: is a product of a scattering state in coordinate
291: $y_\alpha$ and a bound state in the two-body subsystem $x_\alpha$.
292: 
293: We define the two asymptotic long-range Hamiltonians as 
294: \begin{equation}
295: H_\alpha^{(l)}=H^0 + v_\alpha^{C}  + v_\beta^{(l)} +v_\gamma^{(l)}
296: \end{equation}
297: and 
298: \begin{equation}
299: \widetilde{H}_\alpha=H^0 + v_\alpha^{C} + u_\alpha^{(l)},
300: \label{htilde}
301: \end{equation}
302: where $u_\alpha^{(l)}$ is an auxiliary potential in 
303: coordinate $y_\alpha$,
304: and it is required to have the asymptotic form 
305: \begin{equation}
306: u_\alpha^{(l)} \sim  {Z_\alpha (Z_\beta+Z_\gamma) }/{y_\alpha}
307: \end{equation}
308: as ${y_\alpha \to \infty}$. In fact, $u_\alpha^{(l)}$ 
309: is an effective Coulomb-like
310: interaction between the center of mass of the subsystem 
311: $\alpha$ (with
312: charge $Z_\beta+Z_\gamma$) and the third particle 
313: (with charge $Z_\alpha$). We introduced this potential in order that
314: we compensate the long range Coulomb
315: tail of $v_\beta^{(l)} +v_\gamma^{(l)}$ in $\Omega_\alpha$.
316: 
317: Let us introduce the resolvent operators: 
318: \begin{equation}
319: G(z)=(z-H)^{-1},
320: \end{equation}
321: \begin{equation}
322: G_\alpha ^{(l)}(z)=(z-H_\alpha^{(l)})^{-1},
323: \end{equation}
324: \begin{equation}
325: \widetilde{G}_\alpha (z)=(z-\widetilde{H}_\alpha )^{-1}.
326: \end{equation}
327: The operator $G_\alpha^{(l)}$ is the long-range channel 
328: Green's operator
329: and $\widetilde{G}_\alpha $ is the channel 
330: distorted long-range Green's
331: operator. 
332: These operators are connected via
333: the following resolvent relations: 
334: \begin{equation}
335: G(z)=G_\alpha^{(l)}(z)+G_\alpha^{(l)}(z) V^\alpha G(z),  \label{g3b}
336: \end{equation}
337: \begin{equation}
338: G_\alpha^{(l)}(z)=\widetilde{G}_\alpha (z)+
339: \widetilde{G}_\alpha (z)U^\alpha
340: G_\alpha^{(l)}(z),  \label{g2b}
341: \end{equation}
342: where $V^\alpha =v^{(s)}_\beta +v^{(s)}_\gamma $ and 
343: $U^\alpha =v_\beta^{(l)}+v_\gamma^{(l)}-u_\alpha ^{(l)}$.
344: 
345: The scattering state, which evolves from the asymptotic
346: state $|\Phi_\alpha \rangle$ under the influence of
347: $H$, is given as
348: \begin{equation}
349: | \Psi_\alpha^{(\pm)} \rangle = 
350: \lim\limits_{\varepsilon\to 0} \mbox{i}
351: \varepsilon G (E_\alpha \pm \mbox{i}\varepsilon)| 
352: \Phi_\alpha \rangle.
353: \label{Psidef}
354: \end{equation}
355: Similarly, we can define the following auxiliary scattering states
356: \begin{equation}
357: | \Phi_\alpha^{{(l)}(\pm)} \rangle = 
358: \lim\limits_{\varepsilon\to 0} \mbox{i}
359: \varepsilon G_\alpha^{(l)} (E\pm \mbox{i}\varepsilon) |
360:  \Phi_\alpha \rangle \label{phil}
361: \end{equation}
362: and
363: \begin{equation}
364: | \widetilde{\Phi}_{\alpha}^{(\pm)}\rangle = 
365: \lim\limits_{\varepsilon\to 0} 
366: \mbox{i}\varepsilon \widetilde{G}_\alpha
367:  (E\pm \mbox{i}\varepsilon) |
368: \Phi_{\alpha} \rangle,  \label{phitild}
369: \end{equation}
370: which describe scattering processes due to Hamiltonians
371: $H^{(l)}_\alpha$ and $\widetilde{H}_\alpha$, respectively.
372: 
373: The S-matrix elements of scattering processes are obtained from the
374: resolvent of the total Hamiltonian by the reduction technique \cite{redu} 
375: \begin{equation}
376: S_{\beta j,\alpha i}=
377: \lim\limits_{t\to \infty }\lim\limits_{\varepsilon \to
378: 0}\mbox{i}\varepsilon \mbox{e}^{\mbox{i}(E_{\beta j}-
379: E_{\alpha i})t}\langle
380: \Phi _{\beta j}|G(E_{\alpha i}+\mbox{i}\varepsilon )|
381: \Phi _{\alpha i}\rangle .  \label{sm}
382: \end{equation}
383: The subscript $i$ and $j$ denotes the $i$-th and 
384: $j$-th eigenstates of the
385: corresponding subsystems, respectively. 
386: If we substitute (\ref{g3b}) into (\ref{sm}) we  get
387: the following two terms:
388: \begin{equation}
389: S_{\beta j,\alpha i}^{(1,2)}=\lim\limits_{t\to \infty
390: }\lim\limits_{\varepsilon \to 0}\mbox{i}\varepsilon 
391: \mbox{e}^{\mbox{i}
392: (E_{\beta j}-E_{\alpha i})t}\langle \Phi _{\beta j}|
393: G_\alpha^{(l)}(E_{\alpha i}+
394: \mbox{i}\varepsilon )|\Phi _{\alpha i}\rangle   \label{s12}
395: \end{equation}
396: \begin{eqnarray}
397: S_{\beta j,\alpha i}^{(3)}&=&\lim\limits_{t\to \infty
398: }\lim\limits_{\varepsilon \to 0}\mbox{i}\varepsilon \mbox{e}^{\mbox{i}
399: (E_{\beta j}-E_{\alpha i})t}\langle \Phi _{\beta j}|
400: G_\alpha^{(l)}(E_{\alpha i}+\mbox{i}\varepsilon ) \nonumber \\
401: && V^\alpha G(E_{\alpha i}+\mbox{i} \varepsilon )|\Phi_{\alpha i}\rangle .
402: \end{eqnarray}
403: Substituting Eq.\ (\ref{g2b}) into (\ref{s12}), 
404: the first term yields two more terms 
405: \begin{equation}
406: S_{\beta j,\alpha i}^{(1)}=\lim\limits_{t\to \infty
407: }\lim\limits_{\varepsilon \to 0}\mbox{i}\varepsilon 
408: \mbox{e}^{\mbox{i}
409: (E_{\beta j}-E_{\alpha i})t}\langle \Phi _{\beta j}|
410: \widetilde{G}_\alpha
411: (E_{\alpha i}+\mbox{i}\varepsilon )|\Phi _{\alpha i}\rangle 
412: \end{equation}
413: \begin{eqnarray}
414: S_{\beta j,\alpha i}^{(2)}&=&\lim\limits_{t\to \infty
415: }\lim\limits_{\varepsilon \to 0}\mbox{i}\varepsilon 
416: \mbox{e}^{\mbox{i}
417: (E_{\beta j}-E_{\alpha i})t}\langle \Phi _{\beta j}|
418: \widetilde{G}_\alpha
419: (E_{\alpha i}+\mbox{i}\varepsilon ) \nonumber \\
420: && U^\alpha G_\alpha ^{(l)}(E_{\alpha i}+\mbox{i}
421: \varepsilon )|\Phi _{\alpha i}\rangle .
422: \end{eqnarray}
423: Using of the properties of the resolvent operators 
424: the limits can be performed and we arrive at the following, physically 
425: plausible, result.
426: The first term, $S_{\beta j,\alpha i}^{(1)}$, is the 
427: S-matrix of a two-body
428: single channel scattering on the potential $u_\alpha ^{(l)}$ 
429: \begin{equation}
430: S_{\beta j,\alpha i}^{(1)}=\delta _{\beta \alpha }
431: \delta _{ji}S(u_\alpha ^{(l)}).
432: \label{s1}
433: \end{equation}
434: If $u_\alpha ^{(l)}$ is a pure Coulomb interaction
435:  $S(u_\alpha ^{(l)})$ falls back
436: to the S-matrix of the Rutherford scattering, if $u_\alpha ^{(l)}$ is
437: identically zero
438: $S_{\beta j,\alpha i}^{(1)}$ equals to unity.
439: The second term, $S_{\beta j,\alpha i}^{(2)}$, describes a two-body 
440: multichannel scattering on the
441: potential $U^\alpha $ 
442: \begin{equation}
443: S_{\beta j,\alpha i}^{(2)}=-2\pi \mbox{i}
444: \delta _{\beta \alpha }\delta
445: (E_{\beta j}-E_{\alpha i})\langle 
446: \widetilde{\Phi }_{\beta j}^{(-)}|U^\alpha
447: |\Phi _{\alpha i}^{(l)(+)}\rangle .  \label{s2}
448: \end{equation}
449: The third term gives account of the complete three-body dynamics 
450: \begin{equation}
451: S_{\beta j,\alpha i}^{(3)}=-2\pi 
452: \mbox{i}\delta (E_{\beta j}-E_{\alpha i})\langle \Phi _{\beta j}^{(l)(-)}|V^\alpha |
453: \Psi _{\alpha i}^{(+)}\rangle .
454: \label{s3}
455: \end{equation}
456: 
457: 
458: \subsection{Lippmann-Schwinger integral equation for 
459: $|\Phi _{\alpha}^{(l)}\rangle$}
460: 
461: 
462: Starting from the definition of $|\Phi _{\alpha}^{(l)}\rangle$, 
463: Eq.\ (\ref{phil}), 
464: by utilizing the resolvent relation (\ref{g2b}) and the definition
465: (\ref{phitild}), we easily derive a Lippmann-Schwinger equation
466: \begin{equation}
467: |\Phi _{\alpha}^{(l)(\pm)}\rangle =| \widetilde{\Phi}_{\alpha}^{(\pm)}\rangle
468: + \widetilde{G}_\alpha (E\pm \mbox{i} \epsilon) U^\alpha
469: |\Phi _{\alpha}^{(l)(\pm)}\rangle,
470: \label{lsphil}
471: \end{equation}
472: where 
473: $| \widetilde{\Phi}_{\alpha}^{(\pm)}\rangle$ are given by
474: \begin{equation}
475: | \widetilde{\Phi}_{\alpha}^{(\pm)}\rangle =
476: | \widetilde{\chi}_{\alpha}^{(\pm)} \rangle
477: | \phi_{\alpha} \rangle.  \label{chipm}
478: \end{equation}
479: The state $|\widetilde{\chi}_{\alpha}^{(\pm)}\rangle$ is
480: a scattering state in the Coulomb-like potential $u_\alpha^{(l)}(y_\alpha)$.
481: 
482: \subsection{Faddeev-Merkuriev integral equations 
483: for the wave function components}
484: 
485: The integral equations for the wave function
486: $|\Psi _{\alpha}^{(\pm)}\rangle$ are arrived at by combining
487: the resolvent  relation (\ref{g3b}) and Eq.\ (\ref{Psidef}). In this case
488: however we have three resolvent relations and therefore we
489: obtain a triad of Lippmann-Schwinger equations
490: \begin{eqnarray}
491: |\Psi _{\alpha}^{(\pm)}\rangle & = |\Phi _{\alpha}^{(l)(\pm)}\rangle +
492:  & G_\alpha^{(l)} (E \pm {\mathrm{i}} 0)
493:  V^\alpha |\Psi _{\alpha}^{(\pm)}\rangle \\
494: |\Psi _{\alpha}^{(\pm)}\rangle  & =  
495: \phantom{|\Phi _{\alpha}^{(l)}\rangle + \ \ \ \ \  }
496: & G_\beta^{(l)}(E \pm {\mathrm{i}} 0)
497:  V^\beta |\Psi _{\alpha}^{(\pm)}\rangle \\
498: |\Psi _{\alpha}^{(\pm)}\rangle & = 
499: \phantom{|\Phi _{\alpha}^{(l)}\rangle + \ \ \ \ \  }
500: & G_\gamma^{(l)}(E \pm {\mathrm{i}} 0)
501:  V^\gamma |\Psi _{\alpha}^{(\pm)}\rangle .
502: \label{lstriad}
503: \end{eqnarray}
504: Although these three equations together provide unique 
505: solutions \cite{gloeckle}, their kernels
506: are not connected therefore they cannot be solved by iterations.
507: The way out of the problem is to use the Faddeev decomposition 
508: which leads to equations with connected kernels, thus they are effectively 
509: Fredholm-type integral equations.
510: 
511: 
512: Multiplying each elements of the triad from left by $G^{(l)} v_\alpha^{(s)}$
513: and utilizing (\ref{fdec}) we get
514: the set of Faddeev-Merkuriev integral equations for the components
515: \begin{eqnarray}
516: |\psi_{\alpha}^{(\pm)} \rangle & = 
517: |\Phi_{\alpha }^{(l)(\pm)}\rangle + 
518: & G_\alpha^{(l)} (E \pm {\mathrm{i}} 0)
519: v^{(s)}_\alpha [ |\psi_{\beta}^{(\pm)} \rangle +
520: |\psi_{\gamma}^{(\pm)} \rangle  ] \label{fm-eq1} \\ 
521: |\psi_{\beta}^{(\pm)} \rangle & =  \phantom{
522: |\Phi_{\beta i}^{(l)(\pm)}\rangle + \ \  }
523: & G_\beta^{(l)} (E \pm {\mathrm{i}} 0)
524: v^{(s)}_\beta [ |\psi_{\gamma}^{(\pm)} \rangle +
525: |\psi_{\alpha}^{(\pm)} \rangle  ] \label{fm-eq2} \\
526: |\psi_{\gamma}^{(\pm)} \rangle & =  \phantom{
527: |\Phi_{\beta i}^{(l)(\pm)}\rangle +  \ \   }
528: & G_\gamma^{(l)} (E \pm {\mathrm{i}} 0)
529: v^{(s)}_\gamma [ |\psi_{\alpha}^{(\pm)} \rangle +
530: |\psi_{\beta}^{(\pm)} \rangle  ].
531: \label{fm-eq3}
532: \end{eqnarray}
533: Merkuriev showed that after a certain number of iterations
534: these equations were reduced to Fredholm integral equations of
535: the second kind
536: with compact kernels for all energies, including energies  below
537: $(E < 0)$ and above $(E  > 0)$ the three-body breakup threshold 
538: \cite{fm-book}. Thus all the nice properties
539: of the original Faddeev equations  established for
540: short-range interactions  remain valid also for the case of Coulomb-like
541: potentials. We note that 
542: the triad of Lippmann-Schwinger equations and the set of 
543: Faddeev equations describe the same physics, the equations have identical
544: spectra and in fact, the Faddeev equations are the adjoint representations
545: of the  triad of Lippmann-Schwinger equations \cite{yakovlev}.
546: 
547: Utilizing the properties of the Faddeev components 
548: the matrix elements in (\ref{s3}) can be rewritten 
549: in a form  better
550: suited for numerical calculations 
551: \begin{equation}
552: \langle \Phi _{\beta j}^{(l)(-)}|V^\alpha |
553: \Psi _{\alpha i}^{(+)}\rangle
554: =\sum_{\gamma \neq \beta }\langle 
555: \Phi _{\beta j}^{(l)(-)}|v^{(s)}_\beta |\psi_{\gamma i}^{(+)}\rangle .  
556: \label{s3v}
557: \end{equation}
558: 
559: Summarizing, in the three-potential formalism, starting from 
560: $| \widetilde{\Phi}_{\alpha}^{(\pm)}\rangle$, by solving a Lippmann-Schwinger
561: equation, we determine $|{\Phi}_{\alpha}^{(l)(\pm)}\rangle$. 
562: Then from $|{\Phi}_{\alpha}^{(l)(+)}\rangle$, by solving the set of
563: Faddeev-Merkuriev equations, we determine the components 
564: $|\psi_{\alpha}^{(+)}\rangle$. Finally using Eqs.\ (\ref{s2}) and (\ref{s3v})
565: we construct the $S$-matrix.
566: 
567: 
568: \section{Coulomb-Sturmian separable expansion approach to the
569:  three-body integral equations }
570: 
571: In order to solve operator equations in quantum mechanics 
572: one needs a suitable representation for the operators. For solving 
573: integral equations it is especially advantageous if one uses a 
574: representation where the Green's operator is simple. 
575: For the two-body Coulomb Green's operator  there exists a Hilbert-space basis
576: in which its representation is very simple. This is the Coulomb-Sturmian (CS)
577: basis. In this representation-space the Coulomb Green's operator can be
578: given by simple and well-computable analytic functions \cite{papp1}. This
579: basis forms a countable set. If we represent the interaction term on a finite
580: subset of the basis it looks like a kind of separable expansion of the
581: potential, and so the integral equation becomes a set of 
582: algebraic equations which can then be solved without any further approximation. 
583: The completeness of the basis ensures the convergence of the method.
584: 
585: This approximation scheme has been thoroughly tested in two-body
586: calculations. Bound- and resonant-state calculations were presented first
587: \cite{papp1}. Then the method was extended to 
588: scattering states \cite{papp2}. Since only 
589: the asymptotically irrelevant short-range interaction is approximated, 
590: the correct Coulomb asymptotic is guaranteed \cite{papp3}.
591: A recent account of this method is presented in Ref.\ \cite{klp}.
592: The method also proved to be very efficient in solving three-body Faddeev-Noble
593: integral equations for bound- \cite{pzwp} and scattering-state \cite{pzsc}
594: problems with repulsive Coulomb interactions.
595: 
596: In subsection A we  define the basis 
597: states in two- and three-particle Hilbert space. In subsection B 
598: we review some of the most important formulae of the two-body 
599: problem. In subsections C and D
600: we describe the calculation of the $S$-matrix and the
601: solution of the Faddeev-Merkuriev integral equations.
602: We follow the line presented in Ref.\ \cite{pzsc}.
603: 
604: 
605: \subsection{Basis states}
606: 
607: The Coulomb-Sturmian  functions \cite{rotenberg} in some 
608: angular momentum state $l$ are defined as 
609: \begin{equation}
610: \langle r|nl\rangle =\left[ \frac{n!}{(n+2l+1)!}\right]
611: ^{1/2}(2br)^{l+1} \exp({-br}) L_n^{2l+1}(2br),  \label{basisr}
612: \end{equation}
613: $n=0,1,2,\ldots $. Here, $L$ represents the
614: Laguerre  polynomials and 
615: $b$ is a fixed parameter.
616: In an angular momentum subspace they form a complete set 
617: \begin{equation}
618: {\bf {1}}=\lim\limits_{N\to \infty }\sum_{n=0}^N|
619: \widetilde{nl}\rangle
620: \langle nl|=\lim\limits_{N\to \infty }{\bf {1}}_N,  \label{unity}
621: \end{equation}
622: where $|\widetilde{nl}\rangle$ in configuration-space 
623: representation reads 
624: $\langle r|\widetilde{nl}\rangle = \langle r|nl\rangle/r$.
625: 
626: The three-body Hilbert space is a direct sum of two-body 
627: Hilbert spaces.
628: Thus, the appropriate basis in angular momentum 
629: representation should be defined as a direct product 
630: \begin{equation}
631: | n \nu l \lambda \rangle_\alpha = 
632: | n l \rangle_\alpha \otimes | \nu
633: \lambda \rangle_\alpha , \ \ \ \ (n,\nu=0,1,2,\ldots),  
634: \label{cs3}
635: \end{equation}
636: with the CS states of Eq.~(\ref{basisr}). Here $l$
637: and $\lambda$ denote the angular momenta associated with Jacobi coordinates
638: $x$ and $y$, respectively. In our three-body Hilbert space basis
639: we take bipolar harmonics in the
640: angular variables and CS functions in the radial coordinates.
641: The completeness relation takes the form (with
642: angular momentum summation implicitly included) 
643: \begin{equation}
644: {\bf 1} =\lim\limits_{N\to\infty} \sum_{n,\nu=0}^N |
645:  \widetilde{n \nu l
646: \lambda} \rangle_\alpha \ \mbox{}_\alpha\langle 
647: {n \nu l \lambda} | =
648: \lim\limits_{N\to\infty} {\bf 1}_{N}^\alpha,
649: \end{equation}
650: where $\langle x_\alpha y_\alpha |
651: \widetilde{ n \nu l \lambda }
652: \rangle_\alpha=  
653: \langle x_\alpha y_\alpha |
654: {\ n \nu l \lambda }\rangle_\alpha /(x_\alpha y_\alpha)$.
655: It should be noted that in the three-particle 
656: Hilbert space we can introduce
657: three equivalent basis sets which belong to fragmentation 
658: $\alpha$, $\beta$
659: and $\gamma$.
660: 
661: 
662: \subsection{Coulomb-Sturmian separable expansion in 
663: two-body scattering problems}
664: 
665: Let us study a two-body case of short-range
666:  plus Coulomb-like interactions 
667: \begin{equation}
668: v_l=v^{(s)}_l+v^C
669: \end{equation}
670: and consider the inhomogeneous Lippmann-Schwinger 
671: equation for the
672: scattering state $|\psi _l\rangle $ in some partial wave $l$ 
673: \begin{equation}
674: |\psi _l\rangle =|\phi _l^C\rangle +g_l^C(E)v^{(s)}_l|\psi _l\rangle .  
675: \label{LS}
676: \end{equation}
677: Here $|\phi _l^C\rangle $ is the regular Coulomb function, 
678: $g_l^C(E)$ is the
679: two-body Coulomb Green's operator 
680: \begin{equation}
681: g_l^C(E)=(E-h_l^0-v^C)^{-1}
682: \end{equation}
683: with the free Hamiltonian $h_l^0$. We make
684:  the following
685: approximation on Eq.~(\ref{LS}) 
686: \begin{equation}
687: |\psi _l\rangle =|\varphi _l^C\rangle +
688: g_l^C(E){\bf {1}}_N v^{(s)}_l {\bf {1}}_N |\psi _l\rangle, \label{LSapp}
689: \end{equation}
690: i.e.\ we approximate the short-range potential 
691: $v_l^{(s)} $ by a separable form 
692: \begin{equation}
693: v_l^{(s)}=\lim_{N\to\infty} {\bf {1}}_N v_l^{(s)} {\bf {1}}_N \approx
694: {\bf {1}}_N v_l^{(s)} {\bf {1}}_N 
695: = \sum_{n, n^{\prime } =0}^N
696: |\widetilde{
697: n l}\rangle  \;
698: \underline{v}_{l}^{(s)}\;\mbox{}
699: \langle \widetilde{n^{\prime }
700: l }|  \label{sepfe2b}
701: \end{equation}
702: where the matrix
703: \begin{equation}
704: \underline{v}_{l_{n n^{\prime }}}^{(s)}=
705: \langle n l|
706: v_l^{(s)}| n^{\prime } l \rangle .  
707: \label{v2b}
708: \end{equation}
709: These matrix elements can always be calculated (numerically)
710: for any reasonable short-range potential. In practice we use
711: Gauss-Laguerre quadrature, which is well-suited to the CS basis.
712: 
713: Multiplied with the CS states $\langle \widetilde{nl}|$
714:  from the left, Eq.~(\ref
715: {LSapp}) turns into a linear system of equations 
716: for the wave-function
717: coefficients $\underline{\psi }_{l_n}=\langle 
718: \widetilde{nl}|\psi _l\rangle $ 
719: \begin{equation}
720: \lbrack (\underline{g}_l^C(E))^{-1}-
721: \underline{v}_l^{(s)}]\underline{\psi }_l=
722:  (\underline{g}_l^C(E))^{-1}  \underline{\varphi }_l^C,  \label{eq18a}
723: \end{equation}
724: where the underlined quantities are matrices with the following elements
725: \begin{equation}
726: \underline{\varphi }_{l_n}^C=\langle 
727: \widetilde{nl}|\varphi _l^C\rangle 
728: \label{phiov}
729: \end{equation}
730: and
731: \begin{equation}
732: \underline{g}_{l_{nn^{\prime }}}^C(E)=
733: \langle \widetilde{nl}|g_l^C(E)|\widetilde{n^{\prime }l}\rangle.  \label{gcme}
734: \end{equation}
735: 
736: 
737: \subsubsection{The matrix elements $\langle \widetilde{nl}|g_l^C(z)|
738: \widetilde{n^{\prime }l}\rangle$}
739: 
740: The key point in the whole procedure is the exact and analytic
741: calculation of the CS matrix elements of the Coulomb Green's operator
742: and of the overlap of the Coulomb and CS functions.
743: For the Green's matrix we have developed two independent, analytic
744: approaches. Both are based on the observation that the Coulomb Hamiltonian
745: possesses an infinite symmetric tridiagonal (Jacobi) matrix structure on
746: CS basis.
747: 
748: Let us consider the radial Coulomb Hamiltonian 
749: \begin{equation}
750:  h^{\rm C}_l=-\frac{\hbar^2}{2m}\left(\frac{\mbox{d}^2 }{\mbox{d} r^2}
751: - \frac{l(l+1)}{r^2}\right) + \frac{Z}{r}\ ,
752: \label{coulham}
753: \end{equation}
754: where $m$, $l$ and $Z$ stands for the mass, angular momentum and
755: charge, respectively.
756: The matrix 
757: $\underline{J}^{\rm C}_{n n^{\prime}}=
758: \langle n |(z- h^{\rm C}_l)|n^{\prime} \rangle$ 
759: possesses a Jacobi structure,
760: \begin{equation}
761: \underline{J}^{\rm C}_{nn}=2(n+l+1) (k^2-b^2 )
762: \frac{\hbar^2}{4mb}- Z 
763: \label{jii}
764: \end{equation}
765: and
766: \begin{equation}
767: \underline{J}^{\rm C}_{nn-1}=-[n(n+2l+1)]^{1/2} (k^2+b^2 )
768: \frac{\hbar^2}{4mb} \ , 
769: \label{jiip1}
770: \end{equation}
771: where $k=(2m z/\hbar^2)^{1/2}$ is the wave number.
772: The main result of Ref.\ \cite{jmp} is that for Jacobi matrix systems 
773: the $N$'th leading submatrix 
774: $\underline{g}^{{\rm C} (N)}_{n n'}$ of the infinite Green's matrix
775: can be determined by the elements of the Jacobi matrix
776: \begin{equation}
777: \label{invn}
778: \underline{g}^{{\rm C} (N)}_{n n'}=[\underline{J}^{\rm C}_{n n'}+ 
779: \delta _{n N}\, \, 
780: \delta _{n' N}\, \, \underline{J}^{\rm C}_{NN+1}\, 
781: \, C ]^{-1}\ ,
782: \end{equation}
783: where $C$ is a  continued fraction
784: \begin{equation}
785: C=-\frac{u_N}{d_N+
786: \frac{\displaystyle u_{N+1}}{\displaystyle d_{N+1}+ 
787: \frac{\displaystyle u_{N+2}}{ \displaystyle d_{N+2} + \cdots }}} \ ,
788: \label{frakk}
789: \end{equation}
790: with coefficients 
791: \begin{equation}
792: u_n=- {\underline{J}^{\rm C}_{n,n-1}}/{\underline{J}^{\rm C}_{n,n+1}}, 
793: \quad d_n=- {\underline{J}^{\rm C}_{n,n}}/{\underline{J}^{\rm C}_{n,n+1}} 
794: \ .
795: \label{egyutth}
796: \end{equation}
797: 
798: In Ref.\ \cite{jmp} it was shown that although the 
799: continued fraction $C$ is convergent only on the upper-half $k$ plane 
800: it can be continued analytically to the whole $k$ plane.
801: This is because the $u_n$ and $d_n$ coefficients satisfy the limit properties
802: \begin{eqnarray}
803: u & \equiv  & \lim_{n\rightarrow \infty }u_n=-1 \\
804: d & \equiv  & \lim_{n\rightarrow \infty} d_n= 2(k^2 -   b^2)/ 
805:  ( k^2 +   b^2)\ . 
806: \label{ud}
807: \end{eqnarray}  
808: Then the continued fraction  appears as
809: \begin{equation}
810: C=-\frac{u_N}{d_N+
811: \frac{\displaystyle u_{N+1}}{\displaystyle d_{N+1}+ \cdots +
812: \frac{\displaystyle u}{ \displaystyle d +  
813: \frac{\displaystyle u}{ \displaystyle d +    \cdots 
814:  }}}}\ .
815: \label{cflim}
816: \end{equation}
817: Therefore the tail  $w$ of $C$ satisfies the implicit
818: relation
819: \begin{equation}
820: w=\frac{u}{d+w}\ , 
821: \label{tail2}
822: \end{equation}
823: which is solved by
824: \begin{equation}
825: \label{wpm}
826: w_{\pm}=(b \pm  {\mathrm{i}}k )^2/(b^2+k^2)  \ .
827: \end{equation}
828: Replacing the tail of the continued fraction by its explicit analytical 
829: form $w_{\pm}$, we can speed up the convergence and, more importantly 
830: turn a non-convergent continued fraction into a
831: convergent one \cite{lorentzen}. Analytic continuation is achieved
832: by using $w_{\pm}$  instead of the non-converging tail.
833: In Ref.\ \cite{jmp} it was 
834: shown that $w_+$ provides an analytic continuation of the Green's matrix 
835: to the physical, while $w_{-}$ to the unphysical Riemann-sheet. This way 
836: Eq.\ (\ref{frakk}) together with (\ref{invn}) provides the CS basis
837: representation of the Coulomb Green's operator on the whole complex
838: $k$ plane.  We note here that with the choice 
839: of $Z=0$ the Coulomb Hamiltonian (\ref{coulham}) reduces to the kinetic 
840: energy operator and our formulas provide the CS basis representation of the 
841: Green's operator of the free particle as well.
842: We emphasize that this procedure 
843: does not truncate the Coulomb Hamiltonian, because
844: all the higher $\underline{J}_{nn'}$  matrix elements are implicitly contained 
845: in the continued fraction. 
846: 
847: We note that $\underline{g}^C$ has already
848: been calculated before \cite{papp1}. From the J-matrix structure a three-term
849: recursion relation follows for the matrix elements $\underline{g}^C_{n n'}$.
850: This recursion relation is solvable if the first element 
851: $\underline{g}^C_{00}$ is known. 
852: It is given in a closed analytic form
853: \begin{eqnarray}
854: \underline{g}^C_{00}&=& \frac{4mb}{\hbar^2} 
855: \frac{1}{(b-\mbox{i}k)^2} 
856: \frac{1}{l+\mbox{i}\eta +1} \nonumber \\
857: &&\times {_2}F_{1}\left(-l+\mbox{i}\eta,1;l+\mbox{i}\eta+2,
858: \left(\frac{b+\mbox{i}k} {b-\mbox{i}k}\right)^2 \right),
859: \label{g00}
860: \end{eqnarray}
861: where  $\eta=Z m/(\hbar^2 k)$ is the Coulomb
862: parameter and ${_2}F_{1}$ is the hypergeometric function. 
863: For those cases where the first or the second index of
864: ${_2}F_{1}$ is equal to unity, there exists a continued fraction representation,
865: which is very efficient in practical calculations. It  was shown that 
866: the two methods lead to numerically identical results for all energies and
867: our numerical continued fraction representation possesses all the 
868: analytic properties  of $g^C$. The exact analytic knowledge of  $g^C$ 
869: allows us to calculate the matrix elements of the full Green's 
870: operator in the whole complex plane 
871: \begin{equation}
872: \underline{g}_l(z)=((\underline{g}_l^C(z))^{-1}-
873: \underline{v}_l^{(s)})^{-1}.
874: \label{2bgreen}
875: \end{equation}
876: 
877: The overlap vector of CS and the Coulomb functions  $\langle 
878: \widetilde{nl}|\varphi _l^C\rangle$ is known analytically \cite{papp2}.
879: It can be calculated by a three-term recursion, derived from the
880: J-matrix, using the starting value
881: \begin{eqnarray}
882: \langle 
883: \widetilde{0l}|\varphi _l^C\rangle &=&
884: \exp(2\eta\arctan(k/b))\sqrt{\frac{2\pi\eta}{\exp(2\pi\eta)-1 }} \nonumber \\
885: &&\times \left(\frac{2k/b}{1+k^2/b^2}\right)^{l+1} 
886: \prod_{i=1}^l \left( \frac{\eta^2+i^2}{i(i+1/2)} \right)^{1/2} .
887: \end{eqnarray}
888: 
889: 
890: \subsection{Calculation of the three-body S-matrix}
891: 
892: The aim of any scattering calculation is to determine the 
893: S-matrix elements. In our case we need to calculate the terms
894: (\ref{s1}), (\ref{s2}) and (\ref{s3v}) of the three-potential
895: picture.
896: 
897: The term $S^{(1)}_{\beta j,\alpha i}$ is trivial because it is just the two-body
898: S-matrix of the Coulomb-like potential $u_\alpha^{(l)}$.
899: 
900: To calculate the second term, 
901: $S^{(2)}_{\beta j,\alpha i}$ of Eq.\ (\ref{s2}), 
902: the matrix elements $\langle \widetilde{\Phi }_{\alpha j}^{(-)}|U^\alpha
903: |\Phi _{\alpha i}^{(l)(+)}\rangle$ are needed. Since 
904: $\langle \widetilde{\Phi }_{\alpha j}^{(-)}|$ contains a two-body
905: bound-state wave function in coordinate $x_\alpha$ this matrix element
906: is confined to $\Omega_\alpha$, where $U^\alpha$ is of short-range type.
907: Therefore a separable approximation is justified
908: \begin{equation}
909: \langle \widetilde{\Phi }_{\alpha j}^{(-)}|U^\alpha
910: |\Phi _{\alpha i}^{(l)(+)}\rangle
911: \approx    \langle \widetilde{\Phi }_{\alpha j}^{(-)}| 
912: {\bf 1}_N^\alpha U^\alpha {\bf 1}_N^\alpha
913: |\Phi _{\alpha i}^{(l)(+)}\rangle,
914: \end{equation}
915: i.e, in this matrix element, we can approximate $U^\alpha$ by
916: a separable form
917: \begin{eqnarray}
918: U^\alpha   & = &
919: \lim_{N\to\infty} {\bf 1}_N^\alpha U^\alpha  {\bf 1}_N^\alpha 
920: \approx {\bf 1}_N^\alpha U^\alpha {\bf 1}_N^\alpha \nonumber \\
921:   & \approx & \sum_{n,\nu ,n^{\prime },
922: \nu ^{\prime }=0}^N|\widetilde{n\nu l\lambda }\rangle _\alpha \;
923: \underline{U}^{\alpha}\;\mbox{}_\alpha \langle \widetilde{n^{\prime }
924: \nu ^{\prime }l^{\prime }\lambda^{\prime }}|  \label{sepfeU}
925: \end{eqnarray}
926: where 
927: \begin{equation}
928: \underline{U}^\alpha_{ n\nu l \lambda,
929:  n^{\prime } \nu^{\prime } l^{\prime} \lambda^{\prime }}=
930: \mbox{}_\alpha \langle n\nu l\lambda |
931: U^\alpha |n^{\prime }\nu^{\prime}l^{\prime}{\lambda}^{\prime }
932: \rangle_\alpha .  
933: \label{Ua}
934: \end{equation}
935: The matrix element appears as
936: \begin{equation}
937: \langle \widetilde{\Phi }_{\alpha j}^{(-)}|U^\alpha
938: |\Phi _{\alpha i}^{(l)(+)}\rangle
939: \approx \sum^N   \langle \widetilde{\Phi }_{\alpha j}^{(-)}| 
940: \widetilde{n\nu l\lambda} \rangle_\alpha  \underline{U}^\alpha 
941: \mbox{}_\alpha\langle \widetilde{n' \nu' l'\lambda'}
942: |\Phi _{\alpha i}^{(l)(+)}\rangle. \label{sumu}
943: \end{equation}
944: 
945: In calculating the third term, $S^{(3)}_{\beta j,\alpha i}$ of (\ref{s3v}), 
946: we have matrix elements of the type 
947: $\langle \widetilde{\Phi }_{\alpha j}^{l(-)}| v_\alpha^{(s)}  
948: |\psi_{\beta i}^{(+)}\rangle$. Here 
949: we can again approximate the short-range potential 
950: $v_\alpha^{(s)} $ in the three-body
951: Hilbert space by a separable form 
952: \begin{eqnarray}
953: v_\alpha^{(s)}   & = &
954: \lim_{N\to\infty} {\bf 1}_N^\alpha v_\alpha^{(s)}  {\bf 1}_N^\beta 
955: \approx {\bf 1}_N^\alpha v_\alpha^{(s)}  {\bf 1}_N^\beta \nonumber \\
956:   & \approx & \sum_{n,\nu ,n^{\prime },
957: \nu ^{\prime }=0}^N|\widetilde{
958: n\nu l\lambda }\rangle _\alpha \;
959: \underline{v}_{\alpha \beta }^{(s)}\;\mbox{}_\beta \langle 
960: \widetilde{n^{\prime } \nu ^{\prime }l^{\prime }\lambda
961: ^{\prime }}|  \label{sepfe}
962: \end{eqnarray}
963: where 
964: \begin{equation}
965: \underline{v}_{{\alpha \beta}_{n \nu l \lambda, 
966: n^{\prime } \nu ^{\prime } l^{\prime } \lambda^{\prime } }}^{(s)}=
967: \mbox{}_\alpha \langle n\nu l\lambda |
968: v_\alpha^{(s)}|n^{\prime }\nu ^{\prime}
969: l^{\prime }{\lambda }^{\prime }\rangle_\beta .  
970: \label{vab}
971: \end{equation}
972: In (\ref{sepfe}) the ket and bra states belong to different 
973: fragmentations depending on the
974: neighbors of the potential operators in the matrix elements.
975: Finally, the matrix elements take the form
976: \begin{equation}
977: \langle {\Phi }_{\alpha j}^{l(-)}| v_\alpha^{(s)}  
978: |\psi_{\beta i}^{(+)}\rangle \approx
979: \sum^N 
980: \langle {\Phi }_{\alpha j}^{l(-)}| 
981: \widetilde{ n\nu  l \lambda}\rangle_\alpha 
982: \;\underline{v}_{\alpha \beta }^{(s)}\;
983: \mbox{}_\beta\langle  \widetilde{ n'\nu'  l' \lambda'} 
984:  |\psi_{\beta i}^{(+)}\rangle. \label{sumv}
985: \end{equation}
986: 
987: We conclude that to calculate the S-matrix of the
988: three-potential formulae we need the CS matrix elements (\ref{Ua})
989: and (\ref{vab}), which can always be evaluated numerically
990: by using the transformation of Jacobi coordinates \cite{bb}.
991: In addition we need the CS wave function components 
992: $\mbox{}_\alpha\langle \widetilde{n\nu l\lambda} 
993: | \widetilde{\Phi }_{\alpha i}^{(\pm)}\rangle$, 
994: $\mbox{}_\alpha\langle \widetilde{n\nu l\lambda} 
995: | {\Phi}_{\alpha i}^{l(\pm)}\rangle$ and
996: $\mbox{}_\alpha \langle  \widetilde{ n\nu l \lambda} 
997:  |\psi_{\alpha }^{(+)}\rangle$. We  determine them in the
998: following section by solving Lippmann-Schwinger and
999: Faddeev-Merkuriev integral equations.
1000: 
1001: It should be noted that  the approximations
1002: (\ref{sepfeU}) and (\ref{sepfe}) used
1003: in calculating the matrix elements (\ref{sumu}) and (\ref{sumv})
1004: become equalities as $N$ goes to infinity.
1005: In practical calculations we increase $N$ until we observe 
1006: numerical convergence in scattering observables.
1007: 
1008: \subsection{Solution of the three-body integral equations}
1009: 
1010: In the set of Faddeev-Merkuriev equations (\ref{fm-eq1}-\ref{fm-eq3}) 
1011: we make the approximation of (\ref{sepfe})
1012: \begin{eqnarray}
1013: |\psi _\alpha \rangle & =|\Phi _{\alpha i}^{(l)}\rangle
1014: + & G_\alpha^{(l)}[{\bf 1}_N^\alpha v^{(s)}_\alpha 
1015: {\bf 1}_N^\beta |\psi _\beta \rangle +
1016: {\bf 1}_N^\alpha v^{(s)}_\alpha {\bf 1}_N^\gamma 
1017: |\psi _\gamma \rangle ]  \label{feqsapp1} \\
1018: |\psi _\beta \rangle & =\phantom{|\Phi _{\alpha i}^{(l)}\rangle
1019: + \  } & G_\beta^{(l)}[{\bf 1}_N^\beta v^{(s)}_\beta 
1020: {\bf 1}_N^\gamma |\psi _\gamma \rangle +
1021: {\bf 1}_N^\beta v^{(s)}_\beta {\bf 1}_N^\alpha 
1022: |\psi _\alpha \rangle ] \label{feqsapp2} \\
1023: |\psi _\gamma \rangle & =\phantom{|\Phi _{\alpha i}^{(l)}\rangle
1024: + \  } & G_\gamma^{(l)}[{\bf 1}_N^\gamma v^{(s)}_\gamma 
1025: {\bf 1}_N^\alpha |\psi_\alpha \rangle +
1026: {\bf 1}_N^\gamma v^{(s)}_\gamma {\bf 1}_N^\beta 
1027: |\psi_\beta \rangle ].
1028: \label{feqsapp3}
1029: \end{eqnarray}
1030: Multiplied  by the CS states 
1031: $\mbox{}_\alpha \langle \widetilde{n\nu l\lambda }|$, 
1032: $\mbox{}_\beta \langle \widetilde{n\nu l\lambda }|$ and
1033: $\mbox{}_\gamma \langle \widetilde{n\nu l\lambda }|$, respectively, 
1034: from the left the set of integral equations
1035: turn into a linear system of algebraic equations
1036: for the coefficients of the Faddeev 
1037: components $\underline{\psi }_{\alpha_{ n\nu l
1038: \lambda} }=\mbox{}_\alpha \langle 
1039: \widetilde{n\nu l\lambda }
1040: |\psi _\alpha \rangle $: 
1041: \begin{equation}
1042: \lbrack (\underline{G}^{(l)})^{-1}-
1043: \underline{v}^{(s)}]\underline{\psi }= (\underline{G}^{(l)})^{-1}  
1044: \underline{\Phi }^{(l)},  \label{fep1}
1045: \end{equation}
1046: with 
1047: \begin{equation}
1048: \underline{G}_{\alpha_{ n \nu l \lambda, n^{\prime} \nu^{\prime }
1049: l^{\prime } \lambda^{\prime }}}^{(l)}=
1050: \ \mbox{}_\alpha \langle \widetilde{n\nu l\lambda }|
1051: G_\alpha ^{(l)}|\widetilde{
1052: n^{\prime }\nu ^{\prime }{l^{\prime }}
1053: {\lambda ^{\prime }}}\rangle _\alpha ,
1054: \label{G}
1055: \end{equation}
1056: and 
1057: \begin{equation}
1058: \underline{\Phi }_{\alpha_{ n \nu l \lambda}}^{(l)}=
1059: \mbox{}_\alpha \langle 
1060: \widetilde{n\nu l\lambda }|\Phi _\alpha ^{(l)}\rangle . 
1061:  \label{P}
1062: \end{equation}
1063: Notice that the matrix elements of the Green's 
1064: operator are needed only
1065: between the same partition $\alpha $ whereas 
1066: the matrix elements of the
1067: potentials occur only between different 
1068: partitions $\alpha $ and $\beta $.
1069: 
1070: 
1071: \subsubsection{The matrix elements 
1072: $ \mbox{}_\alpha \langle \widetilde{n\nu l\lambda }|
1073: G_\alpha ^{(l)}|\widetilde{n^{\prime }\nu ^{\prime }{l^{\prime }}
1074: {\lambda ^{\prime }}}\rangle _\alpha$ and $\mbox{}_\alpha \langle 
1075: \widetilde{n\nu l\lambda }|\Phi _\alpha ^{(l)}\rangle$}
1076: 
1077: Unfortunately neither the matrix elements (\ref{G}) 
1078: nor the overlaps (\ref{P}) are known.
1079: The appropriate Lippmann-Schwinger equation for $G_\alpha^{(l)}$ 
1080: was proposed by Merkuriev \cite{fm-book}
1081: \begin{equation}
1082: G_\alpha^{(l)}(z)=G_\alpha^{as}(z) + G_\alpha^{as}(z) V^{as}_\alpha
1083: G_\alpha^{(l)}(z),
1084: \label{LSass}
1085: \end{equation}
1086: where $G_\alpha^{as}$ and $V^{as}_\alpha$ 
1087: are the asymptotic channel Green's operator and potential, respectively. 
1088: A similar equation is valid for $|\Phi _\alpha ^{(l)}\rangle$
1089: \begin{equation}
1090: |\Phi _\alpha ^{(l)}\rangle=|\Phi_\alpha^{as}\rangle
1091:  + G_\alpha^{as}(z) V^{as}_\alpha |\Phi_\alpha^{(l)}\rangle.
1092: \label{LSassphi}
1093: \end{equation}
1094: Both $G_\alpha^{(l)}$ and $|\Phi _{\alpha}^{(l)}\rangle$
1095: are genuine three-body quantities. One may wonder why a
1096: single Lippmann-Schwinger equation suffices. The Hamiltonian $H_\alpha^{(l)}$
1097: has a peculiar property - it has only $\alpha$-type
1098: two-body asymptotic channels. For such systems a single Lippmann-Schwinger
1099: equation provides a unique solution \cite{sandhas}.
1100: 
1101: The objects $G_\alpha^{as}$, $V^{as}_\alpha$ and $\Phi^{as}_\alpha$
1102: are very complicated. Their leading order terms were
1103: constructed in configurations space in the different asymptotic regions. 
1104: The potential $V^{as}$,  as $|X|\to \infty $,
1105: decays faster than the Coulomb potential
1106: in all directions of the three-body configuration space:
1107: $ V^{as} \sim {\cal O} (|X|^{-1-\epsilon}),\ \epsilon > 0$ \cite{fm-book}.
1108: Therefore we may express the solutions of Eqs.\ (\ref{LSass})
1109: and (\ref{LSassphi}) formally as 
1110: \begin{equation}
1111: (\underline{G}^{(l)}_\alpha )^{-1}= 
1112: (\underline{{G}}^{as}_\alpha )^{-1} -
1113: \underline{V}^{as}_\alpha
1114: \end{equation}
1115: and 
1116: \begin{equation}
1117: [ (\underline{G}^{as}_\alpha )^{-1} -
1118: \underline{V}^{as}_\alpha ]    \underline{\Phi}_\alpha^{(l)} =  
1119: (\underline{G}^{as}_\alpha )^{-1} 
1120: \underline{\Phi}^{as}_\alpha ,
1121: \label{eqphil1}
1122: \end{equation}
1123: respectively, where 
1124: \begin{equation}
1125: \underline{{G}}^{as}_{\alpha_{ n \nu l \lambda, n^{\prime}\nu^{\prime}
1126: l^{\prime} {\lambda}^{\prime} }} =
1127:  \mbox{}_\alpha\langle n
1128: \nu l \lambda | {G}^{as}_\alpha |
1129:  n^{\prime}\nu^{\prime}l^{\prime}{
1130: \lambda}^{\prime}\rangle_\alpha ,  \label{gasme}
1131: \end{equation}
1132: \begin{equation}
1133: \underline{V}^{as}_{\alpha_{ n \nu l \lambda, n^{\prime}\nu^{\prime}
1134:  l^{\prime} {\lambda}^{\prime}}} =
1135:  \mbox{}_\alpha\langle n
1136: \nu l \lambda | V^{as}_\alpha | n^{\prime}\nu^{\prime}
1137: l^{\prime}{\lambda}
1138: ^{\prime}\rangle_\alpha \label{vasme}
1139: \end{equation}
1140: and 
1141: \begin{equation}
1142: \underline{\Phi}^{as}_{\alpha_{ n \nu l \lambda}}  = 
1143: \mbox{}_\alpha\langle \widetilde{ n \nu l \lambda }| 
1144: {\Phi}^{as}_\alpha \rangle.
1145: \end{equation}
1146: Here, ${G}^{as}_\alpha$, ${V}^{as}_\alpha$ and $\Phi^{as}_\alpha$ 
1147: appear between finite number of 
1148: of square-integrable CS states, which confine the domain of integration 
1149: to $\Omega_\alpha$. 
1150: In this region, however, $G_\alpha^{as}$ coincides 
1151: with $\widetilde{G}_\alpha$, $V^{as}_\alpha$ with $U^\alpha$ and
1152: $\Phi^{as}_\alpha$ with $\widetilde{\Phi}_\alpha$ \cite{fm-book}.
1153: Finally we have
1154: \begin{equation}
1155: (\underline{G}^{(l)}_\alpha )^{-1}= 
1156: (\underline{\widetilde{G}}_\alpha )^{-1} -
1157: \underline{U}^\alpha ,
1158: \label{gleq}
1159: \end{equation}
1160: where 
1161: \begin{equation}
1162: \underline{\widetilde{G}}_{\alpha_{ n \nu l \lambda, 
1163:  n^{\prime}\nu^{\prime}l^{\prime} {\lambda}^{\prime}}} =
1164:  \mbox{}_\alpha\langle n
1165: \nu l \lambda | \widetilde{G}_\alpha |
1166:  n^{\prime}\nu^{\prime}l^{\prime}{
1167: \lambda}^{\prime}\rangle_\alpha  \label{gtilde}
1168: \end{equation}
1169: and 
1170: \begin{equation}
1171: \underline{U}^\alpha_{ n \nu l \lambda,
1172:  n^{\prime}\nu^{\prime} l^{\prime} {\lambda}^{\prime}} =
1173:  \mbox{}_\alpha\langle n
1174: \nu l \lambda | U^\alpha | n^{\prime}\nu^{\prime}
1175: l^{\prime}{\lambda}
1176: ^{\prime}\rangle_\alpha.
1177: \end{equation}
1178: And in a similar way   
1179: \begin{equation}
1180: [ (\underline{\widetilde{G}}_\alpha )^{-1} -
1181: \underline{U}^\alpha ]    \underline{\Phi}_\alpha^{(l)} =  
1182: (\underline{\widetilde{G}}_\alpha )^{-1} 
1183: \underline{\widetilde{\Phi}}_\alpha ,
1184: \label{eqphil}
1185: \end{equation}
1186: where 
1187: \begin{equation}
1188: \underline{\widetilde{\Phi}}_{\alpha_{ n \nu l \lambda}}  = 
1189: \mbox{}_\alpha\langle \widetilde{ n \nu l \lambda }| 
1190: \widetilde{\Phi}_\alpha \rangle.
1191: \end{equation}
1192: 
1193: We note that from Eq.\ (\ref{gleq}) 
1194: follows that the left side of Eq.\ (\ref{eqphil}) is just the 
1195: inhomogeneous term of Eq.\ (\ref{fep1}). Both  Eqs.\ (\ref{eqphil})
1196: and (\ref{fep1}) are solved with the same inhomogeneous term.
1197: 
1198: 
1199: \subsubsection{The matrix elements 
1200: $ \mbox{}_\alpha\langle n
1201: \nu l \lambda | \widetilde{G}_\alpha |
1202:  n^{\prime}\nu^{\prime}l^{\prime}{
1203: \lambda}^{\prime}\rangle_\alpha $ and  
1204: $\mbox{}_\alpha\langle \widetilde{ n \nu l \lambda }| 
1205: \widetilde{\Phi}_\alpha \rangle$}
1206: 
1207: The three-particle free Hamiltonian
1208: can be written  as a sum of two-particle
1209: free Hamiltonians 
1210: \begin{equation}
1211: H^0=h_{x_\alpha }^0+h_{y_\alpha }^0.
1212: \end{equation}
1213: Then the Hamiltonian $\widetilde{H}_\alpha$ of Eq.\ (\ref{htilde}) appears as 
1214: a sum of two Hamiltonians 
1215: acting on different coordinates 
1216: \begin{equation}
1217: \widetilde{H}_\alpha =h_{x _\alpha }+h_{y_\alpha },
1218: \end{equation}
1219: with $h_{x_\alpha }=
1220: h_{x_\alpha }^0+v_\alpha^C(x _\alpha )$ and $h_{y_\alpha }=
1221: h_{y_\alpha }^0+u_\alpha^{(l)}(y_\alpha )$, which, of course, commute. 
1222: The state $| \widetilde{\Phi}_\alpha \rangle$, which is an eigenstate
1223: of $\widetilde{H}_\alpha$, is a product of a
1224: two-body bound-state wave function in coordinate 
1225: $x_\alpha$ and a two-body scattering-state wave function in coordinate 
1226: $y_\alpha$. Their CS representations are known from the 
1227: two-particle case described before. 
1228: 
1229: The matrix elements of $\widetilde{G}_\alpha$ can be determined by 
1230: making use of the convolution theorem 
1231: \begin{eqnarray}
1232: \widetilde{G}_\alpha (z) & = & (z-h_{x_\alpha }-
1233: h_{y_\alpha })^{-1} \nonumber \\
1234: &=& \frac 1{2\pi \mbox{i}}\oint_C dz' (z-z'-h_{x_\alpha })^{-1}  
1235:  (z'-h_{y_\alpha })^{-1}. 
1236:  \label{contourint}
1237: \end{eqnarray}
1238: The contour $C$ should encircle, in positive direction, the 
1239: spectrum of $h_{y_\alpha }$
1240: without penetrating into the spectrum of $h_{x_\alpha }$. 
1241: 
1242: The convolution theorem follows from a more general formula.
1243: A function of a self adjoint operator $h$ is defined as
1244: \begin{eqnarray}
1245: f(h)=\frac 1{2\pi \mbox{i}} \oint_C dz f(z) (z-h)^{-1},
1246: \end{eqnarray}
1247: where $C$ is a contour around the spectrum of $h$ and $f$ should be
1248: analytic on the region encircled by $C$.
1249: 
1250: In the following we suppose that $u^{(l)}$ either vanishes or is
1251: a repulsive Coulomb-like potential. This assumption is not necessary but
1252: it greatly simplifies the analysis below. Numerical examples show that
1253: there are a great many physical three-body systems where this condition 
1254: is satisfied. This condition ensures that $h_y$ does not have bound
1255: states.
1256: 
1257: To examine the analytic structure of the integrand (\ref{contourint}) let us
1258: shift the spectrum of $g_{x_\alpha }$ by
1259: taking  $z=E +{\mathrm{i}}\varepsilon$  with
1260: positive $\varepsilon$. In doing so,
1261: the two spectra become well separated and
1262: the spectrum of $g_{y_\alpha}$ can be encircled.
1263: The contour $C$ is deformed analytically
1264: in such a way that the upper part descends to the unphysical
1265: Riemann sheet of $g_{y_\alpha}$, while
1266: the lower part of $C$ can be detoured away from the cut
1267:  [see  Fig.~\ref{fig1}]. The contour still
1268: encircles the branch cut singularity of $g_{y_\alpha}$,
1269: but in the  $\varepsilon\to 0$ limit avoids the singularities of $g_{x_\alpha}$.
1270:  Thus, the mathematical conditions for
1271: the contour integral representation of $\widetilde{G}_\alpha (z)$ in
1272: Eq.~(\ref{contourint}) is met. 
1273: The matrix elements $\underline{\widetilde{G}}_\alpha$
1274: can be cast in the form
1275: \begin{equation}
1276: \widetilde{\underline{G}}_\alpha (z)=
1277:  \frac 1{2\pi \mathrm{i}}\oint_C
1278: dz^\prime \,\underline{g}_{x_\alpha }(z-z^\prime)\;
1279: \underline{g}_{y_\alpha}(z^\prime),
1280: \label{contourint2}
1281: \end{equation}
1282: where the corresponding CS matrix elements of the two-body Green's operators in
1283: the integrand are known analytically for all complex energies.
1284: 
1285: 
1286: 
1287: \section{Test of the method}
1288: 
1289: We demonstrate the power of this new method by calculating elastic phase shifts of
1290: $e^++H$ scattering below the $Ps(n=1)$ threshold and cross sections
1291: of the $e^++H$ elastic scattering
1292: as well as $p^++Ps$ reaction channels up to the $Ps(n=2)$ threshold. 
1293: In all examples we have total angular momentum $L=0$ and we have taken 
1294: angular momentum channels up to $l=10$. We use atomic units.
1295: 
1296: Let us numerate the particles $e^+$, $p$ and $e^-$,
1297: with masses $m_{e^\pm}=1m_e$ and $m_{p}=1836.1527m_e$, by $1$, $2$ and $3$,
1298: respectively. In the channel $3$ there are no two-body
1299: asymptotic channels since the particles $e^+$ and $p$ do not form bound states.
1300: Therefore, we can take $v_3^{(s)}\equiv 0$ and include the total
1301: $v_3^C$ in the long range Hamiltonian
1302: \begin{eqnarray}
1303: H &=& H^{(l)}+v_1^{(s)}+v_2^{(s)}, \\
1304: H^{(l)} &=& H^0+v_1^{(l)}+v_2^{(l)}+v_3^C .
1305: \end{eqnarray}
1306: In this case $|\psi_3 \rangle \equiv 0$ and we have the set of
1307: two-component Faddeev-Merkuriev equations
1308: \begin{eqnarray}
1309: |\psi_1 \rangle &=& |\phi^{(l)}_1\rangle +G^{(l)}_1 v^{(s)}_1 |\psi_1\rangle \\
1310: |\psi_2 \rangle &=& \phantom{|\phi^{(l)}_1\rangle + \  } 
1311: G^{(l)}_2 v^{(s)}_2 |\psi_2\rangle .
1312: \end{eqnarray}
1313: The parameters of the splitting function $\zeta$ of Eq.\ (\ref{oma1})
1314: are rather arbitrary. The final converged results should be insensitive 
1315: to their values; our numerical experiences confirm this expectation. 
1316: For the parameters of $\zeta$ we 
1317: have taken $\nu=2.1$, $x^0=3$ and $y^0=10$, whereas
1318: for the parameters
1319: of CS functions we have taken $b=0.9$. We have experienced that the 
1320: rate of convergence is rather
1321: insensitive on the choice of $b$ over a broad interval.
1322: 
1323: First we examine the convergence of
1324: the results for cross sections at incident wave numbers
1325: $k_1=0.71$, $k_1=0.75$ and  $k_1=0.8$,
1326: which correspond to  scattering states  in the Ore gap.
1327: Table \ref{tabconv1} shows the convergence of
1328: $e^++H->e^++H$ elastic scattering
1329: ($\sigma_{11}$) and $e^++H->p^++Ps$ positronium formation ($\sigma_{12}$)
1330: cross sections (in $\pi a_0^2$) with respect to $N$, the number of CS functions in the
1331: expansion, and with respect to increasing the angular momentum
1332: channels in the bipolar expansion. For comparison we provide the results
1333: of Ref.\ \cite{kwh}. We can see that very good accuracy is 
1334: achieved even with relatively low $N$ in the expansion.
1335: 
1336: In Table \ref{tabshift} we compare 
1337: our converged results for phase shifts (in radians) 
1338: below the $Ps(n=1)$ threshold to that of other methods.  
1339: Ref.\ \cite{5} is the best variational calculation. In 
1340: Ref.\ \cite{20} the Schr\"odinger equation was solved by means of
1341: finite-element method. In Refs.\ \cite{14} and \cite{kwh} the 
1342: configurations space Faddeev-Merkuriev differential
1343: equations were solved using the bipolar harmonic expansion method 
1344: and in total angular momentum representation, respectively.
1345: We can report perfect agreements with previous calculations.
1346: 
1347: In Table \ref{4channel} we present partial cross sections 
1348: in the $H(n=2)-Ps(n=2)$ gap (threshold energies 0.7496-0.8745 Ry). 
1349: In Ref.\ \cite{hu99} the configurations space Faddeev-Merkuriev differential
1350: equations were solved using the bipolar harmonic expansion in the angular
1351: variables an quintic spline expansion in the radial coordinates. 
1352: We can report fairly good agreements.
1353: 
1354: \section{Conclusion}
1355: 
1356: We have extended the three-potential formalism for 
1357: treating the three-body scattering problem with all kinds of
1358: Coulomb interactions including attractive ones.
1359: We adopted Merkuriev's approach and split the Coulomb potentials
1360: in the three-body configuration space into short-range and long-range
1361: terms. In this picture the three-body Coulomb
1362: scattering process can be decomposed into a 
1363: single channel Coulomb scattering, a two-body
1364: multichannel scattering on the intermediate-range 
1365: polarization potential and a genuinely three-body scattering due to the
1366: short-range potentials. The formalism provides us a set of 
1367: Lippmann--Schwinger  and Faddeev-Merkuriev integral equations.
1368: 
1369: These integral equations are certainly too complicated for the most of the
1370: numerical methods available in the literature. The Coulomb-Sturmian
1371: separable expansion method can be successfully applied.
1372:  It solves the three-body
1373: integral equations by expanding only the short-range terms
1374: in a separable form on Coulomb-Sturmian basis 
1375: while treating the long-range terms in an exact manner via a proper integral 
1376: representation of the three-body channel distorted Coulomb Green's operator.
1377: The use of the Coulomb-Sturmian basis 
1378: is essential as it allows an exact analytic representation of the two-body 
1379: Green's operator, and thus the contour integral for the channel distorted 
1380: Coulomb Green's operator can be calculated.
1381: The method provides solutions which are
1382: asymptotically correct, at least in $\Omega_\alpha$, which
1383: is sufficient if the scattering process starts from a two-body
1384: asymptotic state. Since the two-body Coulomb Green's operator 
1385: is exactly calculated all  thresholds are automatically in the right location
1386: irrespective of the rank of the separable approximation.
1387: The method possesses good convergence properties and in 
1388: practice it can be made arbitrarily accurate by employing an increasing 
1389: number of terms in the expansion. Certainly, there is plenty of room
1390: for improvement but we are convinced that this method can be a very 
1391: powerful tool for studying three-body systems with Coulomb interactions.
1392: 
1393: 
1394: 
1395: \acknowledgements
1396: This work has been supported by the NSF Grant No.Phy-0088936
1397: and by the OTKA Grant No.\ T026233. We also acknowledge the
1398: generous allocation of computer time at the NPACI, formerly 
1399: San Diego Supercomputing Center, by the National Resource Allocation
1400: Committee and at the Department of Aerospace Engineering
1401: of CSULB.
1402: 
1403: 
1404: \newpage
1405: \begin{table}[tbp]
1406: \caption{Convergence of
1407: $e^++H->e^++H$ elastic scattering
1408: ($\sigma_{11}$) and $e^++H->p+Ps$ positronium formation ($\sigma_{12}$)
1409: cross sections (in $\pi a_0^2$) with respect to $N$, the number of CS functions in the
1410: expansion, and with respect to increasing the angular momentum
1411: channels ($l_{max}$) in the bipolar basis.}
1412: \label{tabconv1}
1413: \begin{tabular}{|l|cc|cc|cc|}
1414:    & \multicolumn{2}{c|}{$l_{max}=6$} & \multicolumn{2}{c|}{$l_{max}=8$} 
1415:    & \multicolumn{2}{c|}{$l_{max}=10$}  \\
1416: $N$& $\sigma_{11}$& $\sigma_{12}$& $\sigma_{11}$& $\sigma_{12}$& 
1417: $\sigma_{11}$& $\sigma_{12}$  \\ \hline
1418: \multicolumn{7}{|c|}{$k_1=0.71$, 
1419: Ref.\ \cite{kwh}: $\sigma_{11}=0.025$, $\sigma_{12}=0.0038$} \\ \hline
1420: 12 & 0.02662 & 0.00423 & 0.02664 & 0.00397 & 0.02665 & 0.00393 \\
1421: 13 & 0.02608 & 0.00424 & 0.02609 & 0.00398 & 0.02610 & 0.00394 \\
1422: 14 & 0.02581 & 0.00423 & 0.02582 & 0.00398 & 0.02583 & 0.00394 \\
1423: 15 & 0.02562 & 0.00424 & 0.02561 & 0.00398 & 0.02562 & 0.00395 \\
1424: 16 & 0.02548 & 0.00425 & 0.02546 & 0.00400 & 0.02547 & 0.00396 \\
1425: 17 & 0.02541 & 0.00426 & 0.02539 & 0.00401 & 0.02539 & 0.00397 \\
1426: 18 & 0.02532 & 0.00427 & 0.02529 & 0.00401 & 0.02530 & 0.00398 \\
1427: 19 & 0.02528 & 0.00427 & 0.02524 & 0.00402 & 0.02525 & 0.00398 \\
1428: 20 & 0.02522 & 0.00428 & 0.02517 & 0.00403 & 0.02518 & 0.00399 \\
1429: \hline
1430: \multicolumn{7}{|c|}{$k_1=0.75$, 
1431: Ref.\ \cite{kwh}: $\sigma_{11}=0.044$, $\sigma_{12}=0.0043$} \\ \hline
1432: 12 & 0.04412 & 0.00441 & 0.04412 & 0.00424 & 0.04413 & 0.00422 \\
1433: 13 & 0.04345 & 0.00440 & 0.04344 & 0.00422 & 0.04345 & 0.00421 \\
1434: 14 & 0.04318 & 0.00440 & 0.04317 & 0.00423 & 0.04318 & 0.00421 \\
1435: 15 & 0.04280 & 0.00440 & 0.04278 & 0.00423 & 0.04279 & 0.00421 \\
1436: 16 & 0.04269 & 0.00440 & 0.04265 & 0.00423 & 0.04266 & 0.00422 \\
1437: 17 & 0.04252 & 0.00441 & 0.04248 & 0.00424 & 0.04249 & 0.00423 \\
1438: 18 & 0.04246 & 0.00442 & 0.04240 & 0.00425 & 0.04241 & 0.00423 \\
1439: 19 & 0.04238 & 0.00442 & 0.04232 & 0.00426 & 0.04232 & 0.00424 \\
1440: 20 & 0.04232 & 0.00442 & 0.04225 & 0.00426 & 0.04226 & 0.00424 \\
1441: \hline
1442: \multicolumn{7}{|c|}{$k_1=0.80$, 
1443: Ref.\ \cite{kwh}: $\sigma_{11}=0.063$, $\sigma_{12}=0.0047$} \\ \hline
1444: 12 & 0.06572 & 0.00475 & 0.06571 & 0.00467 & 0.06572 & 0.00467 \\
1445: 13 & 0.06573 & 0.00481 & 0.06571 & 0.00473 & 0.06572 & 0.00473 \\
1446: 14 & 0.06518 & 0.00483 & 0.06515 & 0.00475 & 0.06517 & 0.00475 \\
1447: 15 & 0.06488 & 0.00485 & 0.06484 & 0.00477 & 0.06486 & 0.00477 \\
1448: 16 & 0.06457 & 0.00486 & 0.06452 & 0.00478 & 0.06453 & 0.00478 \\
1449: 17 & 0.06440 & 0.00487 & 0.06433 & 0.00479 & 0.06435 & 0.00479 \\
1450: 18 & 0.06427 & 0.00487 & 0.06420 & 0.00479 & 0.06422 & 0.00480 \\
1451: 19 & 0.06418 & 0.00487 & 0.06409 & 0.00480 & 0.06411 & 0.00480 \\
1452: 20 & 0.06412 & 0.00488 & 0.06402 & 0.00480 & 0.06404 & 0.00480
1453: \end{tabular}
1454: \end{table}
1455: 
1456: 
1457: \begin{table}[tbp]
1458: \caption{Phase shifts (in radians) 
1459: of $e^++H->e^++H$ elastic scattering below
1460: the positronium formation threshold.}
1461: \label{tabshift}
1462: \begin{tabular}{lccccc}
1463: $k$&  Ref.\ \cite{5} &  Ref.\ \cite{20} &  Ref.\ \cite{14} & 
1464: Ref.\ \cite{kwh} & This work \\ \hline
1465: 0.1 &  0.1483 & 0.152 & 0.149 &  0.149 &  0.1480 \\
1466: 0.2 &  0.1877 & 0.188 & 0.188 &  0.189 &  0.1876 \\
1467: 0.3 &  0.1677 & 0.166 & 0.166 &  0.169 &  0.1673 \\
1468: 0.4 &  0.1201 & 0.118 & 0.120 &  0.121 &  0.1199 \\
1469: 0.5 &  0.0624 & 0.061 & 0.060 &  0.062 &  0.0625 \\
1470: 0.6 &  0.0039 & 0.003 &       &  0.003 &  0.0038 \\
1471: 0.7 & -0.0512 &-0.053 &       & -0.050 & -0.0513
1472: \end{tabular}
1473: \end{table}
1474: 
1475: 
1476: \begin{table}[tbp]
1477: \caption{Partial cross sections (in $\pi a_0^2$) 
1478: in the $H(n=2)-Ps(n=2)$ gap
1479: (threshold energies 0.7496-8745 Ry). Numbers $1$,$2$,$3$ and $4$ denote the
1480: channels $e^++H(1s)$, $e^++H(2s)$, $e^++H(2p)$ and  $p^++Ps(1s)$,
1481: respectively.}
1482: \label{4channel}
1483: \begin{tabular}{lccccc}
1484: $E_1$(Ry) &  & $\sigma_{11}$  & $\sigma_{12}$  & $\sigma_{13}$   
1485:  & $\sigma_{14}$\\ \hline
1486: 0.77 & Ref.\ \cite{hu99} & 0.090  & 0.000702 & 0.000454 & 0.00572 \\
1487: 0.77 & This work         & 0.0951 & 0.000673 & 0.000331 & 0.00558 \\ \hline
1488: 0.80 & Ref.\ \cite{hu99} & 0.096  & 0.00115  & 0.000364 & 0.00585 \\
1489: 0.80 & This work         & 0.1010 & 0.00127  & 0.000371 & 0.00563 \\ \hline
1490: 0.83 & Ref.\ \cite{hu99} & 0.0993 & 0.00170  & 0.000885 & 0.00581 \\
1491: 0.83 & This work         & 0.1063 & 0.00163  & 0.000813 & 0.00566 \\ \hline
1492: 0.84 & Ref.\ \cite{hu99} & 0.101  & 0.00190  & 0.00113 & 0.00580 \\
1493: 0.84 & This work         & 0.1080 & 0.00173  & 0.00105 & 0.00566 
1494: \end{tabular}
1495: \end{table}
1496: 
1497: 
1498: 
1499: \begin{figure}
1500: \psfig{file=vs.eps,width=8.5cm,angle=-90}
1501: \caption{The short-range part $v^{(s)}$ of the $-1/x$ attractive
1502: Coulomb potential. }
1503: \label{vs}
1504: \end{figure}
1505: 
1506: \begin{figure}
1507: \psfig{file=vl.eps,width=8.5cm,angle=-90}
1508: \caption{The long range part $v^{(l)}$ of the $-1/x$ attractive
1509: Coulomb potential. }
1510: \label{vl}
1511: \end{figure}
1512: 
1513: 
1514: \begin{figure}
1515: \psfig{file=cont_merkuriev.eps,width=8.5cm}
1516: 
1517: \caption{Analytic structure of $g_{x_\alpha }(z-z^\prime)\;
1518: g_{y_\alpha}(z^\prime)$ as a function of $z^\prime$ with
1519: $z=E+{\mathrm{i}}\varepsilon$, $E<0$, $\varepsilon>0$.
1520: The contour $C$ encircles the continuous spectrum of
1521: $h_{y_\alpha}$. A part of it, which goes on the unphysical
1522: Riemann-sheet of $g_{y_\alpha}$, is drawn by broken line.}
1523: 
1524: \label{fig1}
1525: \end{figure}
1526: 
1527: 
1528: \begin{references}
1529: 
1530: \bibitem{noble}  J.~V.~Noble, Phys. Rev. {\bf 161}, 945 (1967).
1531: 
1532: \bibitem{fm-book}   L.~D.~Faddeev and S.~P.~Merkuriev, {\it Quantum
1533: Scattering Theory for Several Particle Systems} 
1534: (Kluwer, Dordrecht,1993).
1535: 
1536: \bibitem{pzsc} Z.~Papp, Phys.~Rev.~C {\bf 55}, 1080 (1997).
1537: 
1538: \bibitem{pzwp}  Z.~Papp and W.~Plessas, Phys.~Rev.~C 
1539: {\bf 54}, 50 (1996).
1540: 
1541: \bibitem{zis} Z.~Papp, I.~N.~Filikhin and S.~L.~Yakovlev, 
1542: to be published in Few-Body Systems, nucl-th/9909083. 
1543: 
1544: \bibitem{pzatom} Z.~Papp,  Few-Body Systems, {\bf 24} 263 (1998).
1545: 
1546: \bibitem{vanzani} V.~Vanzani,
1547: {\it Few-Body Nuclear Physics}, (IAEA Vienna), 57 (1978).
1548: 
1549: \bibitem{redu}  E.~O.~Alt, P.~Grassberger, 
1550: and W.~Sandhas, Nucl.~Phys.\ {\bf B 2}, 167 (1967).
1551: 
1552: \bibitem{gloeckle} W.~Gl\"ockle, Nucl. Phys. {\bf A141}, 620 (1970); ibid. 
1553: {\bf A158}, 257 (1970).
1554: 
1555: \bibitem{yakovlev}  S.~L.~Yakovlev, Theor. Math. Phys. {\bf 107}, 835 (1996).
1556: 
1557: \bibitem{papp1} Z.~Papp, J.~Phys.\  A {\bf 20}, 153 (1987).  
1558: 
1559: \bibitem{papp2} Z.~Papp, Phys.\ Rev.\ C {\bf 38}, 2457 (1988).
1560: 
1561: \bibitem{papp3} Z.~Papp,  Phys.\ Rev.\ A {\bf 46}, 4437 (1992).
1562: 
1563: \bibitem{klp} B.~K\'onya, G.~L\'evai, and Z.~Papp,   
1564: Phys.\ Rev.\ C 61, 034302 (2000).
1565: 
1566: \bibitem{rotenberg} M.~Rotenberg,  
1567:  Ann.\ Phys.\ (N.Y.) {\bf 19}, 262  (1962); 
1568:  Adv.\ At.\ Mol.\ Phys.\ {\bf 6}, 233  (1970).
1569: 
1570: \bibitem{jmp} B.~K\'onya, G.~L\'evai, and Z.~Papp, J.\ Math.\ Phys.
1571: {\bf 38}, 4832 (1997).
1572: 
1573: \bibitem{lorentzen}  L.~Lorentzen and H.~Waadeland, {\it  Continued
1574: Fractions with Applications} (Noth-Holland, Amsterdam,   1992).
1575: 
1576: \bibitem{bb}  R.~Balian and E.~Br\'{e}zin,
1577:  Nuovo Cim.\  B {\bf 2}, 403 (1969).
1578: 
1579: \bibitem{sandhas} W.~Sandhas,
1580: {\it Few-Body Nuclear Physics}, (IAEA Vienna), 3 (1978).
1581: 
1582: \bibitem{kwh} A.~A.~Kvitsinsky, A.~Wu, and C.-Y.~Hu,
1583: J.\ Phys.\ B: At.\ Mol.\ Opt.\ Phys.\ {\bf 28} 275 (1995).
1584: 
1585: \bibitem{5} A.~K.~Bhatia, A.~Temkin, R.~J.~Drachman, and H.~Eiserike,
1586: Phys.\ Rev.\ A {\bf 3}, 1328 (1971).
1587: 
1588: \bibitem{20} F.~S.~Levin and J.~Shertzer, Phys.\ Rev.\  Lett.\ {\bf 61},
1589: 1089 (1988).
1590: 
1591: \bibitem{14} A.~A.~Kvitsinsky, J.~Carbonell, and C.~Gignoux,
1592: Phys.\ Rev.\ A {\bf 51}, 2997 (1995).
1593: 
1594: \bibitem{hu99} C.-Y.~Hu, Phys.\ Rev.\ A {\bf 59}, 4813 (1999).
1595: 
1596: \end{references}
1597: 
1598: 
1599: \end{document}
1600: 
1601: