nucl-th0201069/text
1: %  24.01.02
2: \documentclass[epj,final]{svjour}
3: \usepackage{psfig}
4: 
5: \begin{document}
6: %
7: \title{K$^+$ production in proton-nucleus reactions and the role of
8: momentum-dependent potentials}
9: \author{Z.
10: Rudy\inst{1}            \and W. Cassing\inst{2}   \and L. Jarczyk\inst{1}         \and B.
11: Kamys\inst{1}    \and P. Kulessa\inst{1,3}
12: %%%\thanks is optional - remove next line if not needed
13: %%%\thanks{\emph{Present address:} Insert the address here if needed}%
14: }                     % Do not remove
15: %
16: %
17: \institute{ M. Smoluchowski Institute of Physics, Jagellonian University,
18: PL-30059 Cracow, Poland \and
19: %
20: Institut f\"ur Theoretische Physik, Justus Liebig Universit\"at
21: Giessen, D-35392 Giessen, Germany    \and
22: %
23: Institut f\"ur Kernphysik, Forschungszentrum J\"ulich,
24: D-52425 J\"ulich, Germany }
25: %
26: %
27: \date{Received: date / Revised version: date}
28: % The correct dates will be entered by Springer
29: \abstract{ The production of $K^+$  mesons in proton-nucleus
30: collisions from 1.0 to 2.5 GeV  is analyzed with respect to
31: one-step nucleon-nucleon
32:  $(NN\rightarrow N Y K^+$)
33:  and two-step $\Delta$-nucleon $(\Delta N \rightarrow  K^+ Y N$) or
34:  pion-nucleon  $(\pi N \rightarrow  K^+ Y $)
35:  production channels on the basis of a coupled-channel transport
36:  approach (CBUU) including the kaon final state interactions.
37: The influence of momentum-dependent potentials for the nucleon,
38: hyperon and kaon in the final state are studied as well as the
39: importance of $K^+$ elastic rescattering in the target nucleus.
40: The transport calculations are compared to the experimental $K^+$
41: spectra taken at LBL Berkeley, SATURNE, CELSIUS, GSI and
42: COSY-J\"ulich. It is found that the momentum-dependent baryon
43: potentials effect the excitation function of the $K^+$ cross
44: section; at low bombarding energies of $\sim $ 1.0 GeV the
45: attractive baryon potentials in the final state lead to a relative
46: enhancement of the kaon yield whereas the net repulsive potential
47: at bombarding energies $\sim$ 2 GeV causes a decrease of the $K^+$
48: cross section. Furthermore it is pointed out, that especially the
49: $K^+$ spectra  at low momenta (or kinetic energy $T_K$) allow to
50: determine the in-medium $K^+$ potential almost model independently
51: due to a relative shift of the $K^+$ spectra in kinetic energy
52: that arises from the acceleration of the kaons when propagating
53: out of the nuclear medium to free space, i.e. converting potential
54: energy to kinetic energy of the free kaon.
55: %
56: \PACS{
57:       {13.60.Le}{Meson production}   \and
58:       {13.75.Jz}{Kaon-baryon interactions} \and
59:       {14.40.Aq}{Pi, K, and eta mesons}    \and
60:       {24.40.-h}{Nucleon-induced reactions}
61:      } % end of PACS codes
62: } %end of abstract
63: %
64: 
65: \maketitle
66: 
67: %
68: \section{Introduction}
69: The production of  mesons heavier than pions in $p+A$ reactions at
70: bombarding energies far below and close to the free
71: nucleon-nucleon threshold is of specific interest
72: \cite{1}-\cite{cass95} as one hopes to learn either about
73: cooperative nuclear phenomena and/or about high momentum
74: components of the nuclear many-body wave function that arise from
75: nucleon-nucleon correlations. Especially $K^+$ mesons have been
76: considered as promising hadronic probes \cite{5,12} due to the
77: rather moderate final state interaction, which is a consequence of
78:  strangeness conservation and the fact that there are no
79: baryon resonances with anti-strange quarks in nuclei.
80: Anti-hyperons, furthermore, have a much larger production
81: threshold and annihilate very fast in nuclei. On the other hand,
82: the kaon properties might change in the nuclear medium
83: \cite{Kaplan} such that conclusions on cooperative nuclear
84: phenomena require a precise understanding of the (anti-) kaon
85: potentials at finite nuclear density.
86: 
87: Experiments on $K^\pm$ production from nucleus-nucleus collisions
88: at SIS energies of 1-- 2 A$\cdot$ GeV have shown that in-medium
89: properties of the kaons are seen in the collective flow pattern of
90: $K^+$ mesons both, in-plane and out-of-plane, as well as in the
91: abundancy  of antikaons \cite{cass99,Li2001}. Thus in-medium
92: modifications of the mesons have become a topic of substantial
93: interest in the last decade triggered in part by the early
94: suggestion of Brown and Rho \cite{BR}, that the modifications of
95: hadron masses should scale with the scalar quark condensate
96: $<q\bar{q}>$ at finite baryon density.
97: 
98: As demonstrated in the pioneering work of Kaplan and Nelson
99: \cite{Kaplan} kaons and antikaons couple attractively to the
100: scalar nucleon density with a strength proportional to the $KN-
101: \Sigma$ constant,
102: \begin{equation} \label{sigmat}
103: \Sigma_{KN} = \frac{1}{2} (m^0_u + m^0_s) \ <N|\bar{u}u +
104: \bar{s}s|N>,
105: \end{equation}
106: which is not well known at present and may vary from 270 to 450
107: MeV. In (\ref{sigmat}) $m^0_u$ and $m^0_s$ denote the bare masses
108: for the light $u$- and strange $s$-quark while the expression in
109: the brackets is the expectation value of the scalar light and
110: strange quark condensate for the nucleon \cite{cass99}.
111: Furthermore, a vector coupling to the quark 4-current -- for
112: vanishing spatial components -- leads to a repulsive potential
113: term for the kaons; on the other hand this (Weinberg-Tomozawa)
114: term is attractive for the antikaons.
115: 
116: In chiral effective theories the dispersion relation for kaons and
117: antikaons in the nuclear medium -- for low momenta --  can be
118: written as \cite{nelson}
119: \begin{eqnarray}
120: &&\omega_K(\rho_N,{\bf p}) = + \frac{3}{8} \frac{\rho_N}{f_\pi^2}
121: \label{nelson1}\\ && + \left[ {\bf p}^2 + m_K^2 \left( 1 -
122: \frac{\Sigma_{KN}}{f_\pi^2 m_K^2} \rho_s + \left( \frac{3\rho_N}{8
123: f_\pi^2 m_K} \right)^2 \right) \right]^{1/2} , \nonumber \\
124: &&\omega_{\bar K} (\rho_N,{\bf p}) =  - \frac{3}{8}
125: \frac{\rho_N}{f_\pi^2}\label{nelson2} \\ &&+ \left[ {\bf p}^2 +
126: m_K^2 \left( 1 - \frac{\Sigma_{KN}}{f_\pi^2 m_K^2} \rho_s + \left(
127: \frac{3\rho_N}{8 f_\pi^2 m_K} \right)^2 \right) \right]^{1/2}.
128: \nonumber\end{eqnarray} In equations (2), (3)  $m_K$ denotes the
129: bare kaon mass, $f_\pi \approx$ 93 MeV is the pion decay constant,
130: while $\rho_s$ and $\rho_N$ stand for the scalar and vector
131: nucleon densities, respectively. As shown in Ref.
132: \cite{Schaffner}, for $\Sigma_{KN}$= 450 MeV  one ends up with an
133: effective kaon potential which is close to zero below ordinary
134: nuclear matter density $\rho_0$ and becomes  more repulsive above
135: $\rho_0$. On the other hand, using $\Sigma_{KN}$= 270 MeV a
136: repulsive kaon potential of $\approx$ 25 MeV at normal nuclear
137: matter density is obtained. Note, that when extrapolating
138: (\ref{nelson2}) to 3$\rho_0$ and above, the antikaon mass becomes
139: very light. Thus antikaon condensates might occur at high baryon
140: density which, furthermore, are of great interest in the
141: astrophysical context \cite{GB1,muto,sahu01}.
142: 
143: However, the actual kaon and antikaon self energies (or
144: potentials) are quite a matter of debate -- due to higher order
145: terms in the chiral expansion -- especially for the antikaon
146: \cite{Gal,Lutz,Ramos} and the momentum-dependence of their self
147: energies  is widely unknown (except for a dispersion analysis in
148: Ref. \cite{Sib98}) since most Lagrangian models restrict to
149: $s$-wave interactions or only include additional $p$-waves. It is
150: thus mandatory to perform experimental studies of the (anti-) kaon
151: properties under well controlled conditions, e.g. in
152: proton-nucleus reactions, where one probes the (anti-) kaon self
153: energies  at normal nuclear matter density $\rho_0
154: \approx$ 0.16 fm$^{-3}$ and below. Furthermore, by gating on kaon momenta in
155: the laboratory, one might be able to obtain information on the
156: momentum dependence of the self energies, too.
157: 
158:  $K^+$ production in $p+A$ collisions at subthreshold energies
159: has been observed experimentally more than a decade ago by Koptev
160: et al. \cite{12} at bombarding energies from 0.8--1.0 GeV.
161: However, only total $K^+$ yields could be extracted at that time.
162: Nevertheless, the target-mass dependence of the $K^+$ yield ($\sim
163: A$) suggested the dominance of two-step reactions with an
164: intermediate pion or $\Delta$. Detailed folding-model calculations
165: in Refs. \cite{15,cass95} essentially came to the same conclusion.
166: First differential $K^+$ spectra from $p+NaF$ and $p+Pb$ reactions
167: from the LBL Berkeley had been performed at $T_{lab}$ = 2.1 GeV
168: \cite{schnetzer}, i.e. far above the $NN$ threshold of 1.58 GeV in
169: free space. Only in more recent years differential $K^+$ spectra
170: have been measured down to 1.2 GeV for $^{12}C$ targets at 40 $^0$
171: \cite{debowski} (at SATURNE) or 90 $^0$ in the laboratory
172: \cite{badala} (at CELSIUS). Unfortunately, the different
173: experiments have no overlap in acceptance and the interpretation
174: of the data, if compatible at all, remains vague \cite{Markus}.
175: First data on the full momentum distribution at forward angles
176: have been presented very recently by the ANKE Collaboration at
177: COSY-J\"ulich \cite{Barsov} for $K^+$ mesons from $p+^{12}C$
178: reactions at 1.0 GeV \cite{ANKE}. These data show a kinematical
179: focussing of the spectra at finite momentum of $\approx$ 350
180: MeV/c, which appears incompatible with the cross section from Ref.
181: \cite{badala}. Thus a systematic comparison of all data is
182: urgently needed within an adequate theoretical approach that
183: allows to compare the kinematically restricted data on the same
184: footing.
185: 
186: 
187: In this study we use the coupled-channel (CBUU) transport model
188: that has first been developed in Ref. \cite{Wolf} for the
189: description of nucleus-nucleus collisions and later on employed
190: for the simulation of pion- and proton-nucleus reactions
191: \cite{Sib98,zibi95,cass98}, too. For applications to $K^\pm$
192: production in nucleus-nucleus collisions at SIS energies we refer
193: the reader to Ref. \cite{cass97}. In this model the effects of
194: momentum-dependent self energies for all hadrons can be studied
195: explicitly as well as their production and propagation in the
196: nuclear medium.
197: 
198: 
199: The paper is organized as follows: We briefly recapitulate the
200: ingredients of the CBUU model in Section 2, present the extensions
201: performed in this study and investigate in particular the effects
202: from $K^+$ rescattering. In section 3 we compare the transport
203: calculations  to the experimental spectra available from different
204: laboratories and explore the sensitivity of the $K^+$ spectra to
205: the momentum-dependent potentials employed. A summary and
206: discussion of open problems concludes this paper in Section 4.
207: 
208: 
209: \section{Ingredients of the transport model}
210: 
211: In this work we perform our analysis along the line of the
212: CBUU\footnote{Coupled-Channel Boltzmann-Uehling-Uhlenbeck}
213:  approach~\cite{Wolf} which is based on a
214: coupled set of  transport equations for the phase-space
215: distributions $f_{h} (x,p)$ of hadron $h$, i.e.
216: \cite{weber,Ehehalt}
217: \begin{eqnarray}
218: && \left( \Pi_{\mu}-\Pi_{\nu}\partial_{\mu}^p
219: U_{h}^{\nu} -M_{h}^*\partial^p_{\mu} U_{h}^{S}
220: \right) \partial_x^{\mu} f_{h}(x,p) \label{g24} \\
221: &&+ \left( \Pi_{\nu} \partial^x_{\mu}
222: U^{\nu}_{h}+ M^*_{h} \partial^x_{\mu}U^{S}_{h}\right)
223: \partial^{\mu}_p  f_{h}(x,p)   \nonumber\\
224: &&= \sum_{h_2 h_3 h_4\ldots} \int d2 d3 d4 \ldots
225:  [G^{\dagger}G]_{12\to 34\ldots} \nonumber \\
226: &&\phantom{a}\hspace*{2cm}\times\delta^4(\Pi +\Pi_2-\Pi_3-\Pi_4 \ldots ) \nonumber\\
227: && \times \left\{ f_{h_3}(x,p_3)f_{h_4}(x,p_4)\bar{f}_{h}(x,p)
228: \bar{f}_{h_2}(x,p_2) \right.\nonumber \\
229: && - \left. f_{h}(x,p)f_{h_2}(x,p_2)\bar{f}_{h_3}(x,p_3) \bar{f}_{h_4}(x,p_4)
230: \right\} \ldots\ \ .
231: \end{eqnarray}
232: In Eq.~(\ref{g24}) $U_{h}^{S}(x,p)$ and $U_{h}^{\mu}(x,p)$ denote
233: the real part of the scalar and vector hadron selfenergies,
234: respectively, while $[G^\dagger G]_{12\to 34\ldots} \delta^4 (\Pi
235: +\Pi_2-\Pi_3-\Pi_4\ldots )$ is the 'transition rate' for the
236: process $1+2\to 3+4+\ldots$ which is taken to be on-shell in the
237: semiclassical limit adopted. The hadron quasi-particle properties
238: in (\ref{g24}) are defined via the mass-shell constraint
239: \begin{equation}   \label{g25}
240: \delta (\Pi_{\mu}\Pi^{\mu}-M_{h}^{*2} ) \ \ ,
241: \end{equation}
242: with effective masses and momenta (in local Thomas-Fermi
243: approximation) given by \cite{weber}
244: \begin{eqnarray}\label{g26}
245: M_{h}^* (x,p)&=&M_h + U_h^{{S}}(x,p) \nonumber \\ \Pi^{\mu}
246: (x,p)&=&p^{\mu}-U^{\mu}_h (x,p)\ \ ,
247: \end{eqnarray}
248: while the phase-space factors
249: \begin{equation}
250: \bar{f}_{h} (x,p)=1 \pm f_{{h}} (x,p)
251: \end{equation}
252: are responsible for fermion Pauli-blocking or Bose enhancement,
253: respectively, depending on the type of hadron in the final/initial
254: channel. The dots in Eq.~(5) stand for further contributions to
255: the collision term with more than two hadrons in the final/initial
256: channels (cf. Ref. \cite{Cassing01}). The transport approach
257: (\ref{g24}) is fully specified by $U_{h}^{S}(x,p)$ and
258: $U_{h}^{\mu}(x,p)$ $(\mu =0,1,2,3)$, which determine the
259: mean-field propagation of the hadrons, and by the transition rates
260: $G^\dagger G\,\delta^4 (\ldots )$ in the collision term (5), that
261: describes the scattering and hadron production/absorption rates.
262: 
263: The scalar and vector mean fields $U_{h}^{S}$ and $U^\mu_{h}$ for
264: nucleons are modeled as in  Ref.~\cite{Ehehalt}, however, slightly
265: modified in line with Ref. \cite{excita}. In Fig. \ref{bild1} the
266: real part of the Schroedinger equivalent potential (SEP) for
267: nucleons,
268: \begin{equation}
269: U_{SEP}(p) = \Pi^0({\bf p}) - \sqrt{{\bf p}^2 + M_0^2}, \label{Ehg23}
270: \end{equation} is shown at density $\rho_0$ (full line) as a
271: function of the momentum  $p$ relative to the nuclear matter at
272: rest. Whereas we see a net attraction for momenta $p \leq$
273: 0.6~GeV/c, the nucleon potential becomes repulsive for higher
274: momenta and  reaches a maximum repulsion at $p \approx$ 1.5 GeV/c.
275: We mention that at density $\rho_0$ the Schroedinger equivalent
276: potential $U_{SEP}$ compares well with the potential from the data
277: analysis of Hama et al. \cite{Hama} as well as Dirac-Brueckner
278: computations from \cite{Mal} up to a kinetic energy $E_{kin}$ of 1
279: GeV \cite{Ehehalt}.
280: 
281: Apart from the nuclear potentials each charged hadron additionally
282: moves in the background of the Coulomb potential that is generated
283: by the charged hadrons themselves. In case of proton-nucleus
284: reactions -- with the nucleus at rest -- this essentially amounts
285: to a Coulomb acceleration in the final state for positively
286: charged hadrons. Note, that for heavy nuclei like $Pb$ or $Au$ the
287: Coulomb potential in the nuclear interior is about + 20 MeV, i.e.
288: of the same order of magnitude as the 'expected' repulsive $K^+$
289: nuclear potential.
290: 
291: \begin{figure}[h]
292: \centerline{\psfig{figure=fig1.eps,width=7.5cm}}
293:  \caption{The nucleon potential (solid line)  at density
294: $\rho_0$ as a function of momentum relative to the nuclear matter
295: at rest as used in the CBUU transport approach. The hatched area
296: denotes the  $K^+$ potential at $\rho_0$ (within the
297: uncertainties) from the dispersion analysis in Ref. \cite{Sib98}.}
298:  \label{bild1}
299: \end{figure}
300: 
301: 
302:  The hyperon mean fields, furthermore, are assumed to be 2/3
303: of the nucleon potentials. In the present approach, apart from
304: nucleons, $\Delta, N(1440)$, $N(1535)$, $\Lambda, \Sigma$ with
305: their isospin degrees of freedom, we propagate explicitly pions,
306: $K^+, K^-$, and $\eta$'s and assume that the pions as Goldstone
307: bosons do not change their properties in the medium; we also
308: discard self energies for the $\eta$-mesons which have a minor
309: effect on the kaon dynamics.  The kaon and antikaon potentials,
310: however, have to be specified explicitly.
311: 
312: \subsection{$K^+, K^-$ self energies}
313: 
314: Apart from (\ref{nelson1}), (\ref{nelson2}) there is a couple of
315: models for the kaon and antikaon
316: selfenergies~\cite{Kaplan,GB1,waas}, which differ in the actual
317: magnitude, however, agree on the relative signs for kaons and
318: antikaons. Thus in line with the kaon-nucleon scattering amplitude
319: the $K^+$ potential should be slightly repulsive at finite baryon
320: density whereas the antikaon should see an attractive potential in
321: the nuclear medium. Without going into a detailed discussion of
322: the various models (cf. Ref. \cite{Schaffner} and figures therein)
323: we adopt the more practical point of view, that the actual $K^+$
324: and $K^-$ self energies are unknown and as a guide for our
325: analysis use a linear extrapolation of the form,
326: \begin{equation}
327: \label{kmass} m^*_K(\rho_B,p) = m_K^0 \left(1 + \alpha(p)
328: \frac{\rho_B}{\rho_0}\right),
329: \end{equation}
330: with $\alpha(p)$ denoting a momentum-dependent coefficient.
331: Following the dispersion analysis of Sibirtsev et al. \cite{Sib98}
332: the coefficients $\alpha(p)$ can be modelled in line with the
333: $K^\pm$ potentials from Fig. 9 of \cite{Sib98}; the resulting kaon
334: potential $U_{K^+}$ at density $\rho_0$ is shown in Fig.
335: \ref{bild1} in terms of the hatched area and remains repulsive for
336: all momenta considered. In the momentum range of interest in this
337: study, i.e. from 0.1 -- 1.0 GeV/c, the $K^+$ potential at density
338: $\rho_0$ may be approximately represented by a constant of
339: $U_{K^+} \approx 20$ MeV taking into account the relative
340: uncertainty of $\pm$ 10 MeV from the analysis in Ref.
341: \cite{Sib98}. Since the antikaon dynamics has been investigated in
342: Ref. \cite{Sib98} for $p+A$ reactions, we skip a further
343: description of the actual implementation of the $K^-$ potential.
344: 
345: 
346: \subsection{Perturbative treatment of strangeness production}
347: The calculation of 'subthreshold' particle production has to be
348: treated perturbatively in the energy regime of interest due
349: to the small cross sections involved. Since we work within the
350: parallel ensemble algorithm, each parallel run of the transport
351: calculation can be considered approximately as an individual
352: reaction event, where binary reactions in the entrance channel at
353: given invariant energy $\sqrt{s}$ lead to final states with 2
354: (e.g. $K^+ Y$ in $\pi B$ channels), 3 (e.g. for $K^+ YN$ channels
355: in $BB$ collisions) or 4 particles (e.g. $K\bar{K}NN$ in $BB$
356: collisions) with a relative weight $P_i$ for each event $i$ which
357: is defined by the ratio of the  production cross section to the
358: total hadron-hadron cross section\footnote{The actual final states
359: are chosen  according to the 2, 3, or 4-body phase space.}. We
360: thus dynamically gate on all events where a $K^+ Y$ or $K^+ K^-$
361: pair is produced initially.  Each strange hadron production event
362: $i$ is represented by  $\sim 10^3$ testparticles for the final
363: strange hadron $j$ with individual weight $W_j^i$ such that the
364: sum of the weights $W_j^i$ over $j$ reproduces the individual
365: production probability $P_i$ and the distribution in momenta
366: (multiplied with the $NN$ or $\pi N$ cross section) describes the
367: differential production cross section. Then the 'perturbative'
368: hadrons are propagated according to the Hamilton equations of
369: motion including the potentials. Elastic and inelastic reactions
370: with pions, $\eta$'s or nonstrange baryons are computed in the
371: standard way~\cite{Wolf}.
372: 
373: The final differential cross sections are obtained by multiplying
374: each testparticle weight $W_j^i$ by the total inelastic $pA$ cross section
375: and gating on the experimental acceptance of the different
376: detectors. In this way one achieves a realistic simulation of the
377: strangeness production, propagation and reabsorption during the
378: proton-nucleus collision with sufficient statistics to allow for
379: selective cuts also at the low bombarding energy of 1.0 GeV,
380: where the total $K^+$ cross section is in the order of 1 $\mu b$
381: (for $Au$) or below (for $^{12}C$).
382: 
383: 
384: \subsection{Elementary cross sections}
385: For the present study the production of pions by $pN$ collisions
386: in $p+A$ reactions as well the total kaon cross sections in $pN$
387: and $\pi N$ collisions are of relevance. The pion production cross
388: section from $NN$ interactions is based on the parametrization of
389: the experimental data by Ver West and Arndt \cite{21} and
390: implemented in the transport model as described in Ref.
391: \cite{Wolf}. The cross sections for the channels $\pi N
392: \rightarrow K Y$, where $Y$ stands for a hyperon $\Lambda,
393: \Sigma$, are taken from the analysis of Huang et al. \cite{huang}
394: and essentially correspond to the experimental data for the
395: different $\pi N$ channels in vacuum (or 'free' space). Note, that
396: in addition to our early studies in \cite{15,cass95} the channels
397: with a $\Sigma$ hyperon are  taken into account. All cross
398: sections are reparametrized as a function of the invariant energy
399: above threshold $\sqrt{s}-\sqrt{s_0}$ \cite{cass99}, where
400: $\sqrt{s_0}$ denotes the threshold for the individual channel
401: given by the sum of the hadron masses in the final state of the
402: reaction.
403: 
404: Whereas the production cross section of kaons from $pN$ collisions
405: close to threshold has been essentially unknown about a decade
406: ago, the ambiguity in this cross section has been resolved
407: experimentally  by now \cite{COSY11} and more adequate
408: parametrizations of the cross sections can be employed. The
409: experimental data from Refs. \cite{COSY11,LB} on the $pp
410: \rightarrow K^+ + X$ reactions are displayed in Fig. \ref{bild2}
411: in comparison to the current approximation (solid line) and the
412: estimate from Zwermann \cite{zwermann} (dotted line) used earlier
413: \cite{15,cass95}. Thus the problem of 'free' cross sections for
414: $pN$ collisions is sufficiently under control. For $\Delta N$
415: collisions, however, no experimental data are available. We use
416: the same cross section as a function of the invariant energy
417: $\sqrt{s}$ as in the $pN$ case keeping in mind this basic
418: uncertainty.
419: 
420: \begin{figure}[h]
421: \centerline{\psfig{figure=fig2.eps,width=8.5cm,angle=270}}
422:  \caption{Comparison
423: of our parametrization for the inclusive $K^+$ production cross
424: section in $pp$ reactions (solid line) with the data from Refs.
425: \cite{COSY11,LB}. The dotted line corresponds to the
426: parametrization from Zwermann \cite{zwermann} used earlier in
427: Refs. \cite{15,cass95}. }
428:  \label{bild2}
429: \end{figure}
430: %
431: 
432: 
433: 
434: The question arises, furthermore, about the in-medium production
435: cross sections - essentially at density $\rho_0$ - if potentials
436: or self energies are involved. Here we employ the assumption that
437: the production matrix element squared $|M|^2$ does not change in
438: the medium and the change of cross section can be described by a
439: change of the available phase space. This assumption has been
440: employed for years in studies of hadron mass modifications in
441: nucleus-nucleus collisions \cite{cass99}. Since the experimental
442: production cross sections essentially are proportional to the
443: available phase space, it is sufficient to shift the threshold in
444: a $pN$ collision to
445: \begin{equation}
446: \sqrt{s_0^*} = \Pi^0_N(p_N) + \Pi^0_\Lambda (p_\Lambda) + \Pi^0_K
447: (p_K) \label{shift}
448: \end{equation}
449: using (\ref{g26}), where the momenta $p_N, p_\Lambda, p_K$ denote
450: the relative momenta with respect to the nuclear matter rest
451: frame.
452: 
453: \subsection{Folding model and its limitations}
454: We briefly recall the assumptions of the folding model, that is
455: used for the evaluation of momentum-dependent differential
456: production probabilities from secondary pion-nucleon collisions in
457: the CBUU approach and is described in more detail for
458: proton-nucleus reactions in \cite{15,Sib97,paryev}. The underlying
459: picture is that the proton impinging on a nucleus at a bombarding
460: energy $T_{lab}> 400$ MeV is producing a  meson $x$ with momentum
461: $k_x$ only in the first collision due to the available energy in
462: the reaction. The Lorentz-invariant differential cross section to
463: produce a meson in a primary proton-nucleon ($pN$) collision  is
464: given by \cite{Sib97}
465: \begin{equation}
466: \left(E_x {d^3 \sigma ^{NN}_x \over d^3k_x}\right)_{prim.} =
467:  \int d^3p d \omega
468:  \ \left(E'_x {d^3\sigma ^e_x(\sqrt{s}) \over d^3k'_x}\right)
469: \  S ({\bf p},\omega),
470: \label{fold1}
471: \end{equation}
472: \noindent where the Pauli-blocking factor for the final nucleon
473: states is neglected since kinematically the nucleons end up in an
474: unoccupied regime in momentum space at the bombarding energies of
475: interest. In eq. (\ref{fold1}) $S({\bf p},\omega)$ stands for the
476: target spectral function (normalized to 1) which can be taken from
477: experiment, e.g. Ref. \cite{Sick}, or parametrized accordingly. In
478: eq. (\ref{fold1}) the primed indices denote meson momenta in the
479: individual nucleon-nucleon cms frame which have to be
480: Lorentz-transformed to the detection frame. The quantity
481: $\sqrt{s}$ is the invariant energy of the individual NN system,
482: while the elementary differential cross section $d^3\sigma ^e_x
483: (\sqrt{s})/d^3k_x$ in (\ref{fold1}) is  parame\-trized according
484: to phase space (cf. \cite{15}) and assumed to be isotropic in the
485: $NN$ cms frame. The momentum differential production probability
486: per $NN$ collision is obtained from (\ref{fold1}) when dividing by
487: the total $NN$ cross section.
488: 
489: In order to obtain the inclusive differential kaon cross section
490: in a $p + A$  reaction within the folding model one has to
491: multiply the differential cross section (\ref{fold1}) by the
492: number of first-chance collisions $N_1(A)$. This number is
493: approximately given by the area of the target divided by the $p N$
494: cross section, i.e. $N_1(A) \approx \pi R^2_{target}/ \sigma_{p
495: N}$. To be more accurate one can use Glauber theory which leads to
496: $N_1 \approx  7.3$ for p + $^{12}C$ \cite{cass95,debowski}. The
497: contributions to the $K^+$ yield from secondary or further
498: sequential $NN$ collisions is discarded in the folding model
499: at subthreshold energies
500: due to the energy straggling of the impinging proton and due to
501: the fact that already the first-chance collisions only give a
502: minor contribution to the $K^+$ yield observed experimentally.
503: Note, however, that within the CBUU approach employed here the number of
504: collisions is calculated dynamically and thus no assumption or
505: separate model for $N_1(A)$ has to be invoked.
506: 
507: Apart from the primary reaction channels described above, the
508: first NN collision may also lead to the excitation of a $\Delta$-
509: resonance or even higher baryon resonances (e.g. N(1440), N(1535)
510: ..) which decay to nucleons and essentially pions due to their
511: short lifetimes of $\approx $ 1 fm/c or collide with another
512: nucleon before decaying. The $\Delta N$ or resonance-nucleon
513: reactions are treated in the CBUU approach dynamically as
514: described before. Energetic secondary pion-nucleon collisions,
515: however, suffer from very low statistics in transport simulations
516: of $p+A$ reactions. On the other hand, their contribution to the
517: kaon yield is expected to be high at subthreshold energies
518: \cite{cass95}.  We thus implement the secondary pion-nucleon
519: production channels for kaons following concepts of the folding
520: model.
521: 
522:  The differential cross section to produce a $K^+$ meson in a
523: collision of an on-shell pion with a nucleon from the target at
524: invariant energy $\sqrt{s}$ then is given by
525: \begin{equation}
526: \left(E_K {d^3 \sigma^{\pi N}_K \over d^3k}\right) = \sum_c \int
527: {d^3p d \omega}
528:  \ \left(E'_K {d^3\sigma ^{\pi N}_K(\sqrt{s}) \over
529:  d^3k'}\right)_c
530: \  S ({\bf p},\omega)
531: \label{fold1p}
532: \end{equation}
533: similar to (\ref{fold1}), where now the differential cross
534: sections for the reactions $\pi N \rightarrow K^+Y$ enter. Here,
535: the index $c$ denotes all individual channels while the hyperon
536: $Y$ stands again for a $\Lambda$ or $\Sigma$ baryon.
537: 
538: 
539: In order to evaluate the $K^+$ yield from
540: secondary $\pi N \rightarrow  K^+ Y$ collisions  in the folding model one folds
541: the primary pion distribution  -- given by the primary differential pion cross
542: section that is divided by the total $pN$ cross section $\sigma_{tot}$ -- with
543: the nucleon spectral function $S({\bf p}, \omega)$  and the
544: invariant production cross section, $i.e.$
545: \begin{eqnarray}
546: && \left(E_k \frac{d^3\sigma _K}{d^3k}\right)_{\sec .} =  \int
547: \int \frac{d^3p d\omega}{\sigma_{tot}}  \frac{d^3k'_\pi}{E'_\pi}
548:  S ({\bf p},\omega) \ g_\pi(A)  \label{fold2}\\
549: &&\times \left(E'_k {d^3\sigma _{\pi N\rightarrow Y K^+}({\bf
550: p,k_\pi} ) \over d^3k'} \right) \ \left(\frac{d^3 \sigma_{pN
551: \rightarrow \pi X}(\sqrt{s'})}{d^3 k'_\pi}\right)_{prim.}.
552: \nonumber
553: \end{eqnarray}
554: \noindent  In Eq. (\ref{fold2}) the single prime indices denote
555: the system of the intermediate pion and a target nucleon.
556: Moreover,
557: \newline $E'_\pi {d^3 \sigma_{pN \rightarrow \pi
558: X}(\sqrt{s'})}/{d^3 k'_\pi}$ stands for the $\pi$-meson
559: differential cross section, which is calculated in line with
560: (\ref{fold1}), while $\sigma_{tot}$ denotes the total
561: proton-nucleon cross section. The factor $g_\pi(A)$ in
562: (\ref{fold2}) accounts for the probability that the pion interacts
563: again with a target nucleon (cf. \cite{Sib97}). Note, that by Eq.
564: (\ref{fold2}) the intermediate pion is assumed to be on-shell
565: which might not hold for deep subthreshold kaon production. As in
566: case of the primary channel the expression (\ref{fold2}) has to be
567: multiplied by $N_1(A)$ in the folding model. Furthermore, some
568: estimate for the secondary rescattering probability $g_\pi(A)$ has
569: to be employed in the folding model as e.g. described in Ref.
570: \cite{PINOT}. As mentioned before, when using the differential
571: probabilities in a perturbative transport approach, the pion
572: rescattering probabilities are calculated dynamically without
573: approximation on $g_\pi(A)$.
574: 
575: The one- and two-step folding model has been used often in the
576: analysis of kaon or $\eta$-meson production
577: \cite{15,cass95,debowski,Sib97,paryev,Paryev2,Paryev3} initially
578: employing parametrized momentum distributions (cf. Refs.
579: \cite{15,cass95}) instead of spectral functions. They allow for an
580: estimate of differential cross sections in case of weakly
581: interacting probes and may serve as reference calculations for
582: more involved simulations employing all initial and final state
583: interactions during the reaction as the CBUU approach employed
584: here.
585: %
586: \begin{figure}[h]
587: \centerline{\psfig{figure=fig3.eps,width=8.5cm}}
588: \caption{Comparison
589: of the differential $K^+$ spectra for $p+Pb$ at 1.5 GeV at
590: different angles in the laboratory from 15$^0-80^o$ . The solid lines are obtained
591: from CBUU calculations including kaon rescattering whereas the
592: dashed lines are without rescattering.  }
593:  \label{bild3}
594: \end{figure}
595: 
596: 
597: \subsection{$K^+$ rescattering}
598: As indicated above, the folding model is useful for the evaluation
599: of total $K^+$ yields, however, becomes questionable in case of
600: differential spectra especially for heavy targets like $Au$ or
601: $Pb$  since kaon elastic rescattering cannot be described in a
602: straightforward manner. In order to show the influence of $K^+$
603: rescattering on the kaon spectra at different angles we show in
604: Fig. \ref{bild3} a comparison of our CBUU calculations with (solid
605: lines) and without (dashed lines) kaon rescattering for $p+Pb$ at
606: 1.5 GeV. As can be seen from Fig. \ref{bild3} the $K^+$ yield in
607: forward direction ($\theta \leq 15 ^0$) is  reduced by  up to a
608: factor of 2, while for large angles in the laboratory (80$^o$) the
609: kaon spectra become enhanced and also shifted to lower momenta in
610: the laboratory. Thus rescattering has to be included as a
611: necessary ingredient for the calculations if comparisons to
612: differential spectra are made on an absolute scale and especially,
613: if one attempts to extract kaon potentials from the spectral shape
614: (see below). We note that kaon rescattering thus will always be
615: included in the calculations to be shown in the following.
616: Furthermore, we mention that the results from the folding model
617: (\ref{fold1}), (\ref{fold2}) agree with the resulting spectra from
618: the CBUU calculation for $p+Pb$ and $p+C$ at 1.5 GeV within  30\%
619: when neglecting final state interactions as well as nuclear and
620: Coulomb potentials.
621: 
622: 
623: 
624: \section{Comparison to experimental data}
625: In this Section we compare our calculations to the experimental
626: $K^+$ spectra available from 1.0 GeV $-$ 2.5 GeV bombarding energy
627: on different targets.  In order to have an identical assignment of
628: lines in this section the dotted lines in Figs. \ref{bild4} - 9
629: correspond to CBUU calculations without baryon and kaon
630: potentials, the dashed lines show the results with baryon
631: potentials included while the solid lines reflect calculations
632: including both, nucleon and kaon potentials as specified in Fig.
633: \ref{bild1}. In all calculations, however, the Coulomb potential
634: will be included by default.
635: 
636: We start in Fig. \ref{bild4} with a comparison to the data of the
637: KaoS/SPES3 Collaboration for the differential $K^+$ spectra for
638: $p+Pb$ and $p+C$  at 1.5 GeV and $\theta_{lab}= 40 \pm$ 5$^0$ from
639: SATURNE \cite{debowski}. The experimental spectra for the $Pb$
640: target are seen to be described roughly within the error bars for
641: all calculations, i.e. with and without potentials, such that one
642: is not very sensitive to in-medium potentials at 1.5 GeV in the
643: momentum range above 350 MeV/c. For lower kaon momenta the
644: repulsive $K^+$ potential leads to a sizeable decrease (or shift)
645: in the spectra which can be attributed to the additional
646: acceleration of the kaons by the nuclear $K^+$ potential when
647: propagating from the nuclear interior to the vacuum. In case of
648: the $^{12}C$ target the effects from the momentum-dependent
649: nucleon potentials as well as from the $K^+$ potential are similar
650: to the $Pb$ target. It both cases the calculations without
651: potentials (dotted lines) slightly overestimate the data, but it
652: is not possible to draw already some conclusion on the actual size
653: of the $K^+$ potential since a single comparison might suffer from
654: systematic errors.
655: %
656: \begin{figure}[h]
657: \centerline{\psfig{figure=fig4.eps,width=8.5cm}}
658: \caption{Comparison of the CBUU calculations for the differential
659: $K^+$ spectra for $p+Pb$  and $p+C$  at 1.5 GeV and $\theta_{lab}=
660: 40 \pm$ 5$^0$ with the experimental data from Ref.
661: \cite{debowski}. The dotted lines are obtained from CBUU
662: calculations without baryon and kaon potentials, the dashed lines
663: show the results with baryon potentials included while the solid
664: lines correspond to calculations including both, nucleon and kaon
665: potentials. Note, that the effect of the repulsive kaon potential
666: is a reduction of the total $K^+$ yield as well as a shift of the
667: spectra at lower momenta. }
668:  \label{bild4}
669: \end{figure}
670: 
671: In Figs. \ref{bild5a} and \ref{bild5b} we compare  the CBUU calculations for the
672: differential $K^+$ spectra for $p+Pb$  and $p+NaF$
673:  at 2.1 GeV  with the experimental data from the LBL Berkeley
674: \cite{schnetzer} for different laboratory angles from 15 - 80$^0$.
675: At this bombarding energy the nucleon and $\Lambda$ final momenta
676: on average are above 0.6 GeV/c such that their potentials at
677: finite density (cf. Fig. \ref{bild1}) are repulsive. As a
678: consequence the calculated kaon yield decreases when including the
679: baryon potentials in the final state. Taking into account
680: additionally the repulsive $K^+$ potential decreases essentially
681: the spectrum for momenta below 250 MeV/c, but leaves the spectrum
682: practically unmodified for higher momenta since the relative
683: change of the $K^+$ energy by the kaon potential is only small.
684: %
685: \begin{figure}[h]
686: \centerline{\psfig{figure=fig5a.eps,width=8.5cm}}
687: \caption{Comparison of the CBUU calculations for the differential
688: $K^+$ spectra for $p+Pb$  at 2.1 GeV  with the experimental data
689: from Ref. \cite{schnetzer} at different angles in the laboratory.
690: The dotted lines are obtained from CBUU calculations without
691: baryon and kaon potentials, the dashed lines show the results with
692: baryon potentials included while the solid lines correspond to
693: calculations including both, nucleon and kaon potentials.}
694:  \label{bild5a}
695: \end{figure}
696: 
697: \begin{figure}[h]
698: \centerline{\psfig{figure=fig5b.eps,width=8.5cm}}
699:  \caption{Comparison
700: of the CBUU calculations for the differential $K^+$ spectra for
701:  $p+NaF$  at 2.1 GeV  with the
702: experimental data from Ref. \cite{schnetzer} at different angles
703: in the laboratory. The dotted lines are obtained from CBUU
704: calculations without baryon and kaon potentials, the dashed lines
705: show the results with baryon potentials included while the solid
706: lines correspond to calculations including both, nucleon and kaon
707: potentials.}
708:  \label{bild5b}
709: \end{figure}
710: %
711: 
712: In all approximations the experimental spectra are underestimated
713: by  factors $\sim 2-3$ at 15$^o$ and 35$^o$, which at first sight
714: might be attributed to an improper energy dependence of the
715: calculations. However, a comparison of the CBUU calculations  for
716: $p+Au$  with the preliminary data from Ref. \cite{scheinast}
717: (taken at GSI)  and $p+C$  at 2.5 GeV and $\theta_{lab}= 40 \pm$
718: 5$^0$ with the experimental data from SATURNE \cite{debowski} in
719: Fig. \ref{bild6} shows that these spectra  are overestimated by up
720: to a factor of 2--3 at higher kaon momenta.
721:  Note,  that the corresponding
722: data from Ref. \cite{scheinast} so far have to be considered as
723: preliminary. These findings suggest that the spectra from Ref.
724: \cite{schnetzer} are systematically too high by about factors of
725: $2-3$ or the data from the KaoS Collaboration too low (by about
726: the same factor).
727: \begin{figure}[h]
728: \centerline{\psfig{figure=fig6.eps,width=8.5cm}}
729:  \caption{Comparison
730: of the CBUU calculations for the differential $K^+$ spectra for
731: $p+Au$ (upper part) and $p+C$ (lower part) at 2.5 GeV  and
732: $\theta_{lab}= 40 \pm$ 5$^0$ with the experimental data from Refs.
733: \cite{scheinast,debowski}. The dotted lines are obtained from CBUU
734: calculations without baryon and kaon potentials, the dashed lines
735: show the results with baryon potentials included while the solid
736: lines correspond to calculations including both, nucleon and kaon
737: potentials.}
738:  \label{bild6}
739: \end{figure}
740: 
741: 
742: To further test the (over-) underprediction of the transport model
743: we show in Fig. \ref{bild7} a comparison of the CBUU calculations
744: for the differential $K^+$ spectra for $p+C$ at 1.2 GeV with the
745: experimental data from Ref. \cite{badala} at $\theta_{lab}= 90^0$
746: (open circles) taken at CELSIUS and $\theta_{lab}= 40 \pm 5^0$
747: (full squares) from the KaoS/SPES3 Collaboration \cite{debowski}
748: taken at SATURNE. In this particular case the spectra from Ref.
749: \cite{debowski} are slightly overestimated by the calculations
750: when neglecting the kaon potential while the spectra from Ref.
751: \cite{badala} at 90$^o$ are underestimated by about a factor of
752: 5-6 when neglecting the repulsive $K^+$ potential and by about an
753: order of magnitude for the repulsive kaon potential included,
754: which leads again to a sizeable decrease of the spectra at low
755: momentum. It is not clear to the authors where such discrepancies
756: might come from.
757: 
758: The CBUU calculations demonstrate that the kaon spectra at $90^o$
759: and $40^0$ are slightly enhanced (dashed lines)  when taking the
760: nucleon potential effects into account. Contrary to the
761: kinematical situation at $T_{lab}$= 2.1 (or 1.5) GeV the nucleon
762: and $\Lambda$ final momenta here on average are below 0.6 GeV/c,
763: where the potentials are attractive (cf. Fig. \ref{bild1}), such
764: that the $K^+$ production becomes enhanced (dashed lines) relative
765: to the free case (dotted lines). When including additionally the
766: overall repulsive kaon potential the $K^+$ spectrum drops again
767: (solid lines).
768: %
769: \begin{figure}[h]
770: \centerline{\psfig{figure=fig7.eps,width=8.5cm}}
771:  \caption{Comparison
772: of the CBUU calculations for the differential $K^+$ spectra for
773: $p+C$  1.2 GeV   with the experimental data from Ref.
774: \cite{badala} at $\theta_{lab}= 90^0 $ (open circles; lower lines)
775:  and $\theta_{lab}= 40
776: \pm 5^0$ \cite{debowski} (full squares; upper lines). The dotted
777: lines are obtained from CBUU calculations without baryon and kaon
778: potentials, the dashed lines show the results with baryon
779: potentials included while the solid lines correspond to
780: calculations including both, nucleon and kaon potentials.}
781:  \label{bild7}
782: \end{figure}
783: 
784: We now turn to the kinematical conditions of the ANKE experiments
785: at COSY-J\"ulich \cite{Barsov}, that have taken  $K^+$ spectra in
786: forward direction for $\theta_{lab} \leq 12^o$.  The calculated
787: differential $K^+$ spectra for $p+^{12}C$ at 1.0 GeV
788:  for $\theta_{lab} \leq 12^0$  are displayed in Fig.
789: \ref{bild8} in comparison to the data from Ref. \cite{ANKE}. The
790: dotted lines again are obtained from CBUU calculations without
791: baryon and kaon potentials, the dashed lines show the results with
792: baryon potentials included while the solid lines correspond to
793: calculations with both, nucleon and kaon potentials. At this low
794: bombarding energy the net attractive baryon potentials in the
795: final state enhance the $K^+$ yield by about a factor of 2 whereas
796: the additional repulsive $K^+$ potential leads again to a decrease
797: by a factor $\sim$ 3. The data from Ref. \cite{ANKE} are rather
798: well described by the calculations that include the baryon and
799: $K^+$ potentials (solid line), whereas the other limits clearly
800: fail. This might be considered as a first indication for the
801: observation of a repulsive $K^+$ potential in $p+A$ reactions,
802: however, a full systematics in target mass $A$ and laboratory
803: energy $T_{lab}$ will be needed to pin down this effect
804: unambiguously.
805: 
806: 
807: 
808: \begin{figure}[h]
809: \centerline{\psfig{figure=fig9new.eps,width=8.5cm}}
810:  \caption{The
811: calculated differential $K^+$ spectra for $p+C$ at 1.0 GeV for
812: $\theta_{lab} \leq 12^0$  within the acceptance of the ANKE
813: spectrometer in comparison to the data from \cite{ANKE}. The
814: dotted line is obtained from CBUU calculations without baryon and
815: kaon potentials, the dashed line shows the results with baryon
816: potentials included while the solid line corresponds to
817: calculations including both, nucleon and kaon potentials. The
818: shaded area indicates the contributions from the two-step
819: mechanisms $\Delta N \rightarrow K^+ Y N$ and $\pi N \rightarrow
820: K^+ Y$, respectively, for the case of nucleon and kaon potentials,
821: such that the difference to the solid line corresponds to the
822: primary $pN$ production channel. }
823:  \label{bild8}
824: \end{figure}
825: %
826: %
827: 
828: The shaded area in Fig. 9 indicates the contributions from the
829: two-step mechanisms $\Delta N \rightarrow K^+ Y N$ and $\pi N
830: \rightarrow K^+ Y$, respectively, for the case of nucleon and kaon
831: potentials included (solid line). Thus the role of secondary
832: ($\Delta$ and pion induced) reaction channels is clearly visible
833: from Fig. 9 by comparing the shaded area to the solid line, that
834: correspond to the total spectra. At $T_{lab}$ = 1.0 GeV the
835: dominant fraction of the $K^+$ yield is due to the secondary
836: channels in line with the earlier calculations in Refs.
837: \cite{15,cass95}. Consequently, one does not probe high momentum
838: components of the nuclear wave function by $K^+$ spectra in a
839: direct way.
840: 
841: Without explicit representation we mention that at $T_{lab}$ = 2.3
842: GeV the secondary channels in case of a $Au$ target amount to
843: about 30\% and in case of a $C$ target to less than 20\%. This
844: relative change with target mass number is attributed to the fact
845: that for the small $^{12}C$ target only a fraction of the high
846: energy pions rescatters in the target and produces $K^+ Y$ pairs.
847: Moreover, the role of the secondary channels decreases with
848: increasing kaon momentum such that the high momentum $K^+$ tail of
849: the spectra is dominated by the first chance $pN$ production
850: channel as found out before by Paryev \cite{paryev}.
851: 
852: 
853: 
854: Furthermore, it is worth to point out that the contribution of the
855: primary channel $pN \rightarrow K^+ Y N$ is enhanced by up to a
856: factor of 3 within the angular range of $\theta \leq 12^0$ as
857: compared to the angle integrated yield at the energy $T_{lab}$ =
858: 1.0 GeV. Thus the experimental mass dependence, when expressed in
859: terms of a scaling $\sim A^\alpha$, does not allow to disentangle
860: the relative contribution of the different reaction channels in a
861: satisfying manner for narrow cuts in the $K^+$ angular
862: distribution.
863: 
864: 
865: In order to provide some guideline for extrapolations between
866: experiments measuring $K^+$ spectra at different angles in the
867: laboratory we show in Fig. \ref{bild11} the angular distribution
868: of the kaons for momenta 0.2 GeV/c $\leq p_K \leq$ 0.5 GeV/c as
869: calculated within the CBUU approach for both, baryon and kaon
870: potentials for $p+ ^{12}C$ at $T_{lab}$ = 1.0, 1.2, 1.5, 1.8, 2.0
871: and 2.3 GeV. These angular distributions are rather flat within
872: the angular acceptance of the ANKE spectrometer of $\sim 12^0$,
873: however, drop substantially for angles larger than 40$^0$. Thus
874: our calculations (cf. also Fig. 8) do not support the idea of a
875: 'thermal' production mechanism for kaons in case of $p+A$
876: reactions.
877: 
878: 
879: 
880: \begin{figure}[h]
881: \centerline{\psfig{figure=fig11.eps,width=8.5cm}}
882:  \caption{The
883: calculated angular distribution of the $K^+$ spectra for
884: $p+C$ at 1.0, 1.2, 1.5, 1.8, 2.0 and 2.3 GeV
885: for $0.2 \leq p_K \leq 0.5$ GeV/c. The
886: solid lines are obtained from CBUU calculations including both,
887: nucleon and kaon potentials.}
888:  \label{bild11}
889: \end{figure}
890: %
891: 
892: \section{Summary}
893: 
894: In this work we have studied the production of $K^+$  mesons in
895: proton-nucleus collisions from 1.0 to 2.5 GeV   with respect to
896: one-step nucleon-nucleon and two-step $\Delta$-nucleon or
897: pion-nucleon production channels on the basis of a coupled-channel
898: transport approach (CBUU). We have included the kaon final state
899: interactions, which are important for heavy targets like $Au$ or
900: $Pb$, and explored the effects of momentum-dependent potentials
901: for the nucleon, hyperon and kaon in the nucleus. A comparison of
902: the transport calculations to the experimental $K^+$ spectra taken
903: at LBL Berkeley, SATURNE, CELSIUS, GSI and COSY-J\"ulich has shown
904: that the different data sets are not compatible with each other.
905: Thus no clear signal on in-medium potentials could be extracted
906: from our analysis in comparison to experimental data so far.
907: 
908: However, the detailed calculations  demonstrate that precise and
909: complete spectra show a substantial sensitivity to the potentials
910: and their momentum dependence. At low bombarding energies of
911: $\sim$ 1.0 GeV the net attractive potentials for the nucleon and
912: the $\Lambda$-hyperon in the final state lead to a relative
913: enhancement of the $K^+$ spectra while at higher bombarding
914: energies ($\sim$ 2 GeV) the baryon potentials are repulsive and
915: thus suppress $K^+$ production relative to the free case. This
916: phenomenon should be seen in the excitation function of the $K^+$
917: cross section when varying $T_{lab}$ from 1.0 -- 2.5 GeV.
918: Furthermore, the shape of the spectrum for low $K^+$ momenta in
919: the laboratory is very sensitive to both, Coulomb and nuclear kaon
920: potentials, since the kaons are accelerated by both forces when
921: leaving the nuclear environment and propagating to the continuum.
922: The relative strength of this momentum shift in the forward $K^+$
923: spectra is proportional to the square root of the sum of both
924: potentials, i.e. $\Delta p \approx \sqrt{2 M_K (U_{Coul}+U_K)}$ .
925: Thus the $K^+$ spectral shape at low momenta (or kinetic energies
926: $T_K$) allows to determine the strength of the $K^+$ potential
927: from experimental data in an almost model independent way
928: especially when comparing kaon spectra from light and heavy
929: targets at the same bombarding energy \cite{Barsov2} as a function
930: of $T_K$. Since most of the $K^+$ spectra measured so far have
931: been taken at higher momenta in the laboratory (except for Ref.
932: \cite{ANKE}) this finding opens up interesting perspectives for
933: the ANKE Collaboration at COSY-J\"ulich, which has performed a
934: systematic study of $K^+$ production in $p+A$ reactions down to
935: momenta of 150 MeV/c in the laboratory or $T_K \approx $ 23 MeV,
936: respectively.
937: 
938: \vspace{0.5cm} The authors like to acknowledge valuable
939: discussions with  M. B\"uscher, V. Koptev, M. Nekipelov, E. Ya.
940: Paryev, P. Senger, A. Sibirtsev, H. Str\"oher and C. Wilkin on
941: various issues of this study. Financial support has been provided
942: by the German BMBF under grant WTZ-POL 99/001 and the Polish State
943: Committee for Scientific Research under grant 2 P03B 101 19.
944: 
945: 
946: \begin{thebibliography}{99}
947: \bibitem{1}
948: P. Grimm, E. Grosse, { Prog. Part. Nucl. Phys.} {\bf 15}, 339
949: (1985)
950: \bibitem{2}
951:  P. Braun-Munzinger, J. Stachel, { Ann. Rev. Nucl. Part. Sci.} {\bf 37}, 1
952:  (1987)
953: \bibitem{3} H. Nifenecker, J. A. Pinston,  { Prog. Part. Nucl. Phys.}
954:  {\bf 23}, 271 (1989)
955: \bibitem{6}  J. Randrup, C. M.  Ko,  { Nucl. Phys. A} {\bf 343}, 519 (1980);
956:  A {\bf 411}, 537 (1983)
957: \bibitem{7} J. Cugnon, R. M.  Lombard,  {\ Nucl. Phys. A} {\bf 422}, 635 (1984)
958: \bibitem{10}
959:  S. V. Efremov, M. V. Kazarnovsky, E. Ya. Paryev,
960:  { Z. Phys. A} {\bf 344}, 181 (1992)
961: \bibitem{27}  V. I. Komarov et al., { Nucl. Phys. A} {\bf 326}, 397 (1979)
962: \bibitem{28}
963:  M. M. Nesterov, N. A. Tarasov, { Sov. Phys. JETP} {\bf 59}, 226 (1984)
964: \bibitem{29} N. A. Tarasov, V. P. Koptev, M. M. Nesterov,
965:  { Pis'ma Zh. Eksp. Teor. Fiz.} {\bf 43}, 217 (1986)
966: \bibitem{Mueller}  H. M\"uller, { Z. Phys. A} {\bf 339}, 409 (1991);
967:  H. M\"uller, K. Sistemich,  { Z. Phys. A} {\bf 344}, 197 (1992)
968: \bibitem{23}
969: J. Cugnon, P. Deneye, J. Vandermeulen,  {  Phys. Rev. C} {\bf 41},
970: 1339 (1990)
971: \bibitem{38b} A. Sibirtsev, M. B\"uscher, { Z. Phys. A} {\bf 347}, 191 (1994)
972: \bibitem{Teis}  S. Teis, W. Cassing, T. Maruyama, U. Mosel,
973: { Phys. Lett. B} {\bf 319}, 47 (1993)
974: 
975: \bibitem{kiselev} Yu. T. Kiselev et al., J. Phys. G {\bf 25}, 381
976: (1999).
977: %
978: \bibitem{Markus} M. B\"uscher et al., { Phys. Rev.} C {\bf 65}, 014603 (2001).
979: %
980: \bibitem{aichelin} J. Aichelin, C. M. Ko, { Phys. Rev. Lett.} {\bf 55},
981: 2661 (1985).
982: \bibitem{15}  W. Cassing et al.,
983:  { Phys. Lett. B} {\bf 238}, 25 (1990)
984: \bibitem{cass95} W. Cassing et al., { Z. Phys.} A {\bf 349}, 77 (1994)
985: \bibitem{5} W. Cassing, V. Metag, U. Mosel, K. Niita,
986:  { Phys. Rept.} {\bf 188}, 363 (1990)
987: \bibitem{12} V. P.  Koptev et al., { Sov. Phys. JETP} {\bf 67}, 2177 (1988)
988: \bibitem{Kaplan} D. B. Kaplan and A.E. Nelson,
989:  {  Phys. Lett. B} {\bf 175}, 57 (1986).
990: \bibitem{cass99}
991:     W. Cassing, E. L. Bratkovskaya, { Phys. Rept.} {\bf 308}, 65 (1999).
992: \bibitem{Li2001} Bao-An Li, A. T. Sustich, Bin Zhang, C. M. Ko,
993: Int. Jour. Mod. Phys. E {\bf 10}, 267 (2001).
994: \bibitem{BR}
995:     G. E. Brown, M. Rho, { Phys. Rev. Lett.} {\bf 66}, 2720 (1991).
996: \bibitem{nelson} A.E. Nelsen, D. Kaplan, { Phys. Lett.} B {\bf 192}, 193
997: (1987).
998: \bibitem{Schaffner}
999:     J. Schaffner-Bielich, I. N. Mishustin, J. Bondorf, { Nucl.
1000:     Phys. } A {\bf 625}, 325 (1997).
1001: \bibitem{GB1} G. E. Brown, C. M. Ko, Z. G. Wu,  L. H. Xia,
1002:  {Phys. Rev. C}  {\bf 43}, 1881 (1991).
1003: \bibitem{muto} T. Muto, { Nucl. Phys.} A {\bf 691}, 447 (2001).
1004: \bibitem{sahu01} P. K. Sahu, A. Ohnishi, { Nucl. Phys.} A {\bf 691}, 439 (2001).
1005: \bibitem{Gal} A. Gal, { Nucl. Phys.} A {\bf 691}, 268 (2001).
1006: \bibitem{Lutz} M. Lutz, { Phys. Lett.} B {\bf 426}, 12 (1998).
1007: \bibitem{Ramos} A. Ramos, S. Hirenzaki, S. S. Kamalov, T.T.S. Kuo,
1008: Y. Okumura, E. Oset, A. Polls, H. Toki, L. Tolos, { Nucl. Phys.} A
1009: {\bf 691}, 259 (2001).
1010: \bibitem{Sib98} A. Sibirtsev, W. Cassing, { Nucl. Phys.} A {\bf 641}, 476
1011: (1998).
1012: \bibitem{schnetzer} S. Schnetzer et al., { Phys. Rev.} C {\bf 40}, 640
1013: (1989).
1014: \bibitem{debowski} M. Debowski et al., { Z. Phys.} A {\bf 356}, 313 (1996).
1015: \bibitem{badala} A. Badala et al., { Phys. Rev. Lett.} {\bf 80}, 4863 (1998).
1016: \bibitem{Barsov} S. Barsov et al., Nucl. Instrum. Methods Phys.
1017: Res. A {\bf 462}, 364 (2001).
1018: \bibitem{ANKE} V. Koptev et al., { Phys. Rev. Lett.} {\bf 87}, 022310
1019: (2001).
1020: \bibitem{Wolf}  Gy. Wolf et al.,
1021:  { Nucl. Phys. A} {\bf 517}, 615 (1990); { Nucl. Phys. A} {\bf 552}, 549  (1993).
1022: \bibitem{zibi95} Z. Rudy et al., { Z. Phys. A} {\bf 354}, 445 (1996).
1023: \bibitem{cass98} Ye. S. Golubeva, L. A. Kondratyuk, W. Cassing, { Nucl. Phys.} A
1024: {\bf 625}, 832 (1997).
1025: \bibitem{cass97} W. Cassing et al., { Nucl. Phys. A} {\bf 614}, 415 (1997).
1026: \bibitem{weber} K. Weber et al., { Nucl. Phys. A} {\bf 539}, 713 (1992);
1027: T. Maruyama et al., { Nucl. Phys.} A {\bf 573}, 653 (1994).
1028: \bibitem{Ehehalt} W. Ehehalt, W. Cassing,
1029:   { Nucl. Phys. A}  {\bf 602}, 449 (1996).
1030: \bibitem{Cassing01}  W. Cassing, {\it nucl-th/0105069},
1031:   { Nucl. Phys. A}  {\bf 700}, 618 (2002).
1032: \bibitem{excita} W. Cassing, E. L. Bratkovskaya, S. Juchem, { Nucl.
1033: Phys.} A {\bf 674}, 249 (2000).
1034: \bibitem{Hama} S. Hama et al., { Phys. Rev. C} {\bf 41}, 2737 (1990)
1035: \bibitem{Mal} W. Botermans, R. Malfliet, { Phys. Rept.} {\bf 198}, 115
1036: (1990).
1037: \bibitem{waas} T. Waas, N. Kaiser,  W. Weise,
1038:   { Phys. Lett. B} {\bf 379}, 34 (1996).
1039: \bibitem{21}  B. J. Ver West, R. A. Arndt,  { Phys. Rev. C} {\bf 25}, 1979 (1982)
1040: \bibitem{huang} K. Tsushima, S. W. Huang, A. Faessler, { J. Phys.} G
1041: {\bf 21}, 33 (1995); { Phys. Lett.} B {\bf 337}, 245 (1994).
1042: %
1043: \bibitem{COSY11} J. T. Balewski et al., {Phys. Lett.} {\bf B420}, 211 (1998)
1044: %
1045: \bibitem{LB} H. Schopper (Editor), Landolt-B\"ornstein, New
1046: Series, Vol. I/12, Springer-Verlag, 1988.
1047: \bibitem{zwermann} W. Zwermann,  { Mod. Phys. Lett. A} {\bf 3}, 251 (1988)
1048: \bibitem{Sib97} A. Sibirtsev, W. Cassing, U. Mosel, { Z. Phys.} A {\bf 358}, 357
1049: (1997).
1050: \bibitem{paryev} E. Ya. Paryev, { Eur. Phys. J.} A {\bf 5}, 307 (1999).
1051: \bibitem{Sick} I. Sick, S. Fantoni, A. Fabrocini, O. Benhar, { Phys.
1052: Lett.} B {\bf 323}, 267 (1994).
1053: \bibitem{PINOT} E. Vercellin et al., { Nuovo Cimento} A {\bf 106},
1054: 861 (1993).
1055: \bibitem{scheinast} W. Scheinast {\it for the KaoS Collaboration},
1056: { Acta. Phys. Pol.} B {\bf 31}, 2305 (2000).
1057: \bibitem{Paryev2} E. Ya. Paryev, { Eur. Phys. J.} A {\bf 9}, 521 (2000).
1058: \bibitem{Paryev3} E. Ya. Paryev, preprint {\it INR - 1062/2001}.
1059: \bibitem{Barsov2} S. Barsov et al., Acta Phys. Pol. B {\bf 31},
1060: 2159 (2000).
1061: 
1062: \end{thebibliography}
1063: 
1064: 
1065: \end{document}
1066: 
1067: #!/bin/csh -f
1068: # Uuencoded gz-compressed .tar file created by csh script  uufiles
1069: # For more info (11/95), see e.g. http://xxx.lanl.gov/faq/uufaq.html
1070: # If you are on a unix machine this file will unpack itself: strip
1071: # any mail header and call resulting file, e.g., figures.uu
1072: # (uudecode ignores these header lines and starts at begin line below)
1073: # Then say        csh figures.uu
1074: # or explicitly execute the commands (generally more secure):
1075: #    uudecode figures.uu ;   gunzip figures.tar.gz ;
1076: #    tar -xvf figures.tar
1077: # On some non-unix (e.g. VAX/VMS), first use editor to change filename
1078: # in "begin" line below to figures.tar-gz , then execute
1079: #    uudecode figures.uu
1080: #    gzip -d figures.tar-gz
1081: #    tar -xvf figures.tar
1082: #
1083: uudecode $0
1084: chmod 644 figures.tar.gz
1085: gunzip -c figures.tar.gz | tar -xvf -
1086: rm $0 figures.tar.gz
1087: exit
1088: 
1089: