1: %============================================================================%
2: %\documentstyle[prc,aps,eqsecnum,epsf]{revtex}
3: \documentstyle[preprint,prc,aps,eqsecnum,epsf]{revtex}
4: \tighten
5: \begin{document}
6: \draft
7:
8: \title{Parity-Violating Interaction Effects I: the
9: Longitudinal Asymmetry in $p$$p$ Elastic Scattering}
10: %
11: \author{J.\ Carlson}
12: \address{Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545}
13: \author{R.\ Schiavilla}
14: \address{Jefferson Lab, Newport News, Virginia 23606 \\
15: and \\
16: Department of Physics, Old Dominion University, Norfolk, Virginia 23529}
17: \author{V.R.\ Brown}
18: \address{
19: Department of Physics, Massachusetts Institute of Technology,
20: Cambridge, MA 02139 \\
21: and \\
22: Department of Physics, University of Maryland, College Park, MD
23: 20742
24: }
25: \author{B.F.\ Gibson}
26: \address{Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545}
27: \date{\today}
28: %
29: \maketitle
30:
31: \begin{abstract}
32: The proton-proton parity-violating longitudinal
33: asymmetry is calculated in the lab-energy range 0--350 MeV, using
34: a number of different, latest-generation strong-interaction
35: potentials--Argonne $v_{18}$, Bonn-2000, and Nijmegen-I--in combination
36: with a weak-interaction potential consisting of $\rho$- and $\omega$-meson
37: exchanges--the model known as DDH. The complete scattering problem
38: in the presence of parity-conserving, including Coulomb,
39: and parity-violating potentials is solved in both configuration- and
40: momentum-space. The predicted parity-violating asymmetries are found
41: to be only weakly dependent upon the input strong-interaction potential adopted
42: in the calculation. Values for the $\rho$- and $\omega$-meson
43: weak coupling constants $h^{pp}_\rho$ and $h^{pp}_\omega$
44: are determined by reproducing the measured asymmetries at 13.6 MeV,
45: 45 MeV, and 221 MeV.
46: \end{abstract}
47: \pacs{21.30.+y, 24.80.-x, 25.40.Cm}
48:
49: \section{Introduction}
50: \label{sec:intro}
51:
52: A new generation of experiments have recently
53: been completed, or are presently under
54: way or in their planning phase to study the effects
55: of parity-violating (PV) interactions in $p$$p$
56: elastic scattering~\cite{Berdoz01}, $n$$p$ radiative
57: capture~\cite{Snow00} and deuteron electro-disintegration~\cite{Jlab01}
58: at low energies. There is also considerable interest
59: in determining the extent to which PV interactions can
60: affect the longitudinal asymmetry measured by the
61: SAMPLE collaboration in quasi-elastic scattering
62: of polarized electrons off the deuteron~\cite{Hasty00}, and
63: therefore influence the extraction from these data
64: (and those on the proton~\cite{Spayde00}) of the nucleon's
65: strange magnetic and axial form factors at a four-momentum
66: transfer squared of $0.1$ (GeV/c)$^2$.
67:
68: The present is the first in a series of papers
69: dealing with the theoretical investigation of PV interaction
70: effects in two-nucleon systems: it is devoted to $p$$p$ elastic
71: scattering, and presents a calculation of the longitudinal
72: asymmetry induced by PV interactions in the lab-energy range
73: 0--350 MeV.
74:
75: The available experimental data on the $p$$p$ longitudinal
76: asymmetry is rather limited. There are two measurements
77: at 15 MeV~\cite{Nagle79} and 45 MeV~\cite{Balzer84},
78: which yielded asymmetry values of $(-1.7\pm 0.8)\times 10^{-7}$ and
79: $(-2.3\pm 0.9)\times 10^{-7}$, respectively, as well as more
80: precise measurements at 13.6 MeV~\cite{Eversheim91},
81: 45 MeV~\cite{Kistryn87}, and 221 MeV~\cite{Berdoz01} yielding
82: $(-0.95\pm 0.15)\times 10^{-7}$,
83: $(-1.50\pm 0.23)\times 10^{-7}$, and
84: $(+0.84\pm 0.29)\times 10^{-7}$,
85: respectively, and finally a measurement
86: at 800 MeV in Ref.~\cite{Yuan86}, which produced an asymmetry
87: value of $(+2.4\pm 1.1)\times 10^{-7}$.
88:
89: The theoretical (and, in fact, experimental) investigation
90: of PV effects induced by the weak interaction in the $p$$p$ system
91: began with the prediction by Simonius~\cite{Simonius72} that
92: the longitudinal asymmetry would have a broad maximum at energies
93: close to 50 MeV, and that, being dominated
94: by the $J$=0 partial waves, it would be essentially independent
95: of the scattering angle. A number of theoretical studies of varying
96: sophistication followed~\cite{Brown74,Henley75,Oka81}, culminating in the study
97: by Driscoll and Miller~\cite{Driscoll89}, who used a distorted-wave
98: Born-approximation (DWBA) formulation of the PV scattering amplitude in
99: terms of exact wave functions obtained from solutions of
100: the Schr\"odinger equation with Coulomb and strong interactions.
101: In fact, the authors of Ref.~\cite{Driscoll89} investigated the
102: sensitivity of the calculated asymmetry to a number of
103: realistic strong-interaction potentials constructed by the late 1980's.
104: The model adopted for the PV weak-interaction potential, however,
105: was that developed by Desplanques and collaborators~\cite{Desplanques80},
106: the so-called DDH model. In the $p$$p$ sector, this potential
107: is parametrized in terms of $\rho$- and $\omega$-meson exchanges, in which
108: the PV $N$$N$$\rho$ and $N$$N$$\omega$ weak coupling constants
109: are calculated in a quark model approach incorporating symmetry
110: techniques like SU(6)$_W$ and current algebra requirements.
111: Factoring in the limitations inherent to such an approach, however,
112: the authors of Ref.~\cite{Desplanques80} gave rather wide ranges
113: of uncertainty for these weak coupling constants.
114:
115: The present work sharpens and updates that of Ref.~\cite{Driscoll89}.
116: It adopts the DDH model for the PV weak-interaction potential, but
117: uses the latest generation of realistic, parity-conserving (PC)
118: strong-interaction potentials, the Argonne $v_{18}$~\cite{Wiringa95},
119: Nijmegen I~\cite{Stoks94}, and CD-Bonn~\cite{Machleidt01}.
120: Rather than employing the DWBA scheme of Ref.~\cite{Driscoll89}
121: to calculate the PV component of the $p$$p$ elastic scattering
122: amplitude, it solves the complete scattering problem in the presence
123: of these PC and PV potentials (including the Coulomb potential),
124: in either configuration- or momentum-space, depending on whether
125: the Argonne $v_{18}$ and Nijmegen I or CD-Bonn models are used.
126: Such an approach allows us to obtain the PC and PV wave functions
127: explicitly. While this is unnecessary for the calculation reported
128: here--the DWBA estimate, along the lines of Ref.~\cite{Driscoll89},
129: of the PV component of the $p$$p$ amplitude should suffice--it becomes
130: essential for the studies of $n$$p$ radiative
131: capture and deuteron electro-disintegration planned at a later stage.
132:
133: The remainder of the present paper is organized as follows.
134: In Sec.~\ref{sec:pots} the PC and PV potentials used in this work
135: are briefly described, while in Sec.~\ref{sec:app} a self-consistent
136: treatment of the $p$$p$ scattering problem is provided along with a
137: discussion, patterned after that of Ref.~\cite{Driscoll89},
138: of the Coulomb contributions to the longitudinal asymmetries
139: measured in scattering and transmission experiments.
140: In Sec.~\ref{sec:res} the results for the asymmetry are presented;
141: in particular, their sentitivity to changes in the values of the
142: weak coupling constants and/or short-range cutoffs at the
143: strong- and weak-interaction vertices is studied. Finally,
144: Sec.~\ref{sec:cons} contains some concluding remarks.
145: %
146: %
147: \section{Parity-Conserving and Parity-Violating Potentials}
148: \label{sec:pots}
149:
150: The parity-conserving (PC), strong-interaction potentials
151: used in the present work are the Argonne $v_{18}$ (AV18)~\cite{Wiringa95},
152: Nijmegen I (NIJ-I)~\cite{Stoks94}, and CD-Bonn (BONN)~\cite{Machleidt01}
153: models. The AV18 and NIJ-I potentials were fitted to the Nijmegen
154: database of 1992~\cite{Bergervoet90,Stoks93}, consisting of 1787 $p$$p$
155: data, and both produced $\chi^2$ per datum close to one. The latest version
156: of the charge-dependent Bonn potential, however, has been fit to the
157: 1999 database, consisting of 2932 $p$$p$ data, for which it gives a $\chi^2$
158: per datum of 1.01~\cite{Machleidt01}. The substantial increase in
159: the number of $p$$p$ data is due to the development of novel
160: experimental techniques--internally polarized gas targets and stored, cooled
161: beams. Indeed, using this technology, IUCF has produced a large
162: number of $p$$p$ spin-correlation parameters of very high precision,
163: see for example Ref.~\cite{Przewoski98}. It is worth noting that
164: the AV18 potential, as an example, fits the post-1992 and both pre- and
165: post-1992 $p$$p$ data with $\chi^2$'s of 1.74 and 1.35,
166: respectively~\cite{Machleidt01}. Therefore, while the quality of
167: their fits has deteriorated somewhat in regard to the
168: extended 1999-database, the AV18 and NIJ-I models can still be
169: considered \lq\lq realistic\rq\rq.
170:
171: These realistic potentials consist of a long-range part due to
172: one-pion exchange (OPE), and a short-range part either modeled by
173: one-boson exchange (OBE), as in the BONN and NIJ-I models, or
174: parameterized in terms of suitable functions of two-pion range
175: or shorter, as in the AV18 model. While these potentials are (almost)
176: phase-equivalent, they differ in the treatment of non-localities.
177: AV18 is local (in $LSJ$ channels), while BONN and NIJ-I have
178: strong non-localities. In particular, BONN has a non-local
179: OPE component. However, it has been known
180: for some time~\cite{Friar77}, and recently
181: re-emphasized in Ref.~\cite{Forest00}, that the
182: local and non-local OPE terms are related to each other
183: by a unitary transformation. Therefore, the differences between
184: local and non-local OPE cannot be of any consequence for the prediction
185: of observables, such as binding energies or electromagnetic form
186: factors, provided, of course, that three-body interactions and/or
187: two-body currents generated by the unitary transformation are
188: also included~\cite{Coon86}. This fact has been
189: demonstrated~\cite{Schiavilla01} in a calculation of
190: the deuteron structure function $A(q)$ and tensor observable $T_{20}(q)$,
191: based on the local AV18 and non-local BONN
192: models and associated (unitarily consistent) electromagnetic currents.
193: The remaining small differences between the calculated $A(q)$ and
194: $T_{20}(q)$ are due to the additional short-range non-localities
195: present in the BONN model. Therefore, provided
196: that consistent calculations--in the sense above--are performed,
197: present \lq\lq realistic\rq\rq potentials will lead to very similar
198: predictions for nuclear observables, at least to the extent that
199: these are influenced predominantly by the OPE component.
200:
201: As already mentioned in Sec.~\ref{sec:intro}, the form of
202: the parity-violating (PV) weak-interaction potential was derived
203: in Ref.~\cite{Desplanques80}--the DDH model:
204:
205: \begin{eqnarray}
206: v^{\rm PV}=\sum_{\alpha=\rho,\omega}&-&
207: \frac{g_\alpha \, h^{pp}_\alpha}{4\pi} \frac{m_\alpha}{m}
208: \Big\{ m_\alpha (1+\kappa_\alpha)Y^\prime (m_\alpha r)
209: ({\bbox \sigma}_1\times {\bbox \sigma}_2) \cdot \hat{\bf r} \nonumber \\
210: &+&({\bbox \sigma}_1 - {\bbox \sigma}_2) \cdot \left[ {\bf p} \, ,
211: \, Y(m_\alpha r)\right]_{+} \Big\} \>\>,
212: \label{eq:DDH}
213: \end{eqnarray}
214: where the relative position and momentum are defined
215: as ${\bf r}={\bf r}_1-{\bf r}_2$ and ${\bf p}=({\bf p}_1-{\bf p}_2)/2$,
216: respectively, $\left[ \dots \, , \, \dots \right]_{+}$ denotes
217: the anticommutator, and $m$ and $m_\alpha$ are the proton and vector-meson
218: ($\rho$ or $\omega$) masses, respectively. Note that
219: the first term in Eq.~(\ref{eq:DDH}) is usually written
220: in the form of a commutator, since
221:
222: \begin{equation}
223: {\rm i} \left[ {\bf p} \, , \, Y(m_\alpha r) \right]_{-}
224: =m_\alpha Y^\prime (m_\alpha r)\, \hat{\bf r} \>\>\>,
225: \end{equation}
226: where $Y^\prime(x)$ denotes its derivative ${\rm d}Y(x)/{\rm d}x$.
227: The Yukawa function
228: $Y(x_\alpha)$, suitably modified by the inclusion of
229: monopole form factors, is given by
230:
231: \begin{equation}
232: Y(x_\alpha)= \frac{1}{x_\alpha}\Bigg\{ {\rm e}^{-x_\alpha}
233: -{\rm e}^{(\Lambda_\alpha/m_\alpha) x_\alpha} \left[ 1+
234: \frac{1}{2}\frac{\Lambda_\alpha}{m_\alpha}
235: \left( 1-\frac{m^2_\alpha}{\Lambda^2_\alpha} \right)
236: x_\alpha \right] \Bigg\} \>\>,
237: \end{equation}
238: where $x_\alpha \equiv m_\alpha r$. Finally, the values
239: for the strong-interaction $\rho$- and $\omega$-meson vector
240: and tensor coupling constants $g_\alpha$ and $\kappa_\alpha$, as well as
241: for the cutoff parameters $\Lambda_\alpha$, are
242: taken from the BONN model~\cite{Machleidt01}, and
243: are listed in Table~\ref{tb:gs}. The weak-interaction
244: coupling constants $h^{pp}_\rho$ and $h^{pp}_\omega$
245: correspond to the following combinations of DDH parameters
246:
247: \begin{eqnarray}
248: h^{pp}_\rho&=&h_{\rho_0}+h_{\rho_1}+ \frac{h_{\rho_2}}{\sqrt{6}} \>\>, \\
249: h^{pp}_\omega&=&h_{\omega_0}+h_{\omega_1} \>\>.
250: \end{eqnarray}
251: Their values, reported in Table~\ref{tb:gs} in the column labeled
252: (DDH-adj), are obtained by
253: fitting the available data on the longitudinal asymmetry, see Sec.~\ref{sec:res}.
254: The values corresponding to the \lq\lq best\rq\rq estimates
255: for the $h_{\rho_i}$ and $h_{\omega_i}$ suggested in
256: Ref.~\cite{Desplanques80} are also listed in Table~\ref{tb:gs}
257: in the column (DDH-orig).
258: Indeed, one of the goals of the present work is to
259: study the sensitivity of the calculated longitudinal asymmetry
260: to variations in both the PV coupling constants and cutoff
261: parameters. In this respect, it should also be noted that,
262: in the limit $\Lambda_\rho$=$\Lambda_\omega$ and ignoring the
263: small mass difference between $m_\rho$ and $m_\omega$, were it not
264: for the different values of the tensor couplings $\kappa_\rho$ and
265: $\kappa_\omega$, the $\rho$- and $\omega$-meson terms
266: in $v^{\rm PV}$ would collapse to a single term of strength
267: proportional to $g_\rho h^{pp}_\rho + g_\omega h^{pp}_\omega$.
268: %
269: %
270: \section{Formalism}
271: \label{sec:app}
272:
273: In this section we discuss the $p$$p$ scattering problem
274: in the presence of a potential $\overline{v}$ given by
275:
276: \begin{equation}
277: \overline{v}= v^{\rm PC}+v^{\rm PV}+v^{\rm C} \>\>,
278: \end{equation}
279: where $v^{\rm PC}$ and $v^{\rm PV}$ denote the parity-conserving
280: and parity-violating components induced by the strong and weak
281: interactions, respectively, and $v^{\rm C}$ is the Coulomb
282: potential.
283:
284: \subsection{Partial-Wave Expansions of Scattering State, $T$- and $S$-Matrices}
285: \label{sec:pwest}
286:
287: The Lippmann-Schwinger equation for the $p$$p$ scattering
288: state $\mid{\bf p},SM_S\rangle^{(\pm)}$, where ${\bf p}$
289: is the relative momentum and $SM_S$ specifies
290: the spin state, can be written as~\cite{Goldberger64}
291:
292: \begin{equation}
293: \mid{\bf p},SM_S\rangle^{(\pm)}=
294: \mid {\bf p},SM_S\rangle^{(\pm)}_{\rm C} +
295: \frac{1}{E-H_0-v^{\rm C}\pm {\rm i}\epsilon}\, v\,
296: \mid{\bf p},SM_S\rangle^{(\pm)} \>\>,
297: \label{eq:LS}
298: \end{equation}
299: where $H_0$ is the free Hamiltonian, $v=v^{\rm PC}+v^{\rm PV}$, and
300: $\mid \dots\rangle^{(\pm)}_{\rm C}$ are the eigenstates of
301: $H_0+v^{\rm C}$,
302:
303: \begin{equation}
304: \left( E-H_0-v^{\rm C} \right)
305: \mid{\bf p},SM_S\rangle^{(\pm)}_{\rm C}=0 \>\>,
306: \end{equation}
307: with wave functions given by
308:
309: \begin{eqnarray}
310: \phi^{(\pm)}_{ {\bf p}, SM_S}({\bf r})&=&
311: \langle {\bf r}\mid {\bf p},SM_S\rangle^{(\pm)}_{\rm C} \nonumber \\
312: &=&4\pi\sqrt{2} \sum_{JM_JL} {\rm i}^L\, \epsilon_{LS} \,\,
313: {\rm e}^{\pm {\rm i}\sigma_L} \frac{F_L(\eta;pr)}{pr}
314: [Z_{LSM_S}^{JM_J}(\hat{\bf p})]^* \,
315: {\cal Y}_{LSJ}^{M_J}(\hat{\bf r}) \>\>.
316: \label{eq:pwe}
317: \end{eqnarray}
318: Here $F_L(\eta;\rho)$ denotes the regular Coulomb
319: wave function~\cite{Abramowitz74}, while the parameter $\eta$
320: and Coulomb phase-shift $\sigma_L$ are given by
321:
322: \begin{eqnarray}
323: \eta &=& \alpha \mu/p \>\>, \\
324: \sigma_L&=& {\rm arg}\left[\Gamma(L+1+{\rm i} \eta)\right] \>\>,
325: \end{eqnarray}
326: where $\alpha$ is the fine structure constant and $\mu$ is the
327: reduced mass. Finally, the following definitions have also
328: been introduced:
329:
330: \begin{equation}
331: Z_{LSM_S}^{JM_J}(\hat{\bf p})\equiv
332: \sum_{M_L} \langle LM_L,SM_S\mid JM_J\rangle
333: \, Y_{LM_L}(\hat {\bf p}) \>\>,
334: \end{equation}
335: \begin{equation}
336: \epsilon_{LS} \equiv \frac{1}{2} \Big[1 +(-1)^{L+S} \Big] \>\>.
337: \end{equation}
338: The factor $\epsilon_{LS}$ ensures that the wave functions $\phi^{(\pm)}$
339: are properly antisymmetrized. Note that in the limit $\eta$=0, equivalent
340: to ignoring the Coulomb potential, the latter reduce to (antisymmetrized)
341: plane waves,
342:
343: \begin{equation}
344: \phi^{(\pm)}_{ {\bf p}, SM_S}({\bf r}) \rightarrow \frac{1}{\sqrt{2}}
345: \Big[{\rm e}^{ {\rm i} {\bf p} \cdot {\bf r}} +(-)^{S}
346: {\rm e}^{-{\rm i} {\bf p} \cdot {\bf r}}\Big]
347: \chi^S_{M_S} \>\>.
348: \end{equation}
349:
350: The $\overline{T}$-matrix corresponding
351: to the potential $v+v^{\rm C}$ can be expressed as~\cite{Goldberger64}
352:
353: \begin{equation}
354: \overline{T}({\bf p}^\prime,S^\prime M_S^\prime;{\bf p},SM_S)=
355: T({\bf p}^\prime,S^\prime M_S^\prime;{\bf p},SM_S)+
356: T^{\rm C}({\bf p}^\prime,S^\prime M_S^\prime;{\bf p},SM_S) \>\>,
357: \label{eq:tmas}
358: \end{equation}
359: where $T^{\rm C}$ is the (known) $T$-matrix corresponding only
360: to the Coulomb potential~\cite{Goldberger64}, and
361:
362: \begin{equation}
363: T({\bf p}^\prime,S^\prime M_S^\prime;{\bf p},SM_S)=
364: ^{(-)}_{\rm C}\!\!\langle {\bf p}^\prime,S^\prime M_S^\prime
365: \mid T \mid {\bf p},SM_S\rangle^{(+)} \>\>.
366: \end{equation}
367: Insertion of the complete set of states
368: $\mid {\bf p},SM_S\rangle^{(-)}_{\rm C}$ into the
369: right-hand-side of the Lippmann-Schwinger equation leads to
370:
371: \begin{equation}
372: \mid{\bf p},SM_S\rangle^{(+)}=\mid{\bf p},SM_S\rangle^{(+)}_{\rm C}
373: +\sum_{S^\prime M_S^\prime}\int\frac{{\rm d}{\bf p}^\prime}{(2\pi)^3}
374: \frac{1}{2} \mid{\bf p}^\prime,S^\prime M_S^\prime\rangle^{(-)}_{\rm C}
375: \frac{T({\bf p}^\prime,S^\prime M_S^\prime;{\bf p},SM_S)}
376: {E-p^{\prime 2}/(2\mu) +{\rm i}\epsilon} \>\>,
377: \label{eq:LSa}
378: \end{equation}
379: from which the partial wave expansion of the scattering state
380: is easily obtained by first noting that
381: the potential, and hence the $T$-matrix, can be expanded as
382:
383: \begin{eqnarray}
384: ^{(-)}_{\rm C}\!\langle {\bf p}^\prime,S^\prime M_S^\prime
385: \mid v \mid {\bf p},SM_S\rangle^{(+)}_{\rm C} &=& 2(4\pi)^2\! \sum_{JM_J}
386: \sum_{L L^\prime}
387: \epsilon_{L^\prime S^\prime}\,\, \epsilon_{LS} \,
388: {\rm e}^{{\rm i}\sigma_{L^\prime}}
389: {\rm e}^{{\rm i}\sigma_L}
390: Z_{L^\prime S^\prime M_S^\prime}^{JM_J}(\hat{\bf p}^\prime) \nonumber \\
391: &&[Z_{LSM_S}^{JM_J}(\hat{\bf p})]^*\,
392: v^J_{L^\prime S^\prime,LS} (p^\prime;p) \>\>,
393: \label{eq:tmae}
394: \end{eqnarray}
395: with
396: \begin{equation}
397: v^J_{L^\prime S^\prime,LS}(p^\prime;p)
398: ={\rm i}^{L-L^\prime}\int {\rm d}{\bf r}
399: \frac{F_{L^\prime}(\eta;p^\prime r)}{p^\prime r}
400: {\cal Y}_{L^\prime S^\prime J}^{M_J \dagger}
401: v({\bf r})
402: {\cal Y}_{LSJ}^{M_J}
403: \frac{F_L(\eta;p r)}{p r} \>\>.
404: \end{equation}
405: After insertion of the corresponding expansion for the $T$-matrix
406: into Eq.~(\ref{eq:LSa}) and a number of standard manipulations,
407: the scattering-state wave function can be written as
408:
409: \begin{equation}
410: \psi^{(+)}_{ {\bf p}, SM_S}({\bf r})=
411: 4\pi\sqrt{2} \sum_{JM_J}\, \sum_{L L^\prime S^\prime}
412: {\rm i}^{L^\prime}\, \epsilon_{L^\prime S\prime}
413: \, \epsilon_{LS} \, {\rm e}^{{\rm i}\sigma_L}
414: \, [Z_{LSM_S}^{JM_J}(\hat{\bf p})]^*
415: \frac{w^J_{L^\prime S^\prime,LS}(r;p)}{r}
416: {\cal Y}_{L^\prime S^\prime J}^{M_J}(\hat{\bf r}) \>\>,
417: \end{equation}
418: with
419:
420: \begin{eqnarray}
421: \frac{w^J_{L^\prime S^\prime,LS}(r;p)}{r}&=&
422: \Bigg[ \delta_{L,L^\prime}\delta_{S,S^\prime}
423: \frac{F_{L^\prime}(\eta;pr)}{pr} \nonumber \\
424: &+&\frac{2}{\pi}\int_0^\infty {\rm d}p^\prime p^{\prime 2}
425: \frac{F_{L^\prime}(\eta;p^\prime r)}{p^\prime r}\frac{1}
426: {E-p^{\prime 2}/(2\mu)+{\rm i}\epsilon}
427: T^J_{L^\prime S^\prime,LS}(p^\prime;p) \Bigg] \>\>.
428: \end{eqnarray}
429: The (complex) radial wave function $w(r)$ behaves in the
430: asymptotic region $r \rightarrow \infty$ as
431:
432: \begin{equation}
433: \frac{w^J_{\alpha^\prime,\alpha}(r;p)}{r}\simeq \frac{1}{2} \Big[
434: \delta_{\alpha,\alpha^\prime} h^{(2)}_{L^\prime}(\eta;pr)
435: +h^{(1)}_{L^\prime}(\eta;pr) S^J_{\alpha^\prime , \alpha}(p) \Big] \>\>,
436: \label{eq:asy}
437: \end{equation}
438: where the label $\alpha$ ($\alpha^\prime$) stands for the set of quantum numbers
439: $LS$ ($L^\prime S^\prime$), the on-shell ($p^\prime=p$) $S$-matrix
440: has been introduced,
441:
442: \begin{equation}
443: S^J_{\alpha^\prime , \alpha}(p)=\delta_{\alpha,\alpha^\prime}
444: -4{\rm i}\, \mu p \, T^J_{\alpha^\prime , \alpha} (p;p) \>\>,
445: \label{eq:sma}
446: \end{equation}
447: and the functions $h^{(1,2)}(\eta;\rho)$ are defined in terms
448: of the regular and irregular ($G_L$) Coulomb functions as
449:
450: \begin{equation}
451: h^{(1,2)}_L(\eta;\rho)=\frac{F_L(\eta;\rho)\mp {\rm i}\,
452: G_L(\eta;\rho)}{\rho} \>\>.
453: \end{equation}
454: Again, in the limit $\eta =0$, $F_L(\eta;\rho)/\rho \rightarrow
455: j_L(\rho)$ and $G_L(\eta;\rho)/\rho \rightarrow -n_L(\rho)$, where
456: $j_L(\rho)$ and $n_L(\rho)$ are the spherical Bessel functions,
457: and the familiar expressions for the partial wave-expansion
458: of the scattering state, $S$- and $T$-matrices are
459: recovered~\cite{Goldberger64}.
460:
461: \subsection{Schr\"odinger Equation, Phase-Shifts, and Mixing Angles}
462: \label{sec:phase}
463:
464: The coupled-channel Schr\"odinger equations for the radial wave
465: functions $w(r)$ read:
466:
467: \begin{equation}
468: \Bigg[ -\frac{ {\rm d}^2}{{\rm d}r^2} + \frac{L(L+1)}{r^2}
469: -p^2 \Bigg] w^J_{\alpha^\prime,\alpha}(r;p)
470: +\sum_{\beta} r \, v^J_{\alpha^\prime,\beta}(r)\, \frac{1}{r}
471: w^J_{\beta,\alpha}(r;p) = 0 \>\>,
472: \label{eq:cschr}
473: \end{equation}
474: with
475:
476: \begin{equation}
477: v^J_{\alpha^\prime,\alpha}(r)={\rm i}^{L-L^\prime}\, 2\mu \,
478: \int{\rm d}\Omega\, {\cal Y}_{\alpha^\prime J}^{M_J \dagger}
479: \, v({\bf r})\, {\cal Y}_{\alpha J}^{M_J} \>\>,
480: \label{eq:vpe}
481: \end{equation}
482: where, because of time reversal invariance, the matrix
483: $v^J_{\alpha^\prime,\alpha}$ can be shown to be
484: real and symmetric (this is the reason for
485: the somewhat unconventional phase factor
486: in Eq.~(\ref{eq:vpe}); in order to mantain symmetry for
487: both the $v^{\rm PC}$- and $v^{\rm PV}$-matrices, and
488: hence the $S$-matrix, the states used here differ by a factor ${\rm i}^L$ from
489: those usually used in nucleon-nucleon scattering analyses).
490: The asymptotic behavior of the $w(r)$'s is given
491: in Eq.~(\ref{eq:asy}).
492:
493: The Pauli principle requires that there be a single channel
494: when $J$ is odd, and three coupled channels when $J$ is even,
495: with the exception of $J$=0 in which case there are only two
496: coupled channels, $^1$S$_0$ and $^3$P$_0$. The situation
497: is summarized in Table~\ref{tb:chan}.
498: Again because of the invariance under time-inversion
499: transformations of $v^{\rm PC}+v^{\rm PV}$,
500: the $S$-matrix is symmetric (apart from also
501: being unitary), and can therefore be written, for the coupled
502: channels having $J$ even, as~\cite{Goldberger64}
503:
504: \begin{equation}
505: S^J = U^{\rm T} \, S^J_{\rm D} \, U \>\>,
506: \label{eq:sdiag}
507: \end{equation}
508: where $U$ is a real orthogonal matrix, and $S^J_D$ is a diagonal
509: matrix of the form
510:
511: \begin{equation}
512: S^J_{\rm D;\alpha^\prime,\alpha} = \delta_{\alpha^\prime,\alpha}
513: {\rm e}^{2 {\rm i} \delta^J_\alpha} \>\>.
514: \end{equation}
515: Here $\delta^J_\alpha$ is the (real) phase-shift in
516: channel $\alpha$, which is function of the energy $p=\sqrt{2\mu\, E}$.
517: The mixing matrix $U$ can be written as
518:
519: \begin{eqnarray}
520: U &=& U^{(12)} \quad \qquad \qquad J=0 \>\>, \\
521: &=& \prod_{1 \leq i < j \leq 3} U^{(ij)}
522: \qquad J \geq 2\>\>{\rm with}\>\> J \>\>{\rm even}\>\>,
523: \label{eq:uma}
524: \end{eqnarray}
525: where $U^{(ij)}$ is the $2 \times 2$ or $3 \times 3$
526: orthogonal matrix, that includes the coupling between
527: channels $i$ and $j$ only, for example
528:
529: \noindent
530: \centerline{
531: $
532: U^{(13)}= \left[ \begin{array}{ccc}
533: {\rm cos}\,\epsilon^J_{13} & 0 & {\rm sin}\,\epsilon^J_{13} \\
534: 0 & 1 & 0 \\
535: -{\rm sin}\,\epsilon^J_{13} & 0 & {\rm cos}\,\epsilon^J_{13}
536: \end{array} \right]
537: \simeq 1 +\epsilon^J_{13} \left[ \begin{array}{ccc}
538: 0 & 0 & 1 \\
539: 0 & 0 & 0 \\
540: -1 & 0 & 0 \\
541: \end{array} \right ] \>\>.
542: $ }
543: \vskip 0.7cm
544:
545: \noindent
546: Thus, for $J$=0 there are two phase-shifts and a mixing
547: angle, while for $J$ even $\geq 2$ there are
548: three phases and three mixing angles. Of course, since
549: $| v^{\rm PV} | \ll | v^{\rm PC} |$, the mixing angles $\epsilon^J_{ij}$
550: induced by $v^{\rm PV}$ are $\ll 1$, a fact already exploited in the
551: last expression above for $U$. Given the channel ordering in
552: Table~\ref{tb:chan}, Table~\ref{tb:mixing} specifies
553: which of the channel mixings are induced by $v^{\rm PC}$
554: and which by $v^{\rm PV}$.
555:
556: The reality of the potential matrix elements
557: $v^J_{\alpha^\prime,\alpha}(r)$ makes
558: it possible to construct real solutions of the
559: Schr\"odinger equation~(\ref{eq:cschr}).
560: The problem is reduced to determining the
561: relation between these solutions and the complex $w(r)$'s functions.
562: Using Eq.~(\ref{eq:sdiag}) and $U^{\rm T} U=1$,
563: the $w(r)$'s can be expressed in the asymptotic region as
564:
565: \begin{eqnarray}
566: \frac{w^J_{\alpha^\prime,\alpha}}{r} &\simeq&
567: \sum_\beta (U^{\rm T})_{\alpha^\prime \beta} \, {\rm e}^{ {\rm i}\delta^J_\beta}
568: \frac{ h^{(2)}_{\alpha^\prime} {\rm e}^{-{\rm i}\delta^J_\beta}
569: + h^{(1)}_{\alpha^\prime} {\rm e}^{ {\rm i}\delta^J_\beta} }{2}
570: U_{\beta \alpha} \nonumber \\
571: &=& \sum_\beta (U^{\rm T})_{\alpha^\prime \beta} \,
572: {\rm e}^{ {\rm i}\delta^J_\beta} \,
573: \frac{ {\rm sin}[pr-L^\prime \, \pi/2
574: -\eta {\rm lg}(2pr)+\sigma_{L^\prime}+\delta^J_\beta] }{pr}
575: \, U_{\beta \alpha} \>\>,
576: \end{eqnarray}
577: where the $\epsilon_{L^\prime S^\prime}$ has been
578: dropped for simplicity. The expression above is real apart from
579: the ${\rm exp}({\rm i}\delta^J_\beta)$. To eliminate
580: this factor, the following linear combinations of the
581: $w(r)$'s are introduced
582:
583: \begin{eqnarray}
584: \frac{u^J_{\alpha^\prime,\alpha}}{r} &\equiv& \sum_\beta
585: {\rm e}^{-{\rm i}\delta^J_\beta}
586: \frac{w^J_{\alpha^\prime,\beta}}{r}(U^{\rm T})_{\beta \alpha} \nonumber \\
587: &\simeq&(U^{\rm T})_{\alpha^\prime \alpha}
588: \frac{{\rm cos}\, \delta^J_\alpha \, F_{L^\prime}(\eta;pr)
589: +{\rm sin}\, \delta^J_\alpha \, G_{L^\prime}(\eta;pr)}{pr} \>\>,
590: \label{eq:uasy}
591: \end{eqnarray}
592: and the $u(r)$'s are then the sought real solutions of Eq.~(\ref{eq:cschr}).
593:
594: The asymptotic behavior of the $u(r)$'s can now be read off
595: from Eq.~(\ref{eq:uasy}) once the $U$-matrices above have been
596: constructed. The latter can be written as, up
597: to linear terms in the \lq\lq small\rq\rq mixing angles induced
598: by $v^{\rm PV}$,
599:
600: \noindent
601: \centerline{
602: $
603: U= \left[ \begin{array}{cc}
604: 1 & \epsilon^0_{12} \\
605: -\epsilon^0_{12} & 1
606: \end{array} \right] \quad J=0 \>\>,
607: $}
608: \vskip 0.7cm
609: \centerline{
610: $
611: U= \left[ \begin{array}{ccc}
612: {\rm cos}\epsilon^J_{12}
613: & {\rm sin}\epsilon^J_{12}
614: & \epsilon^J_{13}{\rm cos}\epsilon^J_{12}+\epsilon^J_{23}{\rm sin}\epsilon^J_{12} \\
615: -{\rm sin}\epsilon^J_{12}
616: & {\rm cos}\epsilon^J_{12}
617: & -\epsilon^J_{13}{\rm sin}\epsilon^J_{12}+\epsilon^J_{23}{\rm cos}\epsilon^J_{12} \\
618: -\epsilon^J_{13} & -\epsilon^J_{23} & 1 \\
619: \end{array} \right ] \quad J \geq 2 ,\>\> J\>\>{\rm even} \>\>.
620: $}
621: \vskip 0.7cm
622:
623: Inverting the first line of Eq.~(\ref{eq:uasy}),
624:
625: \begin{equation}
626: \frac{w^J_{\alpha^\prime,\alpha}}{r} = \sum_\beta
627: {\rm e}^{ {\rm i}\delta^J_\beta}
628: \frac{u^J_{\alpha^\prime,\beta}}{r}\,U_{\beta \alpha} \>\>,
629: \end{equation}
630: and inserting the resulting expressions into Eq.~(\ref{eq:cschr})
631: leads to the (in general, coupled-channel) Schr\"odinger equations
632: satisfied by the (real) functions $u(r)$. They are identical to
633: those of Eq.~(\ref{eq:cschr}), but for the $w(r)$'s being replaced
634: by the $u(r)$'s. These equations are then solved by standard numerical
635: techniques. Note that: i) $v^J_{\alpha^\prime,\alpha} =
636: v^{J,\, {\rm PC}}_{\alpha^\prime,\alpha}$, since the diagonal
637: matrix elements of $v^{\rm PV}$ vanish because of parity selection rules;
638: ii) in the coupled equations with $J$ even, terms of the type
639: $r \,v^{J,\, {\rm PV}}_{\alpha^\prime,\beta}(r) u^J_{\beta,\alpha}(r)/r$
640: involving the product of a parity-violating potential matrix element
641: with a $v^{\rm PV}$-induced wave function are neglected.
642:
643: \subsection{Amplitudes, Cross Sections, and the Parity-Violating Asymmetry}
644:
645: The amplitude for $p$$p$ elastic scattering from an initial
646: state with spin projections $m_1$, $m_2$ to a final state
647: with spin projections $m^\prime_1$, $m^\prime_2$ is given
648: by
649:
650: \begin{eqnarray}
651: \langle m_1^\prime m_2^\prime \mid \overline{M} \mid m_1 m_2\rangle&=&
652: \sum_{S^\prime M_S^\prime,SM_S}
653: \langle \frac{1}{2} m_1^\prime ,
654: \frac{1}{2} m_2^\prime \mid S^\prime M_S^\prime\rangle
655: \langle \frac{1}{2} m_1 ,\frac{1}{2} m_2 \mid S M_S\rangle \nonumber \\
656: &&\overline{M}_{S^\prime M_S^\prime, SM_S}(E,\theta) \>\>,
657: \end{eqnarray}
658: where the amplitude $\overline{M}$ is related to the
659: $\overline{T}$-matrix defined in Eq.~(\ref{eq:tmas}) via
660:
661: \begin{equation}
662: \overline{M}_{S^\prime M_S^\prime, SM_S}(E,\theta) =-\frac{\mu}{2\pi}
663: \overline{T}({\bf p}^\prime,S^\prime M_S^\prime;p\hat{\bf z},SM_S) \>\>.
664: \end{equation}
665: Note that the direction of the initial momentum
666: ${\bf p}$ has been taken to define the spin quantization axis
667: (the $z$-axis), $\theta$ is the angle between
668: $\hat {\bf p}$ and $\hat{\bf p}^\prime$, the direction
669: of the final momentum, and the energy $E=p^2/(2\mu)$
670: ($= p^{\prime 2}/(2\mu)$). The amplitude $\overline{M}$
671: is split into two terms, $\overline{M}=M+M^{\rm C}$,
672: as in Eq.~(\ref{eq:tmas}). Using the expansion of
673: the $T$-matrix, Eq.~(\ref{eq:tmae}) with $v^J_{L^\prime S^\prime,LS}$
674: replaced by $T^J_{L^\prime S^\prime,LS}$, and the relation
675: between the $S$- and $T$-matrices, Eq.~(\ref{eq:sma}), the amplitude induced
676: by $v^{\rm PC}+v^{\rm PV}$ can be expressed as
677:
678: \begin{eqnarray}
679: M_{S^\prime M_S^\prime, SM_S}(E,\theta)&=& \sqrt{4\pi} \sum_{JLL^\prime}
680: \sqrt{2L+1}\, \epsilon_{L^\prime S^\prime}\, \epsilon_{LS} \,
681: \langle L^\prime (M_S-M^\prime_S), S^\prime M^\prime_S\mid J M_S\rangle \nonumber \\
682: &&\langle L 0, SM_S\mid J M_S\rangle Y_{L^\prime (M_S-M_S^\prime)}(\theta)\,
683: {\rm e}^{{\rm i}\sigma_{L^\prime}}\,
684: \frac{S^J_{L^\prime S^\prime,LS}(p) - \delta_{L^\prime,L}
685: \,\delta_{S^\prime,S}}{{\rm i} p}
686: \,{\rm e}^{{\rm i}\sigma_L} \>\>,
687: \label{eq:am}
688: \end{eqnarray}
689: while the partial-wave expansion of the amplitude
690: associated with the Coulomb potential reads~\cite{Goldberger64}:
691:
692: \begin{equation}
693: M^{\rm C}_{S^\prime M_S^\prime, SM_S}(E,\theta)=\delta_{S^\prime,S} \,
694: \delta_{M^\prime_S,M_S} \,\sqrt{4\pi} \sum_{L} \sqrt{2L+1}\,\epsilon_{LS}\,
695: Y_{L0}(\theta) \frac{ {\rm e}^{2 {\rm i}\sigma_L}-1}{{\rm i} p}\>\>.
696: \label{eq:amc}
697: \end{equation}
698:
699: The differential cross section for scattering of a proton with initial
700: polarization $m_1$ is then given by
701:
702: \begin{equation}
703: \overline{\sigma}_{m_1}(E,\theta)=\frac{1}{2} \sum_{m_2} \sum_{m_1^\prime m_2^\prime}
704: \mid \langle m_1^\prime m_2^\prime \mid \overline{M} \mid m_1 m_2\rangle\mid^2 \>\>,
705: \end{equation}
706: and the longitudinal asymmetry is defined as
707:
708: \begin{equation}
709: \overline{A}(E,\theta)=\frac{\overline{\sigma}_{+}(E,\theta)
710: - \overline{\sigma}_{-}(E,\theta)}
711: {\overline{\sigma}_{+}(E,\theta) + \overline{\sigma}_{-}(E,\theta) } \>\>,
712: \end{equation}
713: where $\pm$ denote the initial polarizations $\pm 1/2$. Carrying out
714: the spin sums leads to the following expression for the asymmetry:
715:
716: \begin{equation}
717: \overline{A}(E,\theta)=\frac{
718: \sum_{S^\prime M^\prime_S} \left[
719: \overline{M}_{S^\prime M_S^\prime, 00}(E,\theta)
720: \overline{M}^*_{S^\prime M_S^\prime, 10}(E,\theta) + {\rm c.c.}\right] }
721: {\sum_{S^\prime M^\prime_S} \sum_{S M_S}
722: \mid \overline{M}_{S^\prime M_S^\prime, S M_S}(E,\theta) \mid^2 }\>\>,
723: \label{eq:angc}
724: \end{equation}
725: from which it is clear that the numerator would vanish
726: in the absence of parity-violating interactions, since
727: $v^{\rm PC}+v^{\rm C}$, in contrast to $v^{\rm PV}$,
728: cannot change the total spin $S$ of the $p$$p$ pair.
729:
730: Parity-violating scattering experiments typically measure
731: the asymmetry weighted over a range $[\theta_1,\theta_2]$
732: of scattering angles,
733:
734: \begin{equation}
735: \langle\, \overline{A}(E)\, \rangle =
736: \frac{\int_{\theta_1 \leq \theta \leq \theta_2}
737: {\rm d}\Omega\, \overline{\sigma}(E,\theta) \, \overline{A}(E,\theta) }
738: {\int_{\theta_1 \leq \theta \leq \theta_2}{\rm d}\Omega \,
739: \overline{\sigma}(E,\theta)} \>\>,
740: \end{equation}
741: where $\overline{\sigma}=
742: \left( \overline{\sigma}_{+}+\overline{\sigma}_{-} \right)/2$
743: is the spin-averaged differential cross section.
744: In contrast, transmission experiments measure the
745: transmission of a polarized proton beam through a target. A cross
746: section is then inferred from the transmission measurement. Beam
747: particles elastically scattered by angles greater than some
748: small critical angle $\theta_0$ are removed from the beam, thus
749: reducing the observed transmission and adding to the inferred
750: cross section. Beam particles scattered at angles smaller than
751: $\theta_0$ are not distinguished from the beam and do not
752: contribute to the cross section. To derive an expression
753: for the asymmetry in this case, one needs to carefully
754: consider the Coulomb contribution to the cross section--a divergent
755: quantity in the limit $\theta_0 \rightarrow 0$. To this end,
756: following Ref.~\cite{Holdeman65}, one first defines the
757: differential cross sections
758:
759: \begin{eqnarray}
760: \sigma_{S^\prime M^\prime_S,SM_S}(E,\theta) &\equiv&
761: \mid \overline{M}_{S^\prime M^\prime_S,SM_S}(E,\theta) \mid^2
762: -\mid M^{\rm C}_{S^\prime M^\prime_S,SM_S}(E,\theta) \mid^2 \nonumber \\
763: &=&\mid M_{S^\prime M^\prime_S,SM_S}(E,\theta) \mid^2+
764: 2\,\Re\Big[M_{S^\prime M^\prime_S,SM_S}(E,\theta)
765: M^{{\rm C}\, *}_{S^\prime M^\prime_S,SM_S}(E,\theta)\Big]\>\>,
766: \end{eqnarray}
767: and
768:
769: \begin{equation}
770: \sigma^{\rm C}_{S^\prime M^\prime_S,SM_S}(E,\theta)=
771: \mid M^{\rm C}_{S^\prime M^\prime_S,SM_S}(E,\theta) \mid^2 \>\>,
772: \end{equation}
773: and hence
774:
775: \begin{equation}
776: \overline{\sigma}_{S^\prime M^\prime_S,SM_S}(E,\theta)=
777: \sigma_{S^\prime M^\prime_S,SM_S}(E,\theta)+
778: \sigma^{\rm C}_{S^\prime M^\prime_S,SM_S}(E,\theta) \>\>.
779: \end{equation}
780: In transmission experiments, the quantity of interest is
781:
782: \begin{eqnarray}
783: \overline{\sigma}_{SM_S,>}(E)&\equiv&
784: 2\pi \int_{\theta_0}^{\pi/2}{\rm d}\theta \,{\rm sin}\theta
785: \sum_{S^\prime M^\prime_S}
786: \sigma_{S^\prime M^\prime_S,SM_S}(E,\theta) \nonumber \\
787: &=&\sigma_{SM_S,>}(E)+\sigma^{\rm C}_{SM_S,>}(E) \>\>,
788: \label{eq:s_tra}
789: \end{eqnarray}
790: where $\sigma^{\rm C}_{SM_S,>}(E)$ is explicitly given by
791:
792: \begin{equation}
793: \sigma^{\rm C}_{SM_S,>}(E)=\pi \left(\frac{\eta}{p}\right)^2
794: \Bigg[ \frac{1}{{\rm sin}^2\theta_0/2} - \frac{1}{{\rm cos}^2\theta_0/2}
795: -\frac{(-)^S}{\eta} {\rm sin}
796: \left[ 2\eta\, {\rm lg}\, {\rm tg}(\theta_0/2)\right]\Bigg] \>\>.
797: \label{eq:scc}
798: \end{equation}
799: To evaluate $\sigma_{SM_S,>}(E)$, one writes, following Ref.~\cite{Holdeman65}:
800:
801: \begin{eqnarray}
802: \sigma_{SM_S,>}(E)&=&\sigma_{SM_S}(E)-
803: 2\pi\int_{\epsilon \rightarrow 0}^{\theta_0}{\rm d}\theta \,
804: {\rm sin}\theta \, \sum_{S^\prime M^\prime_S}
805: \Bigg[ \mid M_{S^\prime M^\prime_S,SM_S}(E,\theta) \mid^2 \nonumber \\
806: &+&2\,\Re\Big[M_{S^\prime M^\prime_S,SM_S}(E,\theta)
807: M^{{\rm C}\, *}_{S^\prime M^\prime_S,SM_S}(E,\theta)\Big] \Bigg] \>\>.
808: \label{eq:sbc}
809: \end{eqnarray}
810: Application of the optical theorem to the total cross sections
811: $\overline{\sigma}$ and $\sigma^{\rm C}$ allows
812: one to deduce
813:
814: \begin{equation}
815: \sigma_{SM_S}(E)=\overline{\sigma}_{SM_S}(E)-\sigma^{\rm C}_{SM_S}(E)
816: =\frac{4\pi}{p}\Im \Big[ M_{SM_S,SM_S}(E,0) \Big] \>\>,
817: \end{equation}
818: and the determination of the cross section $\sigma_{SM_S,>}(E)$ is
819: reduced to evaluating the integral on the right-hand-side of Eq.~(\ref{eq:sbc}).
820: For sufficiently small $\theta_0$ and by appropriately taking the limit
821: $\epsilon \rightarrow 0$ in the integral of the interference
822: term $MM^{{\rm C}\, *}$, which essentially entails taking the limit
823: term by term in the partial-wave expansion of $M^{\rm C}$, one finds
824:
825: \begin{equation}
826: \sigma_{SM_S,>}(E)=\frac{4\pi}{p} \Im \Bigg[ M_{SM_S,SM_S}(E,0)
827: {\rm e}^{2{\rm i}[ \eta \, {\rm lg}\, {\rm sin}(\theta_0/2)-\sigma_0]}\Bigg] \>\>,
828: \label{eq:sc}
829: \end{equation}
830: neglecting terms of order $\theta_0^2$ and higher. Using
831: Eqs.~(\ref{eq:scc}) and~(\ref{eq:sc}), the longitudinal asymmetry
832: measured in transmission experiments is obtained as
833:
834: \begin{equation}
835: \overline{A}_{>}(E)=
836: \frac{\Im \Big[
837: {\rm e}^{ {\rm i}\phi}\,\left[ M_{10,00}(E,0) + M_{00,10}(E,0) \right] \Big]}
838: {\Im \Big[{\rm e}^{ {\rm i}\phi} \sum_{SM_S} M_{SM_S,SM_S}(E,0)\Big]
839: +(p/4\pi) \sum_{SM_S} \sigma^{\rm C}_{SM_S,>}(E)} \>\>,
840: \label{eq:asy_t}
841: \end{equation}
842: with ${\rm exp}({\rm i}\phi)\equiv
843: {\rm exp}\Big[ 2{\rm i}[ \eta \, {\rm lg}\, {\rm sin}(\theta_0/2)-\sigma_0]\Big]$.
844: %
845: \subsection{Momentum-Space Formulation}
846:
847: In order to consider the (parity-conserving) Bonn
848: potential~\cite{Machleidt01}, it is necessary to develop
849: techniques to treat the $p$$p$ scattering problem in
850: momentum space. A method first proposed in Ref.~\cite{Vincent74}
851: and most recently applied in Ref.~\cite{Machleidt01} is used
852: here. It consists in separating the potential into short-
853: and long-range parts $\overline{v}_{\rm S}$ and
854: $\overline{v}_{\rm L}$, respectively,
855:
856: \begin{equation}
857: \overline{v}=\overline{v}_{\rm S} + \overline{v}_{\rm L} \>\>,
858: \end{equation}
859: where
860:
861: \begin{eqnarray}
862: \overline{v}_{\rm S}&=&\left[ v^{\rm PC}+v^{\rm PV}
863: +v^{\rm C}\right]\theta(R-r) \>\>, \\
864: \overline{v}_{\rm L}&=&v^{\rm C}\, \theta(r-R) \>\>,
865: \end{eqnarray}
866: and $\theta(x)$ is the Heaviside step function,
867: $\theta(x)$=1 if $x > 0$, =0 otherwise. The radius $R$ is
868: chosen large enough, so that $v^{\rm PC}+v^{\rm PV}$ vanishes
869: for $r > R$ (in the present work, $R=20$ fm).
870:
871: Since $\overline{v}_{\rm S}$ is of finite range,
872: standard momentum-space techniques can now be used to solve for the
873: $K_{\rm S}$-matrix in the $J$ channel(s):
874:
875: \begin{eqnarray}
876: K^J_{{\rm S};\alpha^\prime,\alpha}(p^\prime;p)&=&
877: \overline{v}^J_{{\rm S};\alpha^\prime,\alpha}(p^\prime;p)\nonumber \\
878: &+&\frac{2}{\pi} \int_0^\infty {\rm d}kk^2 \sum_\beta
879: \overline{v}^J_{{\rm S};\alpha^\prime,\beta}(p^\prime;k)
880: \frac{\cal P}{p^2/(2\mu)-k^2/(2\mu)}K^J_{{\rm S};\beta,\alpha}(k;p) \>\>,
881: \label{eq:kma}
882: \end{eqnarray}
883: where ${\cal P}$ denotes a principal-value integration,
884: and the momentum-space matrix elements of the potential
885: $\overline{v}_{\rm S}$ are defined as in Eq.~(\ref{eq:vpe}),
886: but for the replacements $v\rightarrow \overline{v}_{\rm S}$
887: and $F_L(\eta;x)/x \rightarrow j_L(x)$. Note that performing
888: the Bessel transforms of a Coulomb potential
889: truncated at $r=R$ poses no numerical problem. The integral
890: equations~(\ref{eq:kma}) are discretized, and the resulting
891: systems of linear equations are solved by direct numerical
892: inversion. The principal-value integration is eliminated by
893: a standard subtraction technique~\cite{Gloeckle83}.
894:
895: The asymptotic wave functions associated with $\overline{v}_{\rm S}$
896: have the form:
897:
898: \begin{equation}
899: \frac{\overline{w}^J_{{\rm S};\alpha^\prime,\alpha}(r;p)}{r}
900: \simeq \frac{a_\alpha}{2}
901: \Big[ \delta_{\alpha,\alpha^\prime} \hat{h}^{(2)}_{L^\prime}(pr)
902: +\hat{h}^{(1)}_{L^\prime}(pr)
903: \overline{S}^J_{{\rm S};\alpha^\prime , \alpha}(p) \Big] \>\>,
904: \label{eq:awc}
905: \end{equation}
906: where
907:
908: \begin{equation}
909: \hat{h}^{(1,2)}_L(\rho)\equiv j_L(\rho) \pm {\rm i}\, n_L(\rho) \>\>,
910: \end{equation}
911: $j_L$ and $n_L$ being the regular and irregular
912: spherical Bessel functions, respectively, and the constants $a_\alpha$ can
913: only depend upon the entrance channel $\alpha$, see the Schr\"odinger
914: equations~(\ref{eq:cschr}). These wave functions
915: should match smoothly, at $r$=$R$, those associated with the full potential
916: $\overline{v}_{\rm S}+\overline{v}_{\rm L}$, which behave
917: asymptotically as in Eq.~(\ref{eq:asy}). Carrying out the matching
918: for the functions and their first
919: derivatives leads to a relation between the $S$-matrices
920: $S_{{\rm S};\alpha^\prime,\alpha}$ and $S_{\alpha^\prime,\alpha}$,
921: corresponding to $\overline{v}_{\rm S}$ and
922: $\overline{v}_{\rm S}+\overline{v}_{\rm L}$, respectively.
923: In terms of $K$-matrices, related to the corresponding $S$-matrices
924: via
925:
926: \begin{equation}
927: S^J(p)=\left[ 1+2{\rm i} \, \mu p\, K^J(p;p) \right]^{-1}
928: \left[ 1-2{\rm i} \, \mu p\, K^J(p;p) \right] \>\>,
929: \label{eq:skma}
930: \end{equation}
931: and similarly $S_{\rm S}$ and $K_{\rm S}$,
932: this relation reads in matrix notation~\cite{Machleidt01a}:
933:
934: \begin{eqnarray}
935: 2 \mu p\, K^J&=&
936: \Bigg[ {\rm G} -\left[ {\rm J}+2 \mu p \, {\rm N} \cdot K^J_{\rm S}\right]
937: \left[{\rm J}^\prime+2 \mu p \,{\rm N}^\prime \cdot K^J_{\rm S}\right]^{-1}
938: \cdot {\rm G}^\prime \Bigg]^{-1} \nonumber \\
939: &\times& \Bigg[ \left[{\rm J}+2 \mu p \, {\rm N}\cdot K^J_{\rm S}\right]
940: \left[ {\rm J}^\prime+2 \mu p \, {\rm N}^\prime \cdot
941: K^J_{\rm S}\right]^{-1}\cdot {\rm F}^\prime-{\rm F} \Bigg] \>\>,
942: \end{eqnarray}
943: where the dependence upon $p$ is understood, and
944: the diagonal matrices ${\rm X}$ and ${\rm X}^\prime$ have been defined as
945:
946: \begin{eqnarray}
947: {\rm X}_{\alpha^\prime,\alpha} &\equiv&
948: \delta_{\alpha^\prime,\alpha}\, X_\alpha(R) \>\>, \\
949: {\rm X}^\prime_{\alpha^\prime,\alpha} &\equiv&
950: \delta_{\alpha^\prime,\alpha}\, \Bigg[ \frac{ {\rm d}X_\alpha(r)}
951: { {\rm d}r}\Bigg]_{r=R} \>\>,
952: \end{eqnarray}
953: with the functions
954:
955: \begin{equation}
956: X_\alpha(R)=j_L(pR), \>\> \frac{F_L(\eta;pR)}{pR},\>\> n_L(pR),\>\>
957: {\rm and}\> \frac{G_L(\eta;pR)}{pR}
958: \end{equation}
959: when ${\rm X}$=${\rm J}$, ${\rm F}$, ${\rm N}$, and ${\rm G}$, respectively.
960: Once the $K$-matrices in the various channels have been determined,
961: the corresponding $S$-matrices are obtained from Eq.~(\ref{eq:skma}), from
962: which the amplitude $M_{S^\prime M^\prime_S,SM_S}(E,\theta)$~(\ref{eq:am})
963: is constructed.
964: %
965: \subsection{Matrix elements of $v^{\rm PV}$ in channel $J$}
966:
967: To evaluate the radial functions $v^{J,\, {\rm PV}}_{\alpha^\prime,\alpha}(r)$
968: of the PV potential in Eq.~(\ref{eq:vpe})--those associated with
969: the PC potential are well known--one needs the matrix
970: elements of $({\bbox \sigma}_1 \times{\bbox \sigma}_2)\cdot \hat{\bf r}$
971: and $({\bbox \sigma}_1- {\bbox \sigma}_2) \cdot {\bf p}$ between spin-angle
972: functions. Using the notation
973:
974: \begin{equation}
975: \langle J;L^\prime, S^\prime \mid O \mid J; L,S\rangle \equiv
976: \int{\rm d}\Omega\, {\cal Y}_{L^\prime S^\prime J}^{M_J \dagger}
977: \, O({\bf r}) \, {\cal Y}_{LSJ}^{M_J} \>\>,
978: \end{equation}
979: and writing
980:
981: \begin{equation}
982: ({\bbox \sigma}_1- {\bbox \sigma}_2) \cdot {\bf p}=-{\rm i} \,
983: ({\bbox \sigma}_1- {\bbox \sigma}_2) \cdot \left[\hat{\bf r}
984: \frac{\vec{\partial}}{\partial r} + \frac{1}{r} \frac{\vec{\partial}}
985: {\partial {\bbox \Omega}} \right] \>\>,
986: \end{equation}
987: where the $\vec{\partial}$ symbol indicates that the partial
988: derivatives must act to the right, one finds that the
989: non-vanishing matrix elements are:
990:
991: \begin{eqnarray}
992: \langle J;J,0 \mid
993: ({\bbox \sigma}_1 \times{\bbox \sigma}_2)\cdot \hat{\bf r} \mid J;J\mp 1,1\rangle
994: &=&\pm {\rm i}\, \sqrt{\frac{2\, J+1\mp 1}{J +1/2}} \>\>, \label{eq:vpv1}\\
995: \langle J;J,0 \mid
996: ({\bbox \sigma}_1 - {\bbox \sigma}_2)\cdot \hat{\bf r} \mid J;J\mp 1,1\rangle
997: &=&\pm \sqrt{\frac{2\, J+1\mp 1}{J +1/2}} \>\>,
998: \label{eq:vpv2}
999: \end{eqnarray}
1000:
1001: \begin{eqnarray}
1002: \langle J;J,0 \mid
1003: ({\bbox \sigma}_1 - {\bbox \sigma}_2)\cdot \frac{\vec{\partial}}
1004: {\partial {\bbox \Omega}} \mid J;J\mp 1,1\rangle &=&
1005: -\frac{2\, J+1 \mp 3}{2} \sqrt{\frac{2\, J+1\mp 1}{J +1/2}} \>\>, \label{eq:vpv3} \\
1006: \langle J;J\mp 1,1 \mid
1007: ({\bbox \sigma}_1 - {\bbox \sigma}_2)\cdot \frac{\vec{\partial}}
1008: {\partial {\bbox \Omega}} \mid J;J,0\rangle &=&
1009: \frac{2\, J+1 \pm 1}{2} \sqrt{\frac{2\, J+1\mp 1}{J +1/2}} \>\>.
1010: \label{eq:vpv4}
1011: \end{eqnarray}
1012: Note that the operators in Eqs.~(\ref{eq:vpv1})--(\ref{eq:vpv2}) are Hermitian,
1013: while those in Eqs.~(\ref{eq:vpv3})--(\ref{eq:vpv4}) are not.
1014: The complete Hamiltonian is, of course, Hermitian.
1015:
1016:
1017:
1018: %
1019: %
1020: \section{Results and Discussion}
1021: \label{sec:res}
1022:
1023: In this section we present results for the longitudinal asymmetry in
1024: the lab energy range 0--350 MeV. The calculations use any of the
1025: modern strong-interaction potentials, either Argonne $v_{18}$ (AV18)~\cite{Wiringa95},
1026: or CD-Bonn (BONN)~\cite{Machleidt01}, or Nijmegen-I (NIJ-I)~\cite{Stoks94},
1027: in combination with the DDH weak-interaction potential parameterized in terms of
1028: $\rho$- and $\omega$-meson exchanges~\cite{Desplanques80}. The values
1029: for the $\rho$- and $\omega$-meson coupling constants and cutoff parameters
1030: are listed in Table~\ref{tb:gs}. The strong-interaction coupling
1031: constants and cutoff parameters are taken from the BONN potential, while
1032: the weak-interaction coupling constants $h^{pp}_\rho$ and
1033: $h^{pp}_\omega$ have been determined by an AV18-based fit to the
1034: observed asymmetry. In Table~\ref{tb:gs} we also list
1035: the $h^{pp}_\rho$ and $h^{pp}_\omega$ values corresponding to the \lq\lq best\rq\rq
1036: estimates for the $h_{\rho_i}$ and $h_{\omega_i}$ suggested in Ref.~\cite{Desplanques80},
1037: column labeled DDH-orig.
1038:
1039: The data points for the longitudinal asymmetry at 13.6 MeV, 45 MeV,
1040: and 221 MeV are those reported in Ref.~\cite{Berdoz01,vanOers01}, and their values
1041: are $(-0.97\pm 0.20)\times 10^{-7}$, $(-1.53\pm 0.21)\times 10^{-7}$, and
1042: $(+0.84\pm 0.34)\times 10^{-7}$, respectively. The first point at 13.6
1043: MeV has been obtained~\cite{vanOers01} by taking the weighted mean--and accounting
1044: for the square-root energy dependence--of the latest result from the the Bonn
1045: experiment at 13.6 MeV, as reported by Eversheim (Ref.~[14]
1046: in Ref.~\cite{Berdoz01}), and the 15 MeV result from Ref.~\cite{Nagle79}.
1047: The point at 45 MeV has also been obtained~\cite{vanOers01} by combining
1048: results from measurements at 45 MeV~\cite{Kistryn87}, 46 MeV, and 47 MeV
1049: (these last two both from Ref.~\cite{vanOers01}). The last point at
1050: 221 MeV is that reported in Ref.~\cite{Berdoz01}. Finally, the errors
1051: include both statistical and systematic errors added in quadrature.
1052:
1053: The total longitudinal asymmetry, shown in Fig.~\ref{fig:asy}
1054: for a number of combinations of strong- and weak-interaction
1055: potentials, is defined as
1056:
1057: \begin{equation}
1058: A(E)= \frac{\Im \Big[ M_{10,00}(E,0) + M_{00,10}(E,0)\Big]}
1059: {\Im \Big[\sum_{SM_S} M_{SM_S,SM_S}(E,0)\Big]} \>\>,
1060: \end{equation}
1061: where the amplitudes $M_{S^\prime M_S^\prime ,SM_S}(E,\theta)$
1062: are those given in Eq.~(\ref{eq:am}). The expression above
1063: for $A(E)$ ignores the contribution of the Coulomb amplitude,
1064: Eq.~(\ref{eq:amc}), divergent in the limit $\theta$=0, and for this reason
1065: $A(E)$ will be referred to as the \lq\lq nuclear\rq\rq asymmetry.
1066: Of course, one should note that Coulomb potential effects
1067: enter into $A(E)$ explicitly through the Coulomb phase shifts, present
1068: in the partial-wave expansion for $M_{S^\prime M_S^\prime,SM_S}(E,\theta)$,
1069: and implicitly through the wave functions, from which the $S$-matrix
1070: elements are calculated. The effect of including explicitly
1071: the amplitude induced by the Coulomb potential is discussed below.
1072:
1073: The calculated nuclear asymmetries in Fig.~\ref{fig:asy} were
1074: obtained by retaining in the partial-wave expansion for
1075: $M_{S^\prime M_S^\prime ,SM_S}(E,\theta)$ all channels with
1076: $J$ up to $J_{\rm max}=8$. The curves labeled AV18, BONN, and
1077: NIJ-I all use the DDH potential with the coupling constants $h^{pp}_\rho$
1078: and $h^{pp}_\omega$ determined by a rough fit to data (the AV18 is used
1079: in the fitting procedure). There is very little sensitivity
1080: to the input strong-interaction potential, certainly
1081: much less than one would infer from Fig.~1 of Ref.~\cite{Driscoll89}.
1082: This is undoubtedly a consequence of the more extended $p$$p$ and
1083: $p$$n$ scattering data-base to which present potentials are fitted, as
1084: well as of the much higher accuracy achieved in these fits.
1085: An analysis of the extracted $h^{pp}_\rho$ and $h^{pp}_\omega$
1086: coupling constants and their errors is presented later in this section.
1087:
1088: We also show in Fig.~\ref{fig:asy} the AV18 results which
1089: correspond to a DDH potential using the \lq\lq best\rq\rq estimates for the
1090: $h^{pp}_\rho$ and $h^{pp}_\omega$ coupling constants~\cite{Desplanques80}
1091: (values in column DDH-orig in Table~\ref{tb:gs}), with the remaining $\rho$-
1092: and $\omega$-meson strong-interaction coupling constants and cutoff
1093: parameters as given in Table~\ref{tb:gs}. A number of comments are now
1094: in order. The data point at 221 MeV essentially determines the value
1095: of $h^{pp}_\rho$. As pointed out by Simonius~\cite{Simonius88} (see also
1096: below), the dominant contributions to the total asymmetry in the energy
1097: range under consideration here are those associated with the $^1$S$_0$-$^3$P$_0$
1098: and $^3$P$_2$-$^1$D$_2$ partial waves. At energies close to 225 MeV
1099: the $^1$S$_0$-$^3$P$_0$ contribution, which can easily be shown to be
1100: proportional to ${\rm cos}\left[\delta(E;^1{\rm S}_0) + \sigma_1(E) +\sigma_0(E) \right]
1101: -{\rm cos}\left[\delta(E;^3{\rm P}_0) + \sigma_1(E) +\sigma_0(E) \right]$ using
1102: Eq.~(\ref{eq:am}) (here $\delta(E;^1{\rm S}_0)$ and $\delta(E;^3{\rm P}_0)$
1103: are the strong-interaction phases), vanishes. As a result, the total asymmetry
1104: in this energy region is almost entirely due to the $^3$P$_2$-$^1$D$_2$ contribution,
1105: which is known~\cite{Simonius88} to be approximately proportional
1106: to the following combination of coupling constants,
1107: $h^{pp}_\rho g_\rho \kappa_\rho+ h^{pp}_\omega g_\omega \kappa_\omega$.
1108: In the BONN model, the $\omega$-meson tensor coupling constant is taken to be zero,
1109: and hence the data point at 221 MeV fixes $h^{pp}_\rho$ (for given
1110: $g_\rho$, $\kappa_\rho$, and $\Lambda_\rho$). This is the reason for the
1111: $\simeq 44$ \% increase (in magnitude) of $h^{pp}_\rho$ with respect
1112: to the DDH \lq\lq best\rq\rq estimate.
1113:
1114: Below 50 MeV, however, the calculated total asymmetry
1115: is dominated by the $^1$S$_0$-$^3$P$_0$ contribution,
1116: approximately proportional to~\cite{Simonius88}
1117: $h^{pp}_\rho g_\rho ( 2 +\kappa_\rho )
1118: +h^{pp}_\omega g_\omega ( 2 +\kappa_\omega )$. The increase
1119: in magnitude of $h^{pp}_\rho$ required to fit the point at 221 MeV,
1120: now leads to a total asymmetry $\mid\!\! A(E)\!\!\mid$ below 50 MeV, which is too large
1121: when compared to experiment. Thus, in order to reproduce
1122: the 13.6 MeV and 45 MeV data points,
1123: the overall strength of the coupling constant combination above needs
1124: to be reduced significantly. Since $g_\rho ( 2 +\kappa_\rho )$ and
1125: $g_\omega ( 2 +\kappa_\omega )= 2\, g_\omega$ have the same sign, this
1126: requires making the sign of $h^{pp}_\omega$ opposite to that of $h^{pp}_\rho$.
1127:
1128: It is worth pointing out, though, that the changes in value
1129: for $h^{pp}_\rho$ and $h^{pp}_\omega$ advocated here are still compatible
1130: with the \lq\lq reasonable\rq\rq ranges for the $h^{pp}_{\rho_i}$ and
1131: $h^{pp}_{\omega_i}$, determined in Ref.~\cite{Desplanques80}.
1132:
1133: Finally, we show in Fig.~\ref{fig:asy} the total nuclear asymmetry
1134: obtained in a calculation based on the old Reid soft-core
1135: potential~\cite{Reid68,Wiringa01} and a DDH potential using
1136: the following coupling-constant and cutoff values: $g^2_\rho/4\pi = 0.95$,
1137: $g^2_\omega/4\pi = 20$, $\kappa_\rho=6.1$, $\kappa_\omega=0$,
1138: $\Lambda_\rho=1.3$ GeV/c, $\Lambda_\omega=1.5$ GeV/c (these are all
1139: from the old $r$-space version of the Bonn potential~\cite{Machleidt87}),
1140: and the \lq\lq best\rq\rq estimates for $h^{pp}_\rho$ and $h^{pp}_\omega$.
1141: These model interactions are essentially identical to those employed by Driscoll and Miller
1142: in Ref.~\cite{Driscoll89}. Indeed, our calculated total asymmetry
1143: is close to that obtained by these authors. It should be stressed that
1144: in Ref.~\cite{Driscoll89} the strong-interaction phases and mixing angles
1145: were taken from Arndt's analysis of nucleon-nucleon
1146: scattering data~\cite{Arndt87} rather than calculated from
1147: the Reid soft-core potential, as done here. This is presumably
1148: the origin of the remaining small differences between their results
1149: and ours.
1150:
1151: Figure~\ref{fig:asy_j} shows the total nuclear asymmetries
1152: obtained by including only the $J$=0 channel ($^1$S$_0$-$^3$P$_0$)
1153: and, in addition, the $J$=2 channels ($^3$P$_2$-$^1$D$_2$
1154: and $^3$F$_2$-$^1$D$_2$), and finally all (even) $J$-channels
1155: up to $J_{\rm max}$=8. We re-emphasize that in the energy range
1156: 0--350 MeV the asymmetry is dominated by the $J=$0 and 2 contributions
1157: (among the latter, specifically those from the
1158: $^3$P$_2$-$^1$D$_2$ partial waves).
1159:
1160: Figure~\ref{fig:asy_l} illustrates the sensitivity of the total
1161: nuclear asymmetry to modifications of the $\Lambda_{\rho}$ and $\Lambda_\omega$
1162: cutoff parameters in the DDH potential.
1163: Both cot-offs are multiplied by $R_{\rm cut}$, in each case
1164: $h^{pp}_\rho$ and $h^{pp}_\omega$ are then readjusted to approximately reproduce
1165: the AV18+DDH-adj combination. In the near point-like
1166: limit ($R_{\rm cut}$=10), the asymmetry increases in magnitude
1167: by roughly a factor of 2 prior to adjustment.
1168: The resulting couplings used for this case are
1169: $h^{pp}_\rho$=--14.33 and $h^{pp}_\omega$=+3.95.
1170: Results are also shown for $R_{\rm cut}$=1.5 and 0.8, the
1171: latter is an extreme case where the cutoff parameters are
1172: near the meson masses used to determine the ranges.
1173: Nevertheless, the energy dependence of
1174: the asymmetry is in all cases very similar.
1175: The couplings used to generate the two other curves are:
1176: for $R_{\rm cut}$=1.5, $h^{pp}_\rho$=--15.32 and $h^{pp}_\omega$=+3.92;
1177: and for $R_{\rm cut}$=0.8, $h^{pp}_\rho$=--106.7 and $h^{pp}_\omega$=+14.63.
1178:
1179: Figures~\ref{fig:asy_coul}--\ref{fig:asy_tra} illustrate the effects
1180: of Coulomb contributions on the longitudinal asymmetry. Figure~\ref{fig:asy_coul}
1181: compares the total nuclear asymmetry defined above (curve labeled \lq\lq C\rq\rq)
1182: with the total asymmetry obtained by ignoring the Coulomb potential
1183: altogether (curve labeled \lq\lq no-C\rq\rq). As already pointed out
1184: in Ref.~\cite{Driscoll89}, Coulomb contributions to these (non-physical)
1185: quantities are rather small.
1186:
1187: Figure~\ref{fig:asy_ang} compares the angular distribution
1188: of the (physical) longitudinal asymmetry obtained from the amplitudes
1189: $\overline{M}=M+M^{\rm C}$, see Eq.~(\ref{eq:angc}), with that
1190: calculated by replacing $\overline{M} \rightarrow M$ in Eq.(\ref{eq:angc}),
1191: namely ignoring the contribution of the Coulomb amplitude $M^{\rm C}$.
1192: The latter dominates at small scattering angles, and leads to the
1193: peculiar small angle behavior of the angular
1194: distribution shown in Fig.~\ref{fig:asy_ang}, namely its changing of sign
1195: at small $\theta$, and its vanishing at $\theta$=0
1196: (also observed in Ref.~\cite{Driscoll89}). It is interesting
1197: to note that the angular distribution at 230 MeV obtained by
1198: the authors of Ref.~\cite{Driscoll89} is significantly different
1199: from that shown in Fig.~\ref{fig:asy_ang} at 221 MeV.
1200:
1201: Figure~\ref{fig:asy_tra} illustrates the effects of Coulomb contributions
1202: on the total longitudinal asymmetry measured in
1203: transmission experiments, Eq.~(\ref{eq:asy_t}), for
1204: various choices of the critical angle $\theta_0$ ($\theta_0$=2$^\circ$, 5$^\circ$,
1205: and 10$^\circ$), see Eq.~(\ref{eq:s_tra}). Coulomb contributions are
1206: substantial, particularly for small $\theta_0$ and energies below 100 MeV.
1207: We also demonstrate, in Fig.~\ref{fig:asy_bonn}, that the
1208: angular distributions of the longitudinal asymmetry are
1209: only weakly affected by different input (strong-interaction)
1210: potentials.
1211:
1212: Finally, we present an analysis of the extracted coupling constants and
1213: their experimental errors. This analysis employs the AV18 model
1214: and the BONN-derived strong interaction couplings and cut-offs
1215: (Table~\ref{tb:gs}) in the DDH potential. Given the weak sensitivity
1216: described earlier, only small changes should be expected with the
1217: use of other recent strong interaction potentials.
1218: The experimental data at low energies have been combined into the two
1219: data points shown in Fig.~\ref{fig:asy} at 13.6 and 45 MeV.
1220: The asymmetries are --0.97 $\pm$ 0.2 and --1.53 $\pm$ 0.21, respectively,
1221: combining statistical and systematic errors.
1222: We also include the recent TRIUMF result of +0.84 $\pm$ 0.34
1223: at 221 MeV (all in units of $10^{-7}$).
1224:
1225: Figure~\ref{fig:errcontour} shows contours of constant total
1226: $\chi^2$ at levels of 1 through 5 versus the coupling
1227: constants $h^{pp}_\rho$ and $h^{pp}_\omega$.
1228: As is apparent in the
1229: figure, there is a rather narrow band of acceptable values for
1230: $h^{pp}_\rho$ and $h^{pp}_\omega$ at total $\chi^2 = 1$ for the
1231: 3 experimental data points. At this level, $h^{pp}_\rho$ can range
1232: from approximately --14 to --28, with a simultaneous
1233: (and strongly-correlated) variation in
1234: $h^{pp}_\omega$ from rougly --2 to +10, all in units of $10^{-7}$.
1235:
1236: %
1237: \section{Conclusions}
1238:
1239: We have performed an analysis of the $pp$ parity-violating (PV)
1240: longitudinal asymmetry using combinations of modern-day strong
1241: interaction potentials and the DDH PV potential. The
1242: new experimental results from TRIUMF at 221 MeV, in combination
1243: with previous results at lower energy, provide a strong constraint
1244: on allowable linear combinations of $\rho$ and $\omega$
1245: PV coupling constants. Combining the statistical and
1246: systematic errors in quadrature, $h^{pp}_\rho$ is
1247: constrained by present data to approximately 35\%,
1248: at the level of one standard deviation.
1249:
1250: The prime motivation for the present work is to initiate a
1251: systematic and consistent study of many PV
1252: observables in the few-nucleon sector, where accurate
1253: microscopic calculations are feasible.
1254: Recent measurements of nuclear anapole moments
1255: in atomic PV experiments are difficult to reconcile with
1256: earlier PV experiments in light nuclei
1257: using the simple DDH-orig model~\cite{Haxton01}.
1258: Up to now, this approach has been extremely limited
1259: by the available data.
1260:
1261: The present experimental data in the few-nucleon sector
1262: remain rather sparse, with primary constraints coming from
1263: the $p$$p$ longitudinal asymmetry
1264: analyzed here and measurements of the PV asymmetry in
1265: $p-\alpha$ elastic scattering~\cite{Lang86}.
1266: A variety of new results are expected in the next few
1267: years, however, including $^1H(\vec{n},\gamma)^2H$~\cite{Snow00},
1268: the neutron spin rotation in Helium~\cite{Heckelpc}, and electron
1269: scattering measurements at Bates and JLAB.
1270: The combination of these diverse experiments should finally yield
1271: a coherent picture of the NN PV interaction
1272: at the hadronic scale.
1273:
1274: \label{sec:cons}
1275:
1276: %
1277: %
1278: \section*{Acknowledgments}
1279: We wish to thank R.\ Machleidt for providing us with computer codes
1280: generating the latest version of the Bonn potential, and
1281: for illuminating correspondence in regard to the treatment
1282: of the Coulomb potential in momentum space.
1283: The work of J.C. and B.F.G. was supported by the
1284: U.S. Department of Energy under contract W-7405-ENG-36, and
1285: the work of R.S. was supported by DOE contract DE-AC05-84ER40150
1286: under which the Southeastern Universities Research Association (SURA)
1287: operates the Thomas Jefferson National Accelerator Facility.
1288: Finally, some of the calculations were made possible by grants
1289: of computing time from the National Energy Research Supercomputer
1290: Center.
1291: %
1292: %
1293: %
1294: \begin{references}
1295: %
1296: %
1297: \bibitem{Berdoz01} A.R.\ Berdoz {\em et al.},
1298: {\tt nucl-ex/0107014}, submitted to Phys.\ Rev.\ Lett.\ (2001).
1299: %
1300: \bibitem{Snow00} W.\ M.\ Snow, {\em et al.},
1301: Nucl. Instrum. Meth. {\bf A440}, 729 (2000).
1302: %
1303: \bibitem{Jlab01} B.\ Wojtsekhowski and W.T.H.\ van Oers (spokespersons),
1304: JLAB letter-of-intent 00-002.
1305: %
1306: \bibitem{Hasty00} R.\ Hasty {\em et al.},
1307: Science {\bf 290}, 2021 (2000).
1308: %
1309: \bibitem{Spayde00} D.T.\ Spayde {\em et al.},
1310: Phys.\ Rev.\ Lett.\ {\bf 84}, 1106 (2000).
1311: %
1312: \bibitem{Nagle79} D.E.\ Nagle {\em et al.},
1313: in {\it High Energy Physics with Polarized Beams and Polarized Targets--1978},
1314: edited by G.H.\ Thomas, AIP Conference Proceedings No. 51
1315: (American Institute of Physics, New York, 1979), p. 218.
1316: %
1317: \bibitem{Balzer84} R.\ Balzer {\em et al.},
1318: Phys.\ Rev.\ C {\bf 30}, 1409 (1984).
1319: %
1320: \bibitem{Eversheim91} P.D.\ Eversheim {\em et al.},
1321: Phys.\ Lett.\ B {\bf 256}, 11 (1991).
1322: %
1323: \bibitem{Kistryn87} S.\ Kistryn {\em et al.},
1324: Phys.\ Rev.\ Lett.\ {\bf 58}, 1616 (1987).
1325: %
1326: \bibitem{Yuan86} V.\ Yuan {\em et al.},
1327: Phys.\ Rev.\ Lett.\ {\bf 57}, 1680 (19886).
1328: %
1329: \bibitem{Simonius72} M.\ Simonius,
1330: Phys.\ Lett.\ {\bf 41B}, 415 (1972).
1331: %
1332: \bibitem{Brown74} V.R.\ Brown, E.M.\ Henley, and F.R.\ Krejs,
1333: Phys.\ Rev.\ C {\bf 9}, 935 (1973); Phys. Rev. Lett. {\bf 30}, 770 (1973).
1334: %
1335: \bibitem{Henley75} E.M.\ Henley and F.R.\ Krejs,
1336: Phys.\ Rev.\ D {\bf 11}, 605 (1975).
1337: %
1338: \bibitem{Oka81} T.\ Oka,
1339: Prog.\ Theor.\ Phys.\ {\bf 66}, 977 (1981).
1340: %
1341: \bibitem{Driscoll89} D.E.\ Driscoll and G.A.\ Miller,
1342: Phys.\ Rev.\ C {\bf 39}, 1951 (1989).
1343: %
1344: \bibitem{Desplanques80} B.\ Desplanques, J.F.\ Donoghue, and B.R.\ Holstein,
1345: Ann.\ Phys.\ (N.Y.) {\bf 124}, 449 (1980).
1346: %
1347: \bibitem{Wiringa95} R.B.\ Wiringa, V.G.J.\ Stoks, and R.\ Schiavilla,
1348: Phys.\ Rev.\ C {\bf 51}, 38 (1995).
1349: %
1350: \bibitem{Stoks94} V.G.J.\ Stoks, R.A.M.\ Klomp, C.P.F.\ Terheggen, and J.J.\ de Swart,
1351: Phys.\ Rev.\ C {\bf 49}, 2950 (1994).
1352: %
1353: \bibitem{Machleidt01} R.\ Machleidt,
1354: Phys.\ Rev.\ C {\bf 63}, 024001 (2001).
1355: %
1356: \bibitem{Bergervoet90} J.R.\ Bergervoet {\em et al.},
1357: Phys.\ Rev.\ C {\bf 41}, 1435 (1990).
1358: %
1359: \bibitem{Stoks93} V.G.J.\ Stoks, R.A.M.\ Klomp, M.C.M.\ Rentmeester, and J.J.\ de Swart,
1360: Phys.\ Rev.\ C {\bf 48}, 792 (1993).
1361: %
1362: \bibitem{Przewoski98} B.v.\ Przewoski {\em et al.},
1363: Phys.\ Rev.\ C {\bf 58}, 1897 (1998).
1364: %
1365: \bibitem{Friar77} J.L. Friar,
1366: Ann.\ Phys.\ (N.Y.) {\bf 104}, 380 (1977).
1367: %
1368: \bibitem{Forest00} J.L.\ Forest,
1369: Phys.\ Rev.\ C {\bf 61}, 034007 (2000).
1370: %
1371: \bibitem{Coon86} S.A.\ Coon and J.L.\ Friar,
1372: Phys.\ Rev.\ C {\bf 34}, 1060 (1986).
1373: %
1374: \bibitem{Schiavilla01} R.Schiavilla,
1375: Nucl.\ Phys.\ A {\bf 689}, 84c (2001).
1376: %
1377: \bibitem{Goldberger64} M.L.\ Goldberger and K.M.\ Watson,
1378: {\it Collision Theory} (Wiley, New York, 1964).
1379: %
1380: \bibitem{Abramowitz74} M.\ Abramowitz and I.S.\ Stegun,
1381: {\it Handbook of Mathematical Functions} (Dover, New York, 1974).
1382: %
1383: \bibitem{Holdeman65} J.T.\ Holdeman and R.M.\ Thaler,
1384: Phys.\ Rev.\ Lett.\ {\bf 14}, 81 (1965).
1385: %
1386: \bibitem{Vincent74} C.M.\ Vincent and S.C.\ Phatak,
1387: Phys.\ Rev.\ C {\bf 10}, 391 (1974).
1388: %
1389: \bibitem{Gloeckle83} W.\ Gl\"ockle,
1390: {\it The Quantum Mechanical Few-Body Problem} (Springer-Verlag, Berlin, 1983).
1391: %
1392: \bibitem{Machleidt01a} R.\ Machleidt,
1393: private communication.
1394: %
1395: \bibitem{vanOers01} W.T.H.\ van Oers,
1396: private communication.
1397: %
1398: \bibitem{Simonius88} M.\ Simonius,
1399: Can.\ J.\ Phys.\ {\bf 66}, 548 (1988).
1400: %
1401: \bibitem{Reid68} R.V.\ Reid,
1402: Ann.\ Phys.\ (N.Y.) {\bf 50}, 411 (1968).
1403: %
1404: \bibitem{Wiringa01} R.B.\ Wiringa,
1405: private communication.
1406: %
1407: \bibitem{Machleidt87} R.\ Machleidt, K.\ Holinde, and Ch.\ Elster,
1408: Phys.\ Rep.\ {\bf 149}, 1 (1987).
1409: %
1410: \bibitem{Arndt87} R.A.\ Arndt, J.S.\ Hyslop III, and L.D.\ Roper,
1411: Phys.\ Rev.\ D {\bf 35}, 128 (1987).
1412: %
1413: \bibitem{Haxton01} W.C.\ Haxton, C.P.\ Liu, M.J.\ Ramsey-Musolf,
1414: preprint nucl-th/0109014.
1415: %
1416: \bibitem{Lang86} J.\ Lang {\em et al.},
1417: Phys.\ Rev.\ C {\bf34}, 1545 (1986).
1418: %
1419: \bibitem{Heckelpc} B.\ Heckel, private communication.
1420: %
1421: %
1422: \end{references}
1423: %
1424: %
1425: %
1426: \begin{table}
1427: \caption{Values used for the strong- and weak-interaction
1428: coupling constants of the $\rho$- and $\omega$-meson to the nucleon, see text.}
1429: \begin{tabular}{cccccc}
1430: & $g^2_\alpha/4\pi$ & $\kappa_\alpha$ & $10^7 h_\alpha^{pp}$ (DDH-adj) & $10^7 h_\alpha^{pp}$ (DDH-orig) & $\Lambda_\alpha$ (GeV/c) \\
1431: \tableline
1432: $\rho$ & 0.84 & 6.1 & --22.3 & --15.5 & 1.31 \\
1433: $\omega$ & 20. & 0. & +5.17 & --3.04 & 1.50
1434: \end{tabular}
1435: \label{tb:gs}
1436: \end{table}
1437: %
1438: %
1439: %
1440: \begin{table}
1441: \caption{Labeling of channels.}
1442: \begin{tabular}{cccc}
1443: & \multicolumn{3}{c} {$\alpha$} \\
1444: $J$ & 1 & 2 & 3 \\
1445: \tableline
1446: 0 & $^1$S$_0$ & $^3$P$_0$ & \\
1447: 1 & $^3$P$_1$ & & \\
1448: 2 & $^3$P$_2$ & $^3$F$_2$ & $^1$D$_2$ \\
1449: 3 & $^3$F$_3$ & & \\
1450: 4 & $^3$F$_4$ & $^3$H$_4$ & $^1$G$_4$ \\
1451: $\dots$ & $\dots$ & $\dots$ & $\dots$
1452: \end{tabular}
1453: \label{tb:chan}
1454: \end{table}
1455: %
1456: %
1457: \begin{table}
1458: \caption{Classification of channel mixings for $J$ even: PC or PV
1459: if induced by $v^{\rm PC}$ or $v^{\rm PV}$, respectively.}
1460: \begin{tabular}{cccc}
1461: & \multicolumn{3}{c} {coupling} \\
1462: $J$ & 12 & 13 & 23 \\
1463: \tableline
1464: 0 & PV & & \\
1465: 2 & PC & PV & PV \\
1466: 4 & PC & PV & PV \\
1467: $\dots$ & PC & PV & PV
1468: \end{tabular}
1469: \label{tb:mixing}
1470: \end{table}
1471: %
1472: %
1473: %
1474: \begin{figure}[bth]
1475: \let\picnaturalsize=N
1476: \def\picsize{4in}
1477: \def\picfilenamea{asy_pp.eps}
1478: \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi
1479: \let\epsfloaded=Y
1480: \centerline{
1481: \ifx\picnaturalsize N\epsfxsize \picsize\fi \epsfbox{\picfilenamea}
1482: }}\fi
1483: \caption{Total nuclear asymmetries obtained with various combinations
1484: of strong- and weak-interaction potentials are compared to data, see text.}
1485: \label{fig:asy}
1486: \end{figure}
1487: %
1488: %
1489: \begin{figure}[bth]
1490: \let\picnaturalsize=N
1491: \def\picsize{4in}
1492: \def\picfilenamea{asy_pp_j.eps}
1493: \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi
1494: \let\epsfloaded=Y
1495: \centerline{
1496: \ifx\picnaturalsize N\epsfxsize \picsize\fi \epsfbox{\picfilenamea}
1497: }}\fi
1498: \caption{Contributions to the total nuclear asymmetry
1499: obtained by including only the $J$=0 channel, and by adding the
1500: $J$=2 channels, and finally all even $J$-channels up to $J_{\rm max}$=8.
1501: The AV18- and DDH-adj potential combination is used, thin solid line
1502: in Fig.~\protect{\ref{fig:asy}}.}
1503: \label{fig:asy_j}
1504: \end{figure}
1505: %
1506: %
1507: \begin{figure}[bth]
1508: \let\picnaturalsize=N
1509: \def\picsize{4in}
1510: \def\picfilenamea{asy_pp_lnew.eps}
1511: \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi
1512: \let\epsfloaded=Y
1513: \centerline{
1514: \ifx\picnaturalsize N\epsfxsize \picsize\fi \epsfbox{\picfilenamea}
1515: }}\fi
1516: \caption{Sensitivity of the total nuclear asymmetry to modifications of
1517: the cutoff-parameters $\Lambda_\rho$ and $\Lambda_\omega$ in the DDH potential.
1518: Both cut-offs are multiplied by $R_{\rm cut}$, see text. For each case
1519: the couplings are then adjusted to approximately reproduce the
1520: results obtained with the AV18 and DDH-adj potential combination (corresponding
1521: to $R_{\rm cut}$=1) in Fig.~\protect{\ref{fig:asy}}.}
1522: \label{fig:asy_l}
1523: \end{figure}
1524: %
1525: %
1526: \begin{figure}[bth]
1527: \let\picnaturalsize=N
1528: \def\picsize{4in}
1529: \def\picfilenamea{asy_coul.eps}
1530: \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi
1531: \let\epsfloaded=Y
1532: \centerline{
1533: \ifx\picnaturalsize N\epsfxsize \picsize\fi \epsfbox{\picfilenamea}
1534: }}\fi
1535: \caption{The total nuclear asymmetry, see text, is compared to the total
1536: asymmetry obtained by ignoring the Coulomb potential.
1537: The AV18 and DDH-adj potential combination is used, thin solid line
1538: in Fig.~\protect{\ref{fig:asy}}.}
1539: \label{fig:asy_coul}
1540: \end{figure}
1541: %
1542: %
1543: \begin{figure}[bth]
1544: \let\picnaturalsize=N
1545: \def\picsize{4in}
1546: \def\picfilenamea{asy_pp_ang.eps}
1547: \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi
1548: \let\epsfloaded=Y
1549: \centerline{
1550: \ifx\picnaturalsize N\epsfxsize \picsize\fi \epsfbox{\picfilenamea}
1551: }}\fi
1552: \caption{Angular distributions obtained with (curve labeled C)
1553: and without (curve labeled no-C) inclusion of the Coulomb
1554: amplitude at 45 MeV and 221 MeV, see text.
1555: The AV18 and DDH-adj potential combination is used, thin solid line
1556: in Fig.~\protect{\ref{fig:asy}}.}
1557: \label{fig:asy_ang}
1558: \end{figure}
1559: %
1560: %
1561: \begin{figure}[bth]
1562: \let\picnaturalsize=N
1563: \def\picsize{4in}
1564: \def\picfilenamea{asy_pp_tra.eps}
1565: \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi
1566: \let\epsfloaded=Y
1567: \centerline{
1568: \ifx\picnaturalsize N\epsfxsize \picsize\fi \epsfbox{\picfilenamea}
1569: }}\fi
1570: \caption{The total nuclear asymmetry (thick solid line) is compared
1571: with the total asymmetries measured in transmission experiments with critical
1572: angles $\theta_0$=2$^\circ$, 5$^\circ$, and 10$^\circ$ (curves
1573: labeled $A\mid_{>2}$, $A\mid_{>5}$, and $A\mid_{>10}$, respectively).
1574: The AV18 and DDH-adj potential combination is used, thin solid line
1575: in Fig.~\protect{\ref{fig:asy}}.}
1576: \label{fig:asy_tra}
1577: \end{figure}
1578: %
1579: %
1580: \begin{figure}[bth]
1581: \let\picnaturalsize=N
1582: \def\picsize{4in}
1583: \def\picfilenamea{asy_pp_b.eps}
1584: \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi
1585: \let\epsfloaded=Y
1586: \centerline{
1587: \ifx\picnaturalsize N\epsfxsize \picsize\fi \epsfbox{\picfilenamea}
1588: }}\fi
1589: \caption{Angular distributions of the longitudinal asymmetry
1590: obtained with either the AV18 or BONN potential in combination with
1591: the DDH-adj potential.}
1592: \label{fig:asy_bonn}
1593: \end{figure}
1594: %
1595: %
1596: \begin{figure}[bth]
1597: \let\picnaturalsize=N
1598: \def\picsize{4in}
1599: \def\picfilenamea{contour-bw.eps}
1600: \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi
1601: \let\epsfloaded=Y
1602: \centerline{
1603: \ifx\picnaturalsize N\epsfxsize \picsize\fi \epsfbox{\picfilenamea}
1604: }}\fi
1605: \caption{Curves of constant total $\chi^2$ obtained by
1606: analyzing the experimental $pp$ data with the AV18 model,
1607: and $\rho$- and $\omega$-meson strong-interaction couplings in the DDH potential
1608: from the Bonn-2000 model. The curves indicate surfaces
1609: of total $\chi^2$ = 1,2,3,4, and 5 for various values of
1610: $h^{pp}_\rho$ and $h^{pp}_\omega$.}
1611: \label{fig:errcontour}
1612: \end{figure}
1613: %
1614: \end{document}
1615: