hep-ph0312044/ex3.tex
1: %% ****** Start of file template.aps ****** %
2: %%
3: %%
4: %%   This file is part of the APS files in the REVTeX 4 distribution.
5: %%   Version 4.0 of REVTeX, August 2001
6: %%
7: %%
8: %%   Copyright (c) 2001 The American Physical Society.
9: %%
10: %%   See the REVTeX 4 README file for restrictions and more information.
11: %%
12: %
13: % This is a template for producing manuscripts for use with REVTEX 4.0
14: % Copy this file to another name and then work on that file.
15: % That way, you always have this original template file to use.
16: %
17: % Group addresses by affiliation; use superscriptaddress for long
18: % author lists, or if there are many overlapping affiliations.
19: % For Phys. Rev. appearance, change preprint to twocolumn.
20: % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
21: %     important  ===>   $$$   <===   important     %
22: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
23: %  Add 'draft' option to mark overfull boxes with black boxes
24: %  Add 'showpacs' option to make PACS codes appear
25: %  Add 'showkeys' option to make keywords appear
26: \documentclass[aps,prd,preprint,groupedaddress]{revtex4}
27: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
28: %\documentclass[aps,prl,twocolumn,groupedaddress]{revtex4}
29: \usepackage{graphicx}% Include figure files
30: 
31: % You should use BibTeX and apsrev.bst for references
32: % Choosing a journal automatically selects the correct APS
33: % BibTeX style file (bst file), so only uncomment the line
34: % below if necessary.
35: \bibliographystyle{apsrev}
36: 
37: \begin{document}
38: 
39: % Use the \preprint command to place your local institutional report
40: % number in the upper righthand corner of the title page in preprint mode.
41: % Multiple \preprint commands are allowed.
42: % Use the 'preprintnumbers' class option to override journal defaults
43: % to display numbers if necessary
44: \preprint{OU-HEP-459}
45: %hep-ph/0312044
46: 
47: %Title of paper
48: \title{$\mbox{\boldmath $\pi N$}$ sigma-term and chiral-odd twist-3
49: distribution function $\mbox{\boldmath $e(x)$}$\\
50: of the nucleon in the chiral quark soliton model}
51: 
52: % repeat the \author .. \affiliation  etc. as needed
53: % \email, \thanks, \homepage, \altaffiliation all apply to the current
54: % author. Explanatory text should go in the []'s, actual e-mail
55: % address or url should go in the {}'s for \email and \homepage.
56: % Please use the appropriate macro foreach each type of information
57: 
58: % \affiliation command applies to all authors since the last
59: % \affiliation command. The \affiliation command should follow the
60: % other information
61: % \affiliation can be followed by \email, \homepage, \thanks as well.
62: \author{Y.~Ohnishi and M.~Wakamatsu}
63: \email[]{ohnishi@kern1.nclth.osaka-u.ac.jp}
64: \email[]{wakamatu@phys.sci.osaka-u.ac.jp}
65: %\homepage[]{Your web page}
66: %\thanks{}
67: %\altaffiliation{}
68: \affiliation{Department of Physics, Faculty of Science, \\
69: Osaka University, \\
70: Toyonaka, Osaka 560, JAPAN}
71: 
72: %Collaboration name if desired (requires use of superscriptaddress
73: %option in \documentclass). \noaffiliation is required (may also be
74: %used with the \author command).
75: %\collaboration can be followed by \email, \homepage, \thanks as well.
76: %\collaboration{}
77: %\noaffiliation
78: 
79: %\date{\today}
80: 
81: \begin{abstract}
82: The isosinglet combination of the chiral-odd twist-3 distribution function
83: $e^u(x)+e^d(x)$ of the nucleon has outstanding properties that its
84: first moment is proportional to the well-known $\pi N$ sigma-term and
85: that it contains a $\delta$-function singularity at $x=0$.
86: These two features are inseparably connected in that the above sum rule
87: would be violated, if there is no such a singularity in $e^u(x)+e^d(x)$. 
88: Very recently, we found that the physical origin of this 
89: $\delta$-function singularity can be traced back to the long-range
90: quark-quark correlation of scalar type, which signals the spontaneous
91: chiral symmetry breaking of the QCD vacuum.
92: The main purpose of the present paper is to give
93: complete theoretical predictions for the chiral-odd twist-3 distribution
94: function $e^a(x)$ of each flavor $a$ on the basis of the chiral quark
95: soliton model, without recourse to the derivative expansion type
96: approximation. These theoretical predictions are then compared with
97: the empirical information extracted from the CLAS data of the
98: semi-inclusive DIS processes by assuming the Collins mechanism only.
99: A good agreement with the CLAS data is indicative of a
100: sizable violation of the $\pi N$ sigma-term sum rule, or equivalently,
101: the existence of a $\delta$-function singularity in $e^u(x) + e^d(x)$.
102: \end{abstract}
103: 
104: 
105: % insert suggested PACS numbers in braces on next line
106: \pacs{12.39.Fe, 12.39.Ki, 12.38.Lg, 13.40.Em}
107: % insert suggested keywords - APS authors don't need to do this
108: %\keywords{}
109: 
110: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
111: \maketitle
112: 
113: % body of paper here - Use proper section commands
114: % References should be done using the \cite, \ref, and \label commands
115: 
116: 
117: \section{Introduction}
118: 
119: \ \ \ It is a widely accepted common belief now that the nonperturbative
120: dynamics of QCD (chiral dynamics) is an indispensable element to
121: understand high-energy deep inelastic scattering observables.
122: Undoubtedly, the reconfirmation of this natural fact strongly owes
123: to the two remarkable experimental discoveries in this
124: field \cite{EMC88}\nocite{EMC89}-\cite{NMC91}.
125: They are the unexpectedly small quark spin fraction 
126: of the nucleon revealed by the EMC measurement \cite{EMC88},\cite{EMC89}
127: and the light-flavor sea-quark asymmetry confirmed by the NMC
128: measurement \cite{NMC91}. The most successful theoretical studies of
129: parton distribution functions have been carried out within the
130: framework of the chiral quark soliton 
131: model (CQSM) \cite{DPPPW96}\nocite{DPPPW97}\nocite{WGR96}\nocite{WGR97}
132: \nocite{WK98}\nocite{WK99}\nocite{PPGWW99}\nocite{MW00}\nocite{MW01}
133: \nocite{MW03A}\nocite{MW03B}-\cite{DGPW99}, which is an effective model
134: of baryons maximally incorporating the spontaneous chiral symmetry
135: breaking of the QCD vacuum.
136: In fact, we claim that it is so far the only one effective model of
137: baryons which are able to explain the above two remarkable findings
138: simultaneously within the single theoretical framework
139: \cite{WY91}\nocite{WW00A}\nocite{MW92}-\cite{WW00B}.
140: 
141: Very recently, we became aware of another novel example in which
142: nonperturbative QCD dynamics play an unprecedented role in the
143: physics of parton distribution functions.
144: It concerns the possible existence of a delta-function singularity at
145: the Bjorken variable $x = 0$ in the chiral-odd twist-3 distribution
146: function $e(x)$ of the nucleon  \cite{S03},\cite{WO03}.
147: This distribution function itself, together with its first moment 
148: sum rule giving the familiar $\pi N$ sigma-term, have been known
149: for a long time \cite{JLS73}.
150: In spite of several interesting theoretical features, 
151: however, this distribution function has been thought of as an academic 
152: object of study, since, because of its chiral-odd nature, it does not 
153: appear in the cross section formula of inclusive DIS scatterings.
154: The situation changed drastically, however, since the CLAS 
155: Collaboration was able to get the first experimental information
156: on this interesting quantity through the measurement of the azimuthal
157: asymmetry $A_{LU}$ in the electroproduction of pions from deeply
158: inelastic scattering of longitudinally polarized electrons off
159: unpolarized protons \cite{CLAS}\nocite{herm01}-\cite{herm02}.
160: 
161: Some years ago, within the framework of perturbative QCD, Burkardt 
162: and Koike noticed that the first moment sum rule (or the $\pi N$
163: sigma-term sum rule) for $e(x)$ holds only when $e(x)$ has a 
164: $\delta$-function type singularity at the Bjorken variable 
165: $x = 0$ \cite{BK02}.
166: Unfortunately, the physical origin of this singular term is 
167: not very clear in this perturbative analysis. Very recently, two 
168: independent proofs were given to the fact that the physical origin of 
169: this $\delta$-function singularity can be traced back to 
170: the nonvanishing vacuum quark condensate which signals the spontaneous 
171: chiral symmetry breaking of the QCD vacuum~\cite{S03},\cite{WO03}.
172: An interesting question is whether we can verify  experimentally
173: the existence of this $\delta$-function singularity in $e(x)$.
174: Unfortunately, the point $x = 0$ is experimentally inaccessible.
175: This means that, if there really exists such a $\delta (x)$-type
176: singularity in $e (x)$, the experimental measurement would rather 
177: confirm violation of this $\pi N$ sigma-term sum rule.
178: Nonetheless, since $e (x)$ in the region $ x \neq 0$ can in principle
179: be measured, theorists are challenged to explain its behavior.
180: 
181: The first theoretical study of $e (x)$ was done by using the MIT bag 
182: model \cite{JJ92}. (See also \cite{Signal97}.)
183: However, this estimate based on the bag model cannot be taken
184: as a realistic one by the following reasons.
185: First, its prediction for the magnitude of the $\pi N$ sigma-term is far
186: from reliable. Second, more seriously, it cannot reproduce the 
187: $\delta$-function singularity of $e (x)$. Both these features 
188: (they are not actually unrelated) are easily anticipated, since the 
189: MIT bag model is essentially a relativistic quark model with $N_c \,(=3)$
190: valence quark degrees of freedom only, and the reproduction of the 
191: nonzero vacuum quark condensate is beyond the range of applicability
192: of this model. The first realistic investigation of $e(x)$ was carried out
193: by Efremov et al. on the basis of the chiral quark soliton model but
194: within the ``valence" quark only approximations \cite{EGS01},\cite{EGS02}.
195: More recently, the present authors and Schweitzer independently
196: carried out more careful analysis of the contribution of the Dirac sea
197: quarks on the basis of the gradient expansion type approximation and
198: confirmed that the isosinglet combination of $e(x)$ certainly contains
199: $\delta$-function type singularity \cite{S03},\cite{WO03}.
200: After some analysis of higher derivative terms of the expansion,
201: however, Schweitzer retreated to the assumption that the contribution
202: of the Dirac sea quarks is saturated by this $\delta(x)$ term alone.
203: As admitted by himself, however, whether this last assumption is justified
204: or not is far from trivial \cite{S03}.
205: To confirm it, one has to carry out an 
206: exact numerical calculation within the model without recourse to the
207: gradient expansion type approximation. Furthermore, to compare the 
208: predictions of the model with the experimental data of the CLAS
209: collaboration, one must know $e^a (x)$ of each flavor $a$.
210: To this end, only the knowledge 
211: of the isoscalar combination $e^u (x) + e^d (x)$ is not enough.
212: We need another independent combination, i.e. the isovector distribution 
213: $e^u (x) - e^d (x)$. Within the framework of the CQSM, this latter 
214: distribution survives at the next-to-leading order in $1 / N_c$ 
215: expansion and it was left untouched in \cite{S03}.
216: 
217: In view of the above-mentioned circumstances, we think it
218: important to carry out an exact model calculation within the CQSM
219: for both of the isoscalar and isovector combinations of the chiral-odd
220: twist-3 distribution function $e (x)$. 
221: We also think it useful to analyze the first and the second moment sum
222: rule for $e^u (x) + e^d (x)$ and $e^u (x) - e^u (x)$ within the CQSM 
223: in the light of the corresponding sum rule expected in the general
224: framework of perturbative QCD.
225: The predictions of the model for $e^u (x)$ and $e^d (x)$ (as well as 
226: the corresponding distributions for antiquarks) are then used as initial 
227: distributions given at the model energy scale around $600 \,\mbox{MeV}$
228: (or $Q^2 \simeq 0.30 \,\mbox{GeV}^2$), and they are evolved to higher
229: $Q^2$ for the sake of comparison with the phenomenological information
230: obtained by using the CLAS measurement.
231: 
232: The paper is organized as follows. In sect.2, after a brief introduction
233: of the basic idea of the CQSM, the theoretical expression for
234: $e^u (x) + e^d (x)$ and $e^u (x) - e^d (x)$ are given.
235: The fundamental moment sum rules for these distributions are also discussed
236: here in some detail. Sect. 3 is devoted to the discussion of the numerical
237: results. Finally, in sect.4, we summarize what we have found. 
238: 
239: 
240: 
241: \section{$\mbox{\boldmath $e(x)$}$ in the Chiral Quark Soliton Model}
242: 
243: %\subsection{Light-cone representation for $\mbox{\boldmath $e(x)$}$}
244: 
245: The chiral-odd twist-3 quark distribution  $e^a(x)$ of flavor $a$
246: inside a nucleon with 4-momentum $P$, averaged over its spin, is
247: defined by
248: %
249: \begin{equation}
250:  e^a(x)=
251:  P^+\int_{-\infty}^{\infty} \frac{dz^-}{2\pi} \,e^{ixP^+ z^-}\langle
252:  N|\psi_a^{\dagger}(0)\gamma^0\psi_a(z)|N \rangle
253:  \bigr| _{z^+=0,z_{\perp}=0}
254:  \label{eq:dis1} \,\, ,
255: \end{equation}
256: %
257: where $\psi_a$ are quark fields. Similarly, the corresponding
258: antiquark distribution is defined as
259: %
260: \begin{equation}
261:  e^{\bar{a}}(x)=
262:  P^+\int_{-\infty}^{\infty} \frac{dz^-}{2\pi} \,e^{ixP^+ z^-}\langle 
263:  N|\psi_a^{c \dagger} (0) \gamma^0 \psi_a^c (z)|N \rangle
264:  \bigr| _{z^+=0,z_{\perp}=0}
265:  \label{eq:dis2} \,\, ,
266: \end{equation}
267: %
268: with $\psi^c$ being the charge-conjugate field of $\psi$.
269: Here we use the standard light-cone coordinates
270: %
271: \begin{equation}
272:  z^{\pm}=\frac{z^0\pm z^3}{\sqrt{2}},\qquad P^{\pm}=
273:  \frac{P^0\pm P^3}{\sqrt{2}}.
274:  \label{eq:dis3}
275: \end{equation}
276: %
277: The variable $x$ denotes the Bjorken variable, $x=-q^2/(2P\cdot q)$,
278: with $q$ being the 4-momentum transfer to the nucleon.
279: Taking account of the charge-conjugation property of the relevant quark
280: bilinear operator, one can formally extend the domain of quark distribution
281: functions from the interval
282: $0\!\le\! x\! \le\! 1$ to $-1\!\le\! x\!\le\! 1$, such that
283: %
284: \begin{equation}
285:  e^{\bar{a}}(x) = e^a(-x),\quad (0\le x\le 1),
286:  \label{eq:disqa}
287: \end{equation}
288: %
289: which dictates that the distribution function with negative $x$ should be
290: interpreted as antiquark one.
291: 
292: Although the above definitions of the quark and antiquark distribution
293: functions are frame-independent, it is convenient to perform the
294: actual calculation in the nucleon rest frame. In this frame,
295: we have $P^+\!=\!M_N/\sqrt{2}$, and the distribution function is reduced
296: to
297: %
298: \begin{equation}
299:  e^a(x)=M_N\int_{-\infty}^{\infty}\frac{dz_0}{2\pi} \,
300:  e^{ixM_Nz_0}\langle N|
301:  \psi_a^{\dagger}(0)\gamma^0\psi_a(z)|N\rangle|_{
302:  z_3=-z_0 ,z_{\perp}=0}\label{eq:dis4}\, .
303:  \label{eq:disprest}
304: \end{equation}
305: %
306: Throughout the paper, we will confine ourselves to two flavor case of
307: $u$- and $d$-quarks, and neglect strangeness degrees of freedom
308: in the nucleon.
309: Consequently, we have two independent distributions, i.e. the
310: isosinglet distribution $e^{(T=0)}(x) \equiv e^u(x) + e^d(x)$ and the
311: isovector one $e^{(T=1)}(x) \equiv e^u(x) - e^d(x)$.
312: In the case of $e^{(T=0)}(x)$, we simply sum up (\ref{eq:dis4}) over
313: the flavor components. On the other hand, for $e^{(T=1)}(x)$, 
314: we have to sum up the representation after inserting $\tau_3$ matrix
315: in (\ref{eq:dis4}).
316: 
317: For obtaining quark distribution functions, we must generally evaluate
318: nucleon matrix elements of bilocal and bilinear quark operators
319: containing two space-time coordinates with light-cone separation. 
320: The startingpoint of our theoretical analysis is the following path
321: integral representation of the matrix elements of a bilocal and
322: bilinear quark operator between the nucleon states with definite
323: momentum
324: $\mbox{\boldmath $P$}$ :
325: %
326: \begin{eqnarray}
327:   \langle N (\mbox{\boldmath $P$}) \,| \,\psi^\dagger (0) \,\gamma^0 \,
328:   \psi(z) \,| \,N (\mbox{\boldmath $P$})\rangle
329:   &=& \frac{1}{Z} \,\,\int \,\,d^3 x  \,\,d^3 y \,\,
330:   e^{\,- \,i \mbox{\boldmath $P$} \cdot \mbox{\boldmath $x$}} \,\,
331:   e^{\,i \,\mbox{\boldmath $P$} \cdot \mbox{\boldmath $y$}} \,\,
332:   \int {\cal D} U \nonumber \\
333:   &&{}\times \int {\cal D}
334:   \psi \,\,{\cal D} \psi^\dagger \,\,
335:   J_N (\frac{T}{2}, \mbox{\boldmath $x$}) \,\,\psi^\dagger (0) \,
336:   \gamma^0 \,\psi(z) \,\,J_N^\dagger (-\frac{T}{2}, 
337:   \mbox{\boldmath $y$})\nonumber \\
338:   &&{}\times \exp \,[\,\,i \int \,d^4 x \,\,\bar{\psi} \,\,
339:   (\,i \! \not\!\partial \,- \,M
340:   U^{\gamma_5}) \,\psi \, ] \, , \ \ \ \ \
341:   \label{eq:cqsm3} 
342: \end{eqnarray}
343: %
344: where 
345: %
346: \begin{eqnarray}
347:   {\cal L} \ \ = \ \ \bar{\psi} \,(\,i \not\!\partial \ - \ 
348:   M U^{\gamma_5} (x) \,) \,\psi  ,\hspace{10mm}
349:   \label{eq:cqsm4}
350: \end{eqnarray}
351: %
352: with
353: %
354: \begin{equation}
355:  U^{\gamma_5} (x) = \exp [ \,i \gamma_5 \mbox{\boldmath $\tau$}
356:  \cdot \mbox{\boldmath $\pi$} (x) / f_\pi \,] ,
357:  \label{eq:ufield}
358: \end{equation}
359: %
360: being the basic lagrangian of the CQSM with two flavors. The quantity
361: %
362: \begin{equation}
363:  J_N (x) \ \ = \ \ \frac{1}{N_c !} \,\, 
364:  \epsilon^{\alpha_1 \cdots \alpha_{N_c}} \,\,
365:  \Gamma_{J J_3, T T_3}^{\{f_1 \cdots f_{N_c}\}} \,\,
366:  \psi_{\alpha_1 f_1} (x) \cdots \psi_{\alpha_{N_c} f_{N_c}} (x) \,\, ,
367:  \label{eq:intfield}
368: \end{equation}
369: %
370: is a composite operator carrying quantum numbers $J J_3, T T_3$
371: (spin, isospin) of the baryon, where $\alpha_i$ the color indices, while 
372: $\Gamma_{JJ_3,TT_3}^{\{f_1\cdots f_{N_c}\}}$ is a symmetric matrix in spin 
373: flavor indices $f_i$. We start with a stationary pion field configuration
374: of hedgehog shape :
375: %
376: \begin{equation}
377:  \mbox{\boldmath $\pi$}(x)=f_{\pi}\hat{\mbox{\boldmath $r$}}F(r).
378:  \label{eq:hedgehog}
379: \end{equation}
380: %
381: Next we carry out the path integral over $\mbox{\boldmath $\pi$}(x)$ in
382: a saddle point approximation by taking care of two zero-energy modes,
383: i.e. the ``translational zero-modes'' and ``rotational zero-modes''.
384: Under the assumption of ``slow rotation'' as compared with the intrinsic
385: quark motion, the answers can be obtained in a perturbative series in
386: $\Omega$, which can also be regarded as a $1/N_c$ expansion.
387: Up to the first order in the collective rotational velocity
388: $\Omega$, the only surviving contribution to
389: $e^{(T=0)}(x)$ arises at the ${\cal O}(\Omega^0)$ term of this
390: expansion, since the ${\cal O}(\Omega^1)$ term vanishes identically
391: due to the hedgehog symmetry. On the other hand,
392: the first nonvanishing contribution to $e^{(T=1)}(x)$ arises
393: at the ${\cal O}(\Omega^1)$, since the leading ${\cal O}(\Omega^0)$
394: contribution vanishes due to the hedgehog symmetry.
395: Then, between the magnitude of the above two distributions, one may
396: expect the following large-$N_c$ relation :
397: %
398: \begin{equation}
399:  |e^u(x)+e^d(x)| \sim N_c \,|e^u(x)-e^d(x)|\, .
400:  \label{eq:N_c-rel}
401: \end{equation}
402: 
403: 
404: 
405: \subsection{Isosinglet distribution \mbox{\boldmath $e^{(T=0)}(x)$}}
406: 
407: The isosinglet combination of the chiral-odd twist-3 unpolarized
408: distribution is given by
409: %
410: \begin{equation}
411:  e^{(T=0)}(x) \equiv e^u(x)+e^d(x)=M_N\int^{\infty}_{-\infty}
412:  \frac{dz_0}{2\pi} \,e^{ixM_Nz_0} \langle N|\bar{\psi}(0)\psi(z)|
413:  N \rangle |_{z_3=-z_0,z_{\perp}=0}.
414:  \label{eq:e0}
415: \end{equation}
416: %
417: Following the general formalism developed
418: in~\cite{DPPPW96},\cite{DPPPW97},\cite{WK99}, the isosinglet
419: distribution in the CQSM is given in the following form :
420: %
421: \begin{eqnarray}
422:  e^{(T=0)}(x) &=& -N_c M_N
423:  \sum_{n > 0}  \,
424:  \langle n \vert \gamma^0  \delta (xM_N - \hat{p}_3 - E_n) 
425:  \vert n \rangle 
426:  \label{eq:e0xunoccu}
427:  \\ 
428:  &=& \ N_c M_N 
429:  \sum_{n \leq 0}  \,
430:  \langle n \vert \gamma^0  \delta (xM_N - \hat{p}_3 - E_n) 
431:  \vert n \rangle ,
432:  \label{eq:e0xoccu}
433: \end{eqnarray}
434: %
435: where, $|n\rangle $ and $E_n$ are the eigenstates and the associated 
436: eigenenergies of the static Dirac Hamiltonian 
437: %
438: \begin{equation}
439:  H = -i
440:  \mbox{\boldmath $\alpha\cdot\nabla $} + \beta M e^{i\gamma_5
441:  \mbox{\boldmath $\tau\!\cdot\!\hat{r}$}F(r)} ,
442:  \label{eq:Dirac.hamil}
443: \end{equation}
444: %
445: with the hedgehog background. Here, the summation $\sum_{n\le 0}$ in
446: (\ref{eq:e0xoccu}) is meant to be taken over the valence-quark orbital
447: (it is the lowest energy eigenstate that emerges from the
448: positive-energy Dirac continuum) plus
449: all the negative-energy Dirac-sea orbitals.
450: On the other hand, the summation $\sum_{n > 0}$ in (\ref{eq:e0xunoccu})
451: is meant to be taken over all the positive-energy Dirac continuum
452: excluding the discrete valence orbital.
453: We recall that the CQSM is defined with some appropriate regularization.
454: In fact, without regularization, $e^{(T=0)}(x)$ is quadratically divergent,
455: and no practical meaning can be given to either
456: of~(\ref{eq:e0xunoccu}) and (\ref{eq:e0xoccu}).
457: The ideal regularization scheme for our purpose is the Pauli-Villars
458: subtraction scheme, since it preserves several fundamental
459: conservation laws of field theory~\cite{DPPPW96},\cite{DPPPW97}.
460: Furthermore, it is also expected to preserve the equivalence of
461: the two ways of computing the quantity in question, by using
462: (\ref{eq:e0xunoccu}) and (\ref{eq:e0xoccu}).
463: In the present study, we use the double subtraction Pauli-Villars scheme
464: as introduced in~\cite{KWW99}, since $e^{(T=0)}(x)$
465: divergees like the vacuum quark condensate.
466: In this scheme the distribution $e^{(T=0)}(x)$ is replaced with a
467: regularized one defined as 
468: %
469: \begin{equation}
470:  e^{(T=0)}(x) \ \equiv \ 
471:  e^{(T=0)}(x)^M - \sum_{i=1}^2 c_i\left( \frac{\Lambda_i}{M}\right)
472:  e^{(T=0)}(x)^{\Lambda_i} .
473:  \label{eq:pv}
474: \end{equation}
475: %
476: Here $e(x)^{\Lambda_i}$ is obtained from $e(x)^M$  by replacing the mass
477: parameter $M$ by $\Lambda_i$. It was shown in~\cite{KWW99}
478: that, if the parameters $c_1, c_2, \Lambda_1$, and 
479: $\Lambda_2$ are chosen to satisfy the two conditions :
480: %
481: \begin{eqnarray}
482:  1 - \sum_{i = 1}^2 \,c_i \,\left(\frac{\Lambda_i}{M} \right)^2 &=& 0 , 
483:  \label{eq:cond1} \\
484:  1 - \sum_{i = 1}^2 \,c_i \,\left(\frac{\Lambda_i}{M} \right)^4 &=& 0 , 
485:  \label{eq:cond2}
486: \end{eqnarray}
487: %
488: the quadratic as well as the logarithmic divergences in the vacuum quark
489: condensate are completely eliminated.
490: 
491: Actually, we are interested in the nucleon observables measured in
492: reference to the physical vacuum, so that $e^{(T=0)}(x)$ should be
493: replaced by
494: %
495: \begin{equation}
496:  e^{(T=0)}(x)\rightarrow e^{(T=0)}(x)\equiv 
497:  e^{(T=0)}_U(x)-e^{(T=0)}_{U=1}(x).
498:  \label{eq:exregl}
499: \end{equation}
500: %
501: Here the vacuum subtraction term $e^{(T=0)}_{U=1}(x)$ is obtained from
502: $e^{(T=0)}_U(x)$ by setting $U=1$ or $F(r)=0$, and by excluding the sum
503: over the discrete valence level.
504: We point out that, due to the energy-momentum conservation
505: embedded in the factor $\delta(x M_N-\hat{p}_3-E_n)$, the vacuum subtraction
506: terms are required only for $x<0$ in the occupied form (\ref{eq:e0xoccu}),
507: and for $x>0$ in the non-occupied form (\ref{eq:e0xunoccu}).
508: This means that the vacuum subtraction terms need not be considered when
509: $e^{(T=0)}(x)$ is evaluated in the following way, i.e. if it is
510: evaluated by using the occupied form for $x\!>\!0$, while by
511: using the nonoccupied form for $x\!<\!0$.
512: 
513: 
514: 
515: \subsubsection{momentum sum rules of $e^{(T=0)}(x)$}
516: 
517: The most important information of the distribution functions are
518: generally contained in their first few moments of lowest orders.
519: This is also the case for the distribution $e^{(T=0)}(x)$.
520: In a recent paper, Efremov and Schweitzer reviewed some of the
521: important sum rules for the chiral-odd twist-3 distribution functions
522: in an enlightening way \cite{ES03}.
523: Their argument starts with the general definition
524: of the distribution with flavor $a$ as
525: %
526: \begin{equation}
527:  e^a(x)=\frac{1}{2M_N}\int\frac{d \lambda}{2\pi} \,e^{i\lambda x}
528:  \langle N|\bar{\psi}_a(0)[0,\lambda n]\psi_a(\lambda n)|N\rangle , 
529:  \label{eq:link}
530: \end{equation}
531: %
532: where $[0,\lambda n]$ denotes the gauge link. By using an operator
533: identity following from the QCD equation of motion, $e^a(x)$ is shown
534: to be decomposed in a gauge invariant way into the following three
535: pieces :
536: %
537: \begin{equation}
538:  e^a(x) = e^a_{sing}(x) + e^a_{tw3}(x) + e^a_{mass}(x).
539:  \label{eq:exdecomp}
540: \end{equation}
541: %
542: Here, $e^a_{sing}(x)$ denotes a singular term given by
543: %
544: \begin{equation}
545:  e^a_{sing}(x) = \delta(x) \,\langle N | \bar{\psi}^a \psi^a |
546:  N \rangle .
547:  \label{eq:exsing}
548: \end{equation}
549: %
550: On the other hand, $e^a_{tw3}(x)$ is a genuine twist-3 part of $e^a(x)$
551: that contains information on quark-gluon-quark correlations.
552: Finally, $e^a_{mass}(x)$ denotes the term arising from the nonzero current
553: quark mass. It is a somewhat peculiar function defined through its
554: Mellin moments as \cite{BBKT96}\nocite{BM97}\nocite{KN97}-\cite{KT99}
555: %
556: \begin{equation}
557:  \int_{-1}^1 \,x^{n-1} \,e^a_{mass}(x) \,dx = \delta_{n>1} \cdot
558:  \frac{m_0}{M_N} \,\int_{-1}^1 \,x^{n-2} \,f^a_1(x) \,dx ,
559:  \label{eq:exmassmom}
560: \end{equation}
561: %
562: with $f^a_1(x)$ being the twist-2 unpolarized distribution with flavor
563: $a$. The presence of the factor $\delta_{n>1}$ here dictates that the
564: first moment of $e^a_{mass}(x)$ vanishes,
565: %
566: \begin{equation}
567:  \int_{-1}^1 \,e^a_{mass}(x) \,dx = 0 .
568:  \label{eq:exmass1stmom}
569: \end{equation}
570: %
571: It is also known  \cite{BBKT96}\nocite{BM97}\nocite{KN97}-\cite{KT99}
572: that the first two basic Mellin moments of $e^a_{t w 3} (x)$
573: vanish, i.e.
574: %
575: \begin{equation}
576:  \int_{-1}^1 x^{n - 1} e^a_{t w 3} (x) d x = 0 \ \ \ \mbox{for} \ \ 
577:  n = 1,2 .
578:  \label{eq:extw3mom}
579: \end{equation}
580: %
581: Putting the above-mentioned properties altogether, the first moment sum 
582: rule for the isoscalar combination of $e^a (x)$, i.e. $e^{(T = 0)} (x)$ 
583: takes the form. 
584: %
585: \begin{equation}
586:  \int_{-1}^1 e^{(T = 0)} (x) d x = \frac{\sum_{\pi N}}{m_0} ,
587:  \label{eq:ex1stmom}
588: \end{equation}
589: %
590: which is nothing but the $\pi N$ sigma-term sum rule. Note that this sum
591: rule is saturated by the first term of (\ref{eq:exdecomp}) alone.
592: On the other hand, the second Mellin moment of $e^{(T = 0)} (x)$ is 
593: given by
594: %
595: \begin{equation}
596:  \int_{-1}^1 x e^{(T = 0)} (x) d x = \frac{m_0}{M_N} \cdot N_c ,
597:  \label{eq:ex2ndmom}
598: \end{equation}
599: %
600: where $N_c$ is the number of color, which coincides with the number of
601: quarks contained in a baryon-number-one system, i.e. $N_c = 3$.
602: We point out that this second Mellin moment of $e^{(T = 0)}$ vanishes
603: in the chiral limit of $m_0 = 0$.
604: 
605: Next, we turn to the discussion of the moment sum rule in the CQSM.
606: Integrating (\ref{eq:e0xoccu}) over $x$, the first moment of
607: $e^{(T = 0)} (x)$ is given as
608: %
609: \begin{equation}
610:  \int_{-1}^1 e^{(T = 0)} (x) d x = N_c \sum_{n \leq 0}
611:  \langle n | \gamma^0 | n \rangle .
612:  \label{eq:ex1stmom-cqsm1}
613: \end{equation}
614: %
615: Since the r.h.s. of this equation is nothing but the scalar charge 
616: $\bar{\sigma}$ of the nucleon within the CQSM, the sigma-term sum rule 
617: immediately follows
618: %
619: \begin{equation}
620:  \int_{-1}^1 e^{(T = 0)} (x) d x = \bar{\sigma} = 
621:  \frac{\sum_{\pi N}}{m_0} .
622:  \label{eq:ex1stmom-cqsm2}
623: \end{equation}
624: %
625: The way of this sum rule being satisfied is far more delicate in the CQSM
626: than in the above QCD-motivated analysis. As shown by our previous study,
627: although the model certainly predicts the $\delta (x)$-type singularity in
628: $e^{(T = 0)} (x)$, this term alone does not saturate the $\pi N$
629: sigma-term sum rule. The model also predicts nontrivial structure of
630: $e^{(T = 0)} (x)$ at $x \neq 0$, which may contribute to the first
631: moment sum rule. We shall discuss this point in more detail
632: in the next section.
633: 
634: Turning to the second moment, it is easy to show from (\ref{eq:e0xoccu})
635: that
636: %
637: \begin{equation}
638:  \int_{-1}^1 x e^{(T = 0)} (x) d x = \frac{N_c}{M_N}
639:  \sum_{n \leq 0} \langle n | \gamma^0 (\hat{p}_3 + E_n ) | n \rangle .
640:  \label{eq:ex2ndmom-cqsm1}
641: \end{equation}
642: %
643: Owing to the hedgehog symmetry of the soliton, the term containing 
644: $\gamma^0 \hat{p}_3$ vanishes, and we are left with
645: %
646: \begin{equation}
647:  \int_{-1}^1 x e^{(T = 0)} (x) d x = \frac{N_c}{M_N}
648:  \sum_{n \leq 0} E_n \langle n | \gamma^0 | n \rangle .
649:  \label{eq:ex2ndmom-cqsm2}
650: \end{equation}
651: %
652: Following \cite{S03}, it is convenient to rewrite the r.h.s. of the
653: above equation in the following manner.
654: First, notice the identity
655: %
656: \begin{equation}
657:  E_n \langle n | \gamma^0 | n \rangle
658:  = \frac{1}{2} \langle n | \{ \hat{H}, \gamma^0 \} | n \rangle
659:  = m_0 + M \langle n | \frac{1}{2} (U + U^{\dagger}) | n \rangle .
660:  \label{eq:anticom}
661: \end{equation}
662: %
663: Here, we have tentatively restored the current quark mass term in the 
664: model Hamiltonian $H$, just for the sake of explanation here only, i.e.
665: we have used here
666: \begin{equation}
667:  H = -i \mbox{\boldmath $\alpha$} \cdot \nabla 
668:  + \beta M e^{i \gamma_5 \mbox{\boldmath $\tau$} \cdot 
669:  \hat{\mbox{\boldmath $r$}} F(r)} 
670:  + m_0 .
671:  \label{eq:dirac.hamil.m}
672: \end{equation}
673: %
674: Then, the second moment sum rule in the CQSM takes the following form :
675: \begin{equation}
676:  \int_{-1}^1 x e^{(T = 0)} (x) d x = \frac{N_c}{M_N} \,(m_0 + \beta M) ,
677:  \label{eq:ex2ndmom-cqsm3}
678: \end{equation}
679: %
680: with
681: %
682: \begin{equation}
683:  \beta \equiv \sum_{n \leq 0} \langle n | \frac{1}{2} (U + U^{\dagger})
684:  | n \rangle .
685:  \label{eq:beta}
686: \end{equation}
687: %
688: It is clear now that r.h.s. of this sum rule does not vanish even in the
689: chiral limit of $m_0 = 0$, in contrary to the sum rule derived from the
690: QCD-equation-of-motion method.
691: We shall return to this point in the next section.
692: 
693: 
694: 
695: \subsection{Isovector distribution \mbox{\boldmath $e^{(T=1)}(x)$}}
696: 
697: The isovector distribution is defined by
698: %
699: \begin{equation}
700:  e^{(T=1)}(x) \equiv e^u(x)-e^d(x)=M_N\int^{\infty}_{-\infty}
701:  \frac{dz_0}{2\pi} \,e^{ixM_Nz_0} \langle N|\bar{\psi}(0)\tau_3\psi(z)|
702:  N \rangle |_{z_3=-z_0,z_{\perp}=0} .
703:  \label{eq:e1}
704: \end{equation}
705: %
706: Within the framework of the CQSM,  $e^{(T=1)}(x)$ survives
707: only in the next-to-leading order in the collective angular velocity
708: $\Omega$. Following the formalism derived in~\cite{WK99},\cite{PPGWW99},
709: the final answer is written in the form :
710: \begin{eqnarray}
711:  e^{(T=1)}(x)&=& -\langle 
712:  2T_3\rangle_p M_N \frac{N_c}{2I}\frac13\sum_{a=1}^3
713:    \sum_{m = all, n > 0}  
714:  \langle n | \tau_a | m \rangle
715:  \langle m | \tau_a  \gamma^0  
716:  \left( \frac{\delta_n}{E_m - E_n} - \frac{1}{2M_N} \,
717:  \delta_n^\prime \right)
718:  | n \rangle \nonumber\\
719:  &=&\langle 
720:  2T_3\rangle_p M_N \frac{N_c}{2I}\frac13\sum_{a=1}^3
721:  \sum_{m = all, n \leq 0}
722:  \langle n | \tau_a | m \rangle
723:  \langle m | \tau_a  \gamma^0  
724:  \left( \frac{\delta_n}{E_m - E_n} - \frac{1}{2M_N} \,
725:  \delta_n^\prime \right)
726:  | n \rangle , \ \ 
727:  \label{eq:dis31}
728: \end{eqnarray}
729: %
730: with $\delta_n\equiv \delta(x M_N-E_n-p^3)$ and 
731: $\delta'_n=\frac{\partial}{\partial x}\delta(x M_N-E_n-p^3)$.
732: Here $I$ in the r.h.s. of (\ref{eq:dis31}) is the moment of inertia
733: of the soliton, given by
734: %
735: \begin{equation}
736:  I = \frac{N_c}{6}\sum_{a=1}^3 \sum_{m > 0} \sum_{n \le 0}
737:  \frac{\langle n |\tau_a |m \rangle \langle m |\tau_a | 
738:  n \rangle}{E_m-E_n}.
739:  \label{eq:momentinertia}
740: \end{equation}
741: %
742: In (\ref{eq:dis31}), $\langle O\rangle_p$
743: should be understood as an abbreviated notation of the matrix element
744: of a collective operator $O$ between the (spin-up) proton state, i.e.
745: %
746: \begin{eqnarray}
747:  {\langle O \rangle}_p &\equiv& \int
748:  \Psi_{T = T_3 = 1/2 ; J = J_3 = 1/2}^{*} [\xi_A] \,O [\xi_A] \,
749:  \Psi_{T = T_3 = 1/2 ; J = J_3 = 1/2} [\xi_A] \,d \xi_A \nonumber \\
750:  &=& \ \ \langle p,S_3=1/2|O|p,S_3=1/2\rangle .
751:  \label{eq:matrcollect}
752: \end{eqnarray}
753: %
754: In the present case, we have ${\langle 2T_3 \rangle}_p = 1$.
755: %As for the isosinglet distribution, the vacuum 
756: %subtruction is implied in Eq.(\ref{eq:dis31}).
757: 
758: We immediately notice that the above expressions are not suitable for
759: the actual numerical calculation. Here, we shall proceed as in the
760: previous studies~\cite{WK99},\cite{PPGWW99}. First, note that the term
761: containing the $x$-derivative of the $\delta$-function
762: in (\ref{eq:dis31}) can be rewritten as
763: %
764: \begin{equation}
765:  e_2(x)=-
766:  \frac{d}{dx} \frac{N_c}{4I}\frac13\sum_a
767:  \sum_{m = all, n \leq 0}  
768:  \langle n | \tau_a | m \rangle
769:  \langle m | \tau_a  \gamma^0 
770:  \delta_n 
771:  | n \rangle  = 
772:  \frac{1}{4IM_N}\frac{d}{dx}\,e^{(T=0)}(x).
773:  \label{eq:de0}
774: \end{equation}
775: %
776: Here, we have made of the completeness of the eigenstates
777: $| n \rangle$ of the static Dirac Hamiltonian $H$.
778: (We recall that $e_2(x)$ term originates from the non-locality in time
779: of the operator $\bar{\psi}(0)\tau^a\psi(z)$ in (\ref{eq:e1}).)
780: It should be recognized that the $x$-derivative of the isosinglet
781: distribution $e^{(T=0)}(x)$ appears in the right-hand side.
782: Since we already know that the isosinglet distribution $e^{(T=0)}(x)$
783: has the $\delta(x)$ type singularity connected with the nonvanishing
784: vacuum expectation, it then follows that $e_2(x)$ has the
785: derivative-of-$\delta(x)$ type singularity.
786: However, it is unlikely that the net isovector distribution
787: $e^{(T=1)}(x)$ has such a singularity, because the QCD vacuum should
788: not violate isospin symmetry so that vacuum quark condensate of
789: isovector type must simply vanish.
790: This apparent discrepancy can be resolved as follows.
791: We first divide the double sum of (\ref{eq:dis31}) into the sum over
792: terms with $E_m=E_n$ and with $E_m\ne E_n$.
793: The point is that the sum with $E_m=E_n$ in $e_1(x)$ can be rewritten
794: in a similar form as the corresponding term in $e_2(x)$,
795: %
796: \begin{eqnarray}
797:  e_1(x)&=&
798:  M_N \frac{N_c}{2I}\frac13\sum_a
799:  \sum_{{m = all, n \leq 0 \atop (E_m \ne E_n)}}\frac{1}{E_m-E_n}  
800:  \langle n | \tau_a | m \rangle
801:  \langle m | \tau_a  \gamma^0  
802:  \delta_n 
803:  | n \rangle
804:  \nonumber\\
805:  &&+
806:  \frac{d}{dx}\frac{N_c}{4I}\frac13\sum_a
807:  \sum_{{m\leq 0,n\leq0\atop (E_m=E_n)}}
808:  \langle n | \tau_a | m \rangle
809:  \langle m | \tau_a  \gamma^0  
810:  \delta_n
811:  | n\rangle ,
812:  \label{eq:dble1}\\
813:  e_2(x)&=&- 
814:  \frac{d}{dx}\frac{N_c}{4I}\frac13\sum_a
815:  \sum_{{m = all,n\leq0\atop (E_m \ne E_n)}}
816:  \langle n | \tau_a | m \rangle
817:  \langle m | \tau_a  \gamma^0  
818:  \delta_n
819:  | n\rangle
820:  \nonumber\\
821:  &&- 
822:  \frac{d}{dx}\frac{N_c}{4I}\frac13\sum_a
823:  \sum_{{m\leq 0,n\leq0\atop (E_m=E_n)}}
824:  \langle n | \tau_a | m \rangle
825:  \langle m | \tau_a  \gamma^0  
826:  \delta_n
827:  | n\rangle .
828:  \label{eq:dblsum}
829: \end{eqnarray}
830: %
831: Now, just as argued in~\cite{PPGWW99},\cite{WK99}, $E_m = E_n$
832: contribution in the double sums in $e_1(x)$ and $e_2(x)$ precisely
833: cancel each other.
834: After regrouping the terms in such a way that this cancellation occurs at
835: the level of analytical expressions, the ${\cal O}(\Omega^1)$
836: contribution to the distribution function $e^{(T=1)}(x)=e^u(x)-e^d(x)$
837: can finally be written in the following form:
838: %
839: \begin{equation}
840:  e^{(T=1)}(x)=
841:  M_N \frac{N_c}{2I}\frac13\sum_a
842:  \sum_{{m = all, n \leq 0\atop(E_m \ne E_n)}}  
843:  \langle n | \tau_a | m \rangle
844:  \langle m | \tau_a  \gamma^0  
845:  \left( \frac{\delta_n}{E_m - E_n} - \frac{1}{2M_N} \,
846:  \delta_n^\prime \right)
847:  | n \rangle .
848:  \label{eq:nosing}
849: \end{equation}
850: %
851: The fact is that, in the double sum of (\ref{eq:dblsum}),
852: the singularity connected with the non-zero vacuum quark condensate
853: comes only from $E_m = E_n$ contribution, i.e. the second term of
854: (\ref{eq:dblsum}). As mentioned above, after the $E_m = E_n$
855: contributions in $e_1(x)$ and $e_2(x)$ are canceled, these singularities
856: disappear in (\ref{eq:nosing}). The final theoretical formula
857: (\ref{eq:nosing}) is therefore free from any
858: singularity which contradicts the symmetries of the QCD vacuum, and
859: it provides us with a sound basis for numerical calculation.
860: 
861: 
862: 
863: \subsubsection{first moment sum rule of $e^{(T=1)}(x)$}
864: \label{sum-rule-e1}
865: 
866:  Here we discuss the first moment sum rule of the isovector
867: distribution. Integrating (\ref{eq:e1}) over $x$, we obtain
868: %
869: \begin{equation}
870:  \int_{-1}^1 e^{(T=1)}(x) dx = 
871:  \int_{-1}^1 (e^u(x)-e^d(x))dx=\langle N|\bar{\psi}
872:  \tau_3\psi|N\rangle \, .
873:  \label{eq:dis12}
874: \end{equation}
875: %
876: (Here, $\bar{\psi} \tau_3\psi$ should be taken as an
877: abbreviated notation of $\int \bar{\psi}(\mbox{\boldmath $y$}) 
878: \tau_3\psi (\mbox{\boldmath $y$}) \,d^3 \mbox{\boldmath $y$}$,
879: which gives the isovector scalar charge operator.)
880: An interesting observation is that the first moment of $e^{( T=1)}(x)$ is 
881: related to the non-electromagnetic mass difference of neutron and proton.
882: In fact, the nonelectromagnetic neutron-proton mass difference is
883: thought to be generated by the isospin breaking term in the QCD
884: Hamiltonian : 
885: %
886: \begin{equation}
887:  \Delta H=\frac{m_u-m_d}{2} \,(\bar{\psi}_u\psi_u-\bar{\psi}_d\psi_d)\, .
888:  \label{eq:chiralsb}
889: \end{equation}
890: % 
891: Because of the smallness of all the masses $m_u,m_d,m_d-m_u$ compared
892: with the typical energy-scale of hadron physics, we can treat $\Delta H$
893: as the first order perturbation, thereby being led to the following
894: formula for the non-electromagnetic mass difference between neutron
895: and proton :
896: %
897: \begin{eqnarray}
898:  (M_n-M_p)_{QCD}&=&\langle n |\Delta H|n\rangle - 
899:  \langle p|\Delta H |p\rangle
900:  \nonumber\\
901:  &=& (m_d-m_u)\langle p|\bar{\psi}_u\psi_u-\bar{\psi}_d 
902:  \psi_d|p\rangle ,
903:  \label{eq:npmdif}
904: \end{eqnarray}
905: %
906: where use has been made of the isospin symmetry for the unperturbative
907: state $|p\rangle ,|n\rangle $ (i.e., the invariance under the
908: interchanges $p\leftrightarrow n$ and $u\leftrightarrow d$).
909: Empirically, the neutron-proton mass difference of QCD origin can be
910: estimated from the observed mass difference by taking account of
911: the correction due to the electromagnetic interactions :
912: %
913: \begin{equation}
914:  (M_n-M_p)_{QCD}=(M_n-M_p)_{exp}-(M_n-M_p)_{e.m.}\, .
915:  \label{eq:npmdifcor}
916: \end{equation}
917: %
918: Using the values $(M_n-M_p)_{exp}\simeq 1.29 \,\mbox{MeV},\,
919: (M_n-M_p)_{e.m.}\simeq (-0.76\pm 0.30)\mbox{MeV}$~\cite{GL82},
920: we obtain 
921: %
922: \begin{equation}
923:  (M_n-M_p)_{QCD}\simeq (2.05\pm 0.30) \,\mbox{MeV}\, ,
924:  \label{eq:npmdifqcd}
925: \end{equation}
926: %
927: To extract the first moment of $e^{(T=1)}(x)$ empirically, we
928: need to know the value of $m_d - m_u$. By using
929: $m_d - m_u \simeq 5 \,\mbox{MeV}$, as an order of magnitude
930: estimate, we obtain
931: %
932: \begin{eqnarray}
933:  \int_{-1}^1 e^{(T=1)}(x) \,dx &=& 
934:  \frac{(M_n-M_p)_{QCD}}{m_d-m_u}\simeq 0.41\pm 0.06\, .
935:  \label{eq:dis13}
936: \end{eqnarray}
937: 
938: On the other hand, the theoretical expression for the first moment of
939: $e^{(T=1)} (x)$ is obtained from (\ref{eq:nosing}) as
940: %
941: \begin{equation}
942:  \int^1_{-1} e^{(T=1)}(x) dx = \frac{N_c}{2I}\frac13
943:  \sum_a \sum_{n\le 0}\sum_{m>0}\frac{\langle n |\tau_a |m\rangle
944:  \langle m | \tau_a \gamma^0|n\rangle }{E_m-E_n} \, ,
945:  \label{eq:modelsum}
946: \end{equation}
947: %
948: Here, we have used the fact that, since the contribution $e_2(x)$
949: is a total derivative, it does not contribute
950: to the integral of (\ref{eq:modelsum}). 
951: After integration over $x$, the double sum over levels in 
952: (\ref{eq:modelsum}) is naturally restricted to include only
953: transitions from occupied to non-occupied states.
954: This is reasonable, since
955: the operator appearing in the r.h.s. of (\ref{eq:modelsum})
956: is a local operator, and transitions from occupied to occupied states
957: would violate the Pauli principle.
958: Within the framework of the CQSM, we can evaluate the
959: r.h.s. of (\ref{eq:dis12}), i.e. the isovector scalar charge of the
960: nucleon $\langle N|\bar{\psi}\tau^3\psi|N\rangle $, directly
961: without passing through the distribution function.
962: Since the resultant expression of $\langle N|\bar{\psi}\tau^3\psi|N
963: \rangle $ precisely coincides with the r.h.s. of (\ref{eq:modelsum}),
964: we conclude that the first moment sum rule of $e^{(T=1)}(x)$ is
965: properly satisfied within the model.
966: 
967: 
968: 
969: \section{Numerical results and discussion}
970: 
971: The numerical method used for evaluating $e(x)$ in this paper is
972: essentially the same as the one used for the computing the
973: twist-2 distributions $q(x),\, \Delta q(x),\,
974: \delta q(x)$~\cite{WK98,WK99}.
975: The eigen-energies and the eigen-vectors of the static Dirac
976: Hamiltonian $H$ with the hedgehog background are obtained by diagonalizing
977: it with the so-called Kahana-Ripka plane-wave basis~\cite{KR84}.
978: Following them, the plane-wave states, introduced as a set of
979: eigenstates of the free Hamiltonian 
980: $H_0 = - i \mbox{\boldmath $\alpha$}\cdot\! \nabla\! + \beta M$, 
981: is discretized by imposing an appropriate boundary condition for the 
982: radial wave functions at the radius $D$ chosen to be sufficiently larger
983: than the soliton size. The basis is made finite by
984: retaining only those states with the momentum $k$ satisfying
985: the condition $k<k_{max}$. As a results of using
986: this discretized momentum basis, a resultant distribution
987: becomes a discontinuous function of $x$, due to the factor
988: $\delta(x M_n-E_n-\hat{p}_3)$.
989: In order to get a continuous function with a discretized basis,
990: we introduce a smeared distribution function in the variable $x$
991: as \cite{DPPPW97}
992: %
993: \begin{equation}
994:  e_{\gamma}(x)\equiv \frac{1}{\gamma \sqrt{\pi}}
995:  \int_{-\infty}^{\infty}
996:  e^{-\frac{(x-x')^2}{\gamma^2}}e(x') dx',
997:  \label{eq:smeared}
998: \end{equation}
999: %
1000: with a small but finite value of $\gamma$ ($\gamma\ll 1$).
1001: The smeared distribution is expected to be continuous when
1002: the separation between the discretized momenta is much smaller
1003: than the smearing width $\gamma$. Since the physical distribution
1004: corresponds to the limit $\gamma \rightarrow 0$, this 
1005: forces us to employ a very large box size $D$ to get a continuous
1006: distribution function.
1007: 
1008: This procedure works very well at least for the standard distributions
1009: investigated so far.
1010: However, in the numerical calculation of $e^{(T=0)}(x)$, we
1011: have a new problem which we have not encountered before.
1012: Our expectation is that, if a $\delta(x)$-type singularity really
1013: exists in $e^{(T=0)}(x)$, the corresponding smeared distribution
1014: would have a Gaussian peak centered around $x=0$ with the
1015: width $\gamma$.
1016: The problem here is that the distribution function in question
1017: may also have a piece that is non-singular for all value of $x$.
1018: One might think that the contribution of the singular part can
1019: be disentangled from the total contribution by using the
1020: ``unsmearing method" described in \cite{DPPPW97}.
1021: This is not feasible, however, by the following reasons.
1022: First, although the smearing procedure defined with
1023: (\ref{eq:smeared}) preserves the integral value of the
1024: distribution, we have no {\it ad hoc} way to know the overall
1025: coefficient of the $\delta(x)$ term of the
1026: distribution. Secondly, the small $x$ behavior of the nonsingular part of
1027: the distribution would be hard to know, because it is buried
1028: in the very large contribution of smeared $\delta$ function
1029: singularity. This point will be discussed in more detail in the
1030: following subsection.
1031: 
1032: 
1033: 
1034: \subsection{isosinglet distribution $\mbox{\boldmath $e^{(T=0)}(x)$}$}
1035: 
1036: \ \ \ In the numerical calculation, we fix the pion weak decay constant
1037: $f_{\pi}$ in (\ref{eq:hedgehog})
1038: to its physical value, i.e., $f_{\pi}=93\,$MeV, so that only one
1039: parameter of the model is the dynamical quark mass $M$, which
1040: plays the role of the coupling constant between the pion and the 
1041: effective quark fields. 
1042: Through the present analysis, we use the value of $M=375 \,\mbox{MeV}$,
1043: which is favored from the phenomenology of nucleon low energy
1044: observables.
1045: With $M=375 \,\mbox{MeV}$, we have $\Lambda_1\simeq 627 \,\mbox{MeV}$
1046: and $\Lambda_2\simeq 1589 \,\mbox{MeV}$ from the
1047: conditions (\ref{eq:cond1}), (\ref{eq:cond2}).
1048: The static soliton energy
1049: obtained with these parameters is about $1018 \,\mbox{MeV}$.
1050: We point out that, although the soliton mass emerges about $8 \,\%$
1051: larger than the observed nucleon mass $M_N$, the consistency with
1052: the energy-momentum sum rule of the unpolarized distribution functions
1053: enforces us to use this value for $M_N$ in the following evaluation
1054: of the distribution functions.
1055: 
1056: We start with showing the numerical equivalence of the final answers
1057: based on the non-occupied representation and the occupied one.
1058: The problem here is the dependence on the cut-off momentum $k_{max}$,
1059: which is introduced to make finite the discretized Kahana-Ripka basis set.
1060: Since the distribution $e^{(T = 0)} (x)$ is ultra-violet finite after 
1061: introduction of the double-subtraction Pauli-Villars regularization, 
1062: one might expect that the answers would be stable as far as one takes 
1063: $k_{max}$ much larger than the second Pauli-Villars cut-off mass 
1064: $\Lambda_2 \simeq 1.6 \,\mbox{GeV}$.
1065: This is not the case, however. As clarified in \cite{WO03}, the
1066: $\delta$-function type singularity in 
1067: $e^{(T = 0)} (x)$ is generated by the contribution of the infinitely 
1068: deep Dirac-sea levels, which are naturally contained in either of the 
1069: three terms, i.e. the main term and the two Pauli-Villars subtraction 
1070: terms. This implies that the singularity, which will appear in the
1071: smeared distribution as a Gaussian peak around $x = 0$ with the width 
1072: $\gamma$, would be reproduced only in the ideal limit of 
1073: $k_{max} \rightarrow \infty$.
1074: To achieve this ideal limit, we therefore use an extrapolation method
1075: explained below. For this extrapolation to be done smoothly, 
1076: we first introduce an energy cut-off into the level
1077: sums~(\ref{eq:e0xunoccu}) and (\ref{eq:e0xoccu}) of the form.
1078: %
1079: \begin{eqnarray}
1080:  [e^u(x)+e^d(x)] _{non-occupied}^R &=& -N_c M_N
1081:  \sum_{n > 0}  \,
1082:  \langle n \vert \gamma^0  \delta (xM_N - \hat{p}_3 - E_n) \vert n \rangle
1083:  R(E_n) ,
1084:  \label{eq:e0-nonoccu}
1085:  \\ 
1086:  {}[e^u(x)+e^d(x)] _{occupied}^R
1087:  &=& \ N_c M_N 
1088:  \sum_{n \leq 0}  \,
1089:  \langle n \vert \gamma^0  \delta (xM_N - \hat{p}^3 - E_n) \vert n \rangle
1090:  R(E_n) .
1091:  \label{eq:e0-occu}
1092: \end{eqnarray}
1093: %
1094: Here, $R(E_n)$ is a smooth regulator function with an energy cut-off, 
1095: $E_{max} = \sqrt{k_{max}^2 + M^2}$. For this regulator function, 
1096: we employ here a Gauusian function
1097: %
1098: \begin{equation}
1099:  R(E_n) = \mbox{exp} \left[-{\left(E_n / E_{max}\right)}^2\right] ,
1100:  \label{eq:reglfunc}
1101: \end{equation}
1102: %
1103: following Diakonov et al.~\cite{DPPPW97}.
1104: We first compute the level sums (\ref{eq:e0-nonoccu})
1105: and (\ref{eq:e0-occu}) for several values
1106: of $k_{max}$, in the case of masses $M$, $\Lambda_1$,
1107: and $\Lambda_2$ respectively, and then perform 
1108: the Pauli-Villars subtraction, and finally remove the energy cut-off 
1109: by the numerical extrapolation to infinity pointwise in $x$.
1110: In the present study, we use five data (corresponding to 
1111: $k_{max} / M = 12, 16, 20, 24, \mbox{and} \ 28)$ and perform a least
1112: square fit of these data by using a fourth order function of $1/k_{max}$.
1113: %
1114: \begin{figure}[htb] \centering
1115: \begin{center}
1116:  \includegraphics[width=11.0cm]{occupied.extrp.eps}
1117: \end{center}
1118: \vspace*{-0.5cm}
1119: \renewcommand{\baselinestretch}{1.20}
1120: \caption{The $k_{max}$ dependence of the Dirac-sea contribution
1121: to $e^{(T=0)}(x)$ based on the occupied representation.
1122: The solid curve represents the extrapolated result.} 
1123: \label{fig:e0(x).occupied}
1124: \end{figure}
1125: %
1126: 
1127: Now we are ready to show in Fig.{\ref{fig:e0(x).occupied}
1128: the $k_{max}$ dependence of the 
1129: Dirac-sea contributions based on the occupied representation for all 
1130: values of $x$. Here we use a value of $\gamma = 0.1$.
1131: This figure shows that the peak positions of the Gaussian-like
1132: function obtained with the finite cutoff energy deviate to the
1133: negative $x$ region from the origin $x = 0$.
1134: This deviation of the peak 
1135: position in the smeared distribution may be understood as follows. 
1136: First, when one uses the occupied representation, the vacuum substraction 
1137: as represented by (\ref{eq:exregl}) is necessary only for the region
1138: $x < 0$, while it is not necessary for $x > 0$, since the vacuum term
1139: identically vanishes for $x > 0$ due to the restriction of the factor
1140: $\delta(x M_N - E_n - \hat{p}_3)$.
1141: Secondly, we recall the fact that the singular term of $e^{(T = 0)} (x)$ 
1142: emerges as a delicate cancellation of two large numbers or the 
1143: infinities, i.e. the difference between the main contribution with 
1144: hedgehog background and the vacuum subtraction term obtained with $U = 1$.
1145: These two facts indicate that the use of the occupied form with
1146: some finite value of $\gamma$ can reproduce 
1147: the redistribution of the delta-function strength at $x = 0$ in the 
1148: $x < 0$ region only, but it cannot do it properly in the $x > 0$ region, 
1149: as far as the finite energy cutoff is used. This is the reason why the 
1150: Gaussian-like peak of the smeared distribution is shifted to the negative 
1151: $x$ region. One can however confirm the behavior that the position of the 
1152: Gaussian peak approaches $x = 0$ as the energy cutoff is 
1153: increased. And, finally, with the extrapolation method, we obtain a
1154: reasonable result which shows that the peak of the smeared distribution
1155: is positioned just at $x= 0$.
1156: (In the above analysis, we fix the box size to
1157: be $DM = 20$. As a matter of course, to get a physically acceptable
1158: answers, we must also investigate the dependence of the answers on
1159: the box size $D$. We found that, above 
1160: $DM = 20$, the change of the small $x$ behavior of $e^{(T = 0)} (x)$ as
1161: illustrated in Fig.{\ref{fig:e0(x).occupied} is almost due to
1162: the increase of $k_{max}$, and the answer is stable against the
1163: further increase of $DM$ above 20.)
1164: %
1165: \begin{figure}[htb] \centering
1166: \begin{center}
1167:  \includegraphics[width=11.0cm]{occ.and.nonocc.eps}
1168: \end{center}
1169: \vspace*{-0.5cm}
1170: \renewcommand{\baselinestretch}{1.20}
1171: \caption{Comparison of the Dirac-sea contributions to
1172: $e^{(T=0)}(x)$ based on the occupied (solid curve) and
1173: non-occupied (dashed curve) representations.}
1174: \label{fig:occu.and.non-occu}
1175: \end{figure}
1176: %
1177: After carrying out the similar analysis, this time, with use of the
1178: non-occupied representation, we can now compare the final numerical
1179: results for the Dirac-sea contribution obtained with the two
1180: alternative representations.
1181: Fig.{\ref{fig:occu.and.non-occu} shows this comparison. A reasonable 
1182: agreement between the two ways of evaluating $e^{(T = 0)} (x)$ confirms 
1183: the equivalence of the two representations. At the same time, the
1184: analysis above establishes the existence of the $\delta$-function
1185: singularity in $e^u (x) + e^d (x)$ on the numerical ground.
1186: Some difference between the two curves at the positive and negative $x$
1187: tails of the Gaussian like distributions would be a spurious one
1188: generated by the numerical extrapolation method.
1189: The contributions based on the occupied representation for $x<0$ and
1190: the non-occupied representation for $x>0$ can be obtained after the 
1191: cancellation of two large numbers, i.e., the main contribution and the 
1192: corresponding vacuum subtraction term.
1193: On the other hand, if one uses the occupied representation for $x>0$
1194: and the non-occupied representation for $x<0$, one is free from
1195: the spurious contribution due to the cancellation,
1196: so that the extrapolated curves at these tail region have reasonable
1197: smooth behavior.
1198: 
1199: Although we were able to confirm the existence of 
1200: $\delta (x)$-type singularity in the numerical analysis of 
1201: $e^{(T = 0)} (x)$, we cannot exclude the possibility that the 
1202: $e^{(T = 0)} (x)$ may also contain a regular term which is smooth in all
1203: the range of $x$. Is it possible to disentangle such non-singular term of 
1204: $e^{(T = 0)} (x)$ from the total contribution containing the singular one?
1205: One should recognize that it is not so easy by the following reasons.
1206: First, the deconvolution method as proposed by Diakonov et al. does not 
1207: work because of the very delicate nature of the
1208: singularity~\cite{DPPPW97}.
1209: Second, we have no {\it ad hoc} way to know the coefficient of
1210: $\delta (x)$ term in the original unsmeared distribution.
1211: Nevertheless, we found that the following trick works for obtaining 
1212: the non-singular distribution excluding the $\delta (x)$ term.
1213: That is, as repeatedly emphasized, by using the non-occupied expression 
1214: for $x < 0$ and the occupied one for $x > 0$, we can avoid the vacuum 
1215: subtraction. Interestingly, this also works to remove the singular
1216: contribution in the bare distribution, and the corresponding smeared
1217: distribution would not contain the Gaussian peak corresponding to
1218: $\delta (x)$-type singularity. (One should remember the fact that the
1219: vacuum term plays an indispensable role in reproducing the
1220: $\delta$-function singularity in $e^{(T=0)}(x)$.)
1221: 
1222: 
1223: %
1224: \begin{figure}[htb] \centering
1225: \begin{center}
1226:  \includegraphics[width=11.0cm]{reg.extrp.eps}
1227: \end{center}
1228: \vspace*{-0.5cm}
1229: \renewcommand{\baselinestretch}{1.20}
1230: \caption{The $k_{max}$ dependence of  
1231: the Dirac-sea contributions to $e^{(T=0)}(x)$
1232: based on the occupied representation for $x>0$ and the non-occupied 
1233: representation for $x<0$. The solid curves represents
1234: the extrapolated result.}
1235: \label{fig:reg.extrp}
1236: \end{figure}
1237: 
1238: Fig.~\ref{fig:reg.extrp} shows the $k_{max}$ dependence 
1239: of the Dirac-sea contributions  based on the occupied 
1240: representation for $x>0$ and the non-occupied representation for $x<0$.
1241: One finds that the large and positive Gaussian peak, the reminiscence
1242: of the $\delta$-function singularity in the bare distribution,
1243: does not appear any more.
1244: One can also see that the negative large contributions of the Dirac sea
1245: in the small $x$ region tend to decrease as the cutoff momentum
1246: $k_{max}$ increase.
1247: We again remove the energy cutoff by the numerical extrapolation to
1248: infinity pointwise in $x$. We observe some difference from the
1249: previous case, however. Owing to the feature that the
1250: $\delta$-function singularity is already excluded in the present way
1251: of calculation, the $k_{max}$ dependence is well reproduced by the
1252: linear function of $1 / k_{max}$ as illustrated in
1253: Fig.~\ref{fig:fit.point}.
1254: After this extrapolation procedure, the result
1255: shows a smooth behavior in the whole region of $x$ except the region
1256: $|x|<0.06$ in which the answer is thought to contain some numerical
1257: instability generated by the extrapolation method.  
1258: Neglecting the data in the $|x|<0.06$ region, we make this
1259: extrapolated result smooth.
1260: After deconvoluting the smeared distribution with use of the
1261: Fourier and its inverse transforms, we obtain the final
1262: prediction for the distribution  $e^{(T=0)}(x)$ within the framework
1263: of the CQSM, the normalization point of which may be interpreted as
1264: about $600 \,\mbox{MeV}$.
1265: %
1266: \begin{figure}[htb] \centering
1267: \begin{center}
1268:  \includegraphics[width=11.0cm]{fit.point.eps}
1269: \end{center}
1270: \vspace*{-0.5cm}
1271: \renewcommand{\baselinestretch}{1.20}
1272: \caption{The $k_{max}$ dependence of $e^{(T=0)}_{sea}(x)$ at
1273: $x=0.12$ and its linear extrapolation to $k_{max} \rightarrow \infty$.}
1274: \label{fig:fit.point}
1275: \end{figure}
1276: %
1277: 
1278: Summarizing our analysis up to this point, the isosinglet part of the
1279: chiral-odd twist-3 distribution is given as a sum of the valence quark
1280: and Dirac-sea quark contributions,
1281: %
1282: \begin{equation}
1283:  e^{(T = 0)} (x) = e^{(T=0)}_{val} (x) + e^{(T=0)}_{sea} (x),
1284:  \label{eq:exvalsea}
1285: \end{equation}
1286: %
1287: where the Dirac-sea contribution consists of the singular term and the 
1288: nonsingular (regular) term as 
1289: %
1290: \begin{equation}
1291:  e^{(T=0)}_{sea} (x) = C \delta (x) + e^{(T=0)}_{reg} (x),
1292:  \label{eq:exsea}
1293: \end{equation}
1294: %
1295: \begin{figure}[htb] \centering
1296: \begin{center}
1297:  \includegraphics[width=11.0cm]{e0.t.vl.vc.eps}
1298: \end{center}
1299: \vspace*{-0.5cm}
1300: \renewcommand{\baselinestretch}{1.20}
1301: \caption{The final theoretical predictions of the CQSM for $e^{(T=0)}(x)$.
1302: The dot-dashed curve represents the contribution
1303: of $N_c$ valence level quarks, the dashed curve the
1304: non-singular part of the Dirac-sea contributions, and the solid curve
1305: does their sum. The $\delta$-function singularity at $x=0$ in
1306: the Dirac-sea quarks part is not shown in this figure.}
1307: \label{fig:e0(x)}
1308: \end{figure}
1309: 
1310: Shown in Fig.\ref{fig:e0(x)} are the final theoretical predictions for 
1311: $e^{(T = 0)} (x)$ obtained in the above-explained way.
1312: The dashed curve here represents the contribution of 
1313: $N_c$ valence level quarks, while the dotted curve does the regular part
1314: of Dirac-sea contribution. The sum of these two contributions is shown
1315: by the solid curve.
1316: (We recall that the $\delta (x)$-type singular term is not 
1317: shown in this figure.) One can convince that the regular part of the 
1318: Dirac-sea contribution shows a nontrivial structure in the $x \neq 0$ 
1319: region.
1320: 
1321: After performing the numerical integration of the above distributions
1322: over $x$, one can obtain the contributions of the valence quark term
1323: and the regular part of the Dirac-sea term to the first moment sum rule :
1324: %
1325: \begin{eqnarray}
1326:  \int_{-1}^1 e^{(T=0)}_{val} (x) d x &\simeq& 1.7, \\
1327:  \label{eq:exmomval}
1328:  \int_{-1}^1 e^{(T=0)}_{reg} (x) d x &\simeq& 0.18.
1329:  \label{eq:exmomsea}
1330: \end{eqnarray}
1331: %
1332: Note that the regular part of $e^{(T=0)}_{sea} (x)$ gives small but
1333: non-zero contribution to the sum rule.
1334: To determine the coefficient of the singular 
1335: term in (\ref{eq:exsea}), we use the first moment sum rule
1336: (\ref{eq:ex1stmom-cqsm1}) or (\ref{eq:ex1stmom-cqsm2}) for 
1337: $e^{(T = 0)} (x)$, which was already shown to hold within the framework 
1338: of the CQSM. We first recall that the r.h.s. of the sum rule
1339: (\ref{eq:ex1stmom-cqsm1}) or (\ref{eq:ex1stmom-cqsm2}) is 
1340: the nucleon scalar charge defined by
1341: %
1342: \begin{equation}
1343:  \bar{\sigma} = \langle N | \bar{\psi}_u \psi_u + \bar{\psi}_d \psi_d 
1344:  | N \rangle .
1345:  \label{eq:scharge}
1346: \end{equation}
1347: %
1348: The point is that this low energy observable can be calculated within
1349: the CQSM, without asking for the distribution function $e^{(T = 0)} (x)$.
1350: It is given as
1351: %
1352: \begin{equation}
1353:  \bar{\sigma} = \bar{\sigma}_{val} + \bar{\sigma}_{sea},
1354:  \label{eq:schargevalsea}
1355: \end{equation}
1356: %
1357: with
1358: %
1359: \begin{eqnarray}
1360:  \bar{\sigma}_{val} &=& N_c \,\langle 0 | \gamma^0 | 0 \rangle, \\
1361:  \label{eq:schargeval}
1362:  \bar{\sigma}_{sea} &=& N_c \sum_{n < 0} 
1363:  \langle n | \gamma^0 | n \rangle .
1364:  \label{eq:schargesea}
1365: \end{eqnarray}
1366: %
1367: Numerically, we find that
1368: %
1369: \begin{equation}
1370:  \bar{\sigma}_{val} \simeq 1.7, \ \ \ 
1371:  \bar{\sigma}_{sea} \simeq 10.1 ,
1372:  \label{eq:scharge.num.valsea}
1373: \end{equation}
1374: %
1375: so that
1376: %
1377: \begin{equation}
1378:  \bar{\sigma} = \bar{\sigma}_{val} + \bar{\sigma}_{sea} \simeq 11.8 .
1379:  \label{eq:scharg.num.total}
1380: \end{equation}
1381: %
1382: Then, by admitting the validity of the first moment sum rule,
1383: one can extract the coefficient of the $\delta (x)$ term as follows :
1384: %
1385: \begin{equation}
1386:  C = \bar{\sigma}_{sea} - \int_{-1}^1 e^{(T=0)}_{reg} (x) d x 
1387:  \simeq 9.92 .
1388:  \label{eq:coeffc}
1389: \end{equation}
1390: %
1391: Our procedure for obtaining the coefficient $C$ is different from that
1392: of Schweitzer~\cite{S03}.
1393: After some consideration based on the gradient expansion analysis, he
1394: assumed that the Dirac-sea contribution to $e^{(T = 0)} (x)$ is saturated 
1395: by the $\delta(x)$ term with the coefficient $\Sigma_{\pi N} / m_0$, and 
1396: simply neglected the possible existence of the nonsingular contribution.
1397: In his treatment, then, the nontrivial shape of $e^{(T = 0)} (x)$ at 
1398: $x \neq 0$ solely comes from the contribution of $N_c$ valence level 
1399: quarks. Thus, the total distribution consists of these two terms as
1400: %
1401: \begin{equation}
1402:  e^{(T = 0)} (x) = \frac{\Sigma_{\pi N}}{m_0} \delta (x) 
1403:  + e_{val} (x) .
1404:  \label{eq:exschweitzer}
1405: \end{equation}
1406: %
1407: (Here, for simplicity, we ignore the term proportional to the product
1408: of $m_0$ and the unpolarized distribution function.)
1409: In our opinion, this procedure has a 
1410: danger of double counting. Within the framework of the CQSM, the total 
1411: $\pi N$ sigma term divided by the current quark mass $m_0$ is nothing but 
1412: the total scalar charge $\bar{\sigma}$ of the nucleon, which is made up 
1413: of the two terms, $\bar{\sigma}_{val}$ and $\bar{\sigma}_{sea}$.
1414: The $x$-integral of (\ref{eq:exschweitzer}) would then lead to
1415: %
1416: \begin{equation}
1417:  \int_{-1}^1 e^{(T = 0)} (x) d x 
1418:  = (\bar{\sigma}_{val} + \bar{\sigma}_{sea}) + \bar{\sigma}_{val} ,
1419:  \label{eq:exmomschweitzer}
1420: \end{equation}
1421: %
1422: which is obviously double counting the valence quark contribution to the 
1423: first moment sum rule. From the practical viewpoint, this double counting 
1424: is not so serious, since $\bar{\sigma}_{val}$ term turns out to be order
1425: of magnitude smaller than $\bar{\sigma}_{sea}$. This dominance of the 
1426: Dirac-sea contribution to the nucleon scalar charge is one of the
1427: distinguishing features of the CQSM. One can say that
1428: it is connected with the unique feature of this model, which is able to
1429: describe simultaneously a localized baryonic excitation 
1430: together with the nontrivial QCD vacuum structure with nonzero quark 
1431: condensate (or nonzero scalar quark density).
1432: In any case, we emphasize that the CQSM predicts fairly large scalar
1433: charge for the nucleon, i.e. $\bar{\sigma} \simeq 11.8$.
1434: Using the current quark 
1435: mass of $m_0 \simeq 5 \,\mbox{MeV}$, as an estimate, this gives 
1436: \begin{equation}
1437:  \Sigma_{\pi N} \equiv m_0 \,\bar{\sigma} \simeq 60 \,\mbox{MeV},
1438:  \label{eq:sigmaterm}
1439: \end{equation}
1440: which seems to favor relatively large values obtained from the 
1441: recent analysis of the pion-nucleon scattering
1442: amplitude~\cite{Koch82}\nocite{PASW99}\nocite{Olsson00}\nocite{PASW02}-\cite{Sainio02}.
1443: 
1444: Next we turn to the discussion of the second moment sum rule. We first
1445: point out that the $\delta (x)$ term in $e^{(T = 0)} (x)$ does not 
1446: contribute to the second moment. In the CQSM, then, the second moment of 
1447: $e^{(T = 0)} (x)$ receives contributions from two terms in the 
1448: distribution, i.e. the valence quark term $e^{(T = 0)}_{val} (x)$ and the 
1449: regular part of the vacuum polarization term $e^{(T = 0)}_{reg} (x)$. 
1450: After performing the numerical integration, we find that 
1451: %
1452: \begin{eqnarray}
1453:  \int_{-1}^1 x e^{(T = 0)}_{val} (x) d x &\simeq& 0.23,
1454:  \label{eq:ex2nd1} \\
1455:  \int_{-1}^1 x e^{(T = 0)}_{sea} (x) d x &=& \int_{-1}^1 
1456:  x e^{(T = 0)}_{reg} (x) d x \simeq -0.05 .
1457:  \label{eq:ex2nd2}
1458: \end{eqnarray}
1459: %
1460: The total second moment is therefore given by 
1461: %
1462: \begin{equation}
1463:  \int_{-1}^1 x e^{(T = 0)} (x) d x \simeq 0.23 - 0.05 \simeq 0.18 .
1464:  \label{eq:ex2nd3}
1465: \end{equation}
1466: %
1467: We recall that, within the CQSM, there is another independent method
1468: for evaluating the second moment.
1469: Since we are working in the chiral limit, we rewrite
1470: (\ref{eq:ex2ndmom-cqsm3}), by setting $m_0 = 0$, as
1471: %
1472: \begin{equation}
1473:  \int_{-1}^1 x e^{(T = 0)}(x) d x = N_c \cdot \frac{M}{M_N} \beta ,
1474:  \label{eq:ex2nd4}
1475: \end{equation}
1476: %
1477: or
1478: %
1479: \begin{eqnarray}
1480:  \int_{-1}^1 x e^{(T = 0)}_{val} (x) d x 
1481:  &=& N_c \frac{M}{M_N} \beta_{val}, \\
1482:  \label{eq:ex2nd5}
1483:  \int_{-1}^1 x e^{(T = 0)}_{sea} (x) d x 
1484:  &=& N_c \frac{M}{M_N} \beta_{sea},
1485:  \label{eq:ex2nd6}
1486: \end{eqnarray}
1487: %
1488: with
1489: %
1490: \begin{eqnarray}
1491:  \beta_{val} &=& \langle 0 | \frac{1}{2} (U + U^{\dagger}) | 0 \rangle, \\
1492:  \beta_{sea} &=& \sum_{n < 0} \langle n | \frac{1}{2} (U + U^{\dagger})
1493:  | n \rangle .
1494:  \label{eq:ex2nd7}
1495: \end{eqnarray}
1496: These quantities $\beta_{val}$ and $\beta_{sea}$ can be calculated 
1497: directly within the model, without invoking the corresponding 
1498: distribution functions. Numerically, we find that
1499: \begin{eqnarray}
1500:  N_c \frac{M}{M_N} \beta_{val} &\simeq& 0.23, \\
1501:  \label{eq:betval}
1502:  N_c \frac{M}{M_N} \beta_{sea} &\simeq& -0.06.
1503:  \label{eq:betsea}
1504: \end{eqnarray}
1505: These two numbers are consistent with the corresponding numbers in
1506: (\ref{eq:ex2nd1}) and (\ref{eq:ex2nd2}), obtained through the
1507: distribution functions, A small discrepancy between the numbers in
1508: (\ref{eq:ex2nd2}) and (\ref{eq:betsea}) may be interpreted as giving a
1509: measure of numerical errors introduced by the very delicate interpolation
1510: method for obtaining the vacuum polarization term of $e^{(T = 0)} (x)$.
1511: At any rate, we find that the CQSM predicts relatively small but 
1512: nonzero value for the second moment of $e^{(T = 0)} (x)$. Since we are 
1513: working in the chiral limit ($m_0 = 0$), this appears to contradict the 
1514: corresponding sum rule (\ref{eq:ex2ndmom}) derived on the basis of
1515: the QCD equations of motion, which states that the second moment
1516: of $e^{(T = 0)} (x)$ vanishes in the chiral limit.
1517: Does this discrepancy simply mean the limitation of the CQSM as an 
1518: effective theory of QCD?
1519: In our opinion, this is not necessarily the case by the following reasons.
1520: First of all, we point out that moment sum rules containing quark masses 
1521: are somewhat delicate one, since the masses are generally dependent on
1522: the renormalization scale. Secondly, if the QCD vacuum breaks the chiral
1523: symmetry spontaneously as is generally 
1524: believed, a quark acquires a dynamical mass of several hundred MeV, which 
1525: means that massless quarks are nowhere. Naturally, the situation is not
1526: so simple because of the color confinement.
1527: For instance, according to the picture of the MIT bag model, which realizes quark confinement by hand, at least the vacuum inside the bag is 
1528: perturbative and the quarks inside it remains massless.
1529: According to Shuryak~\cite{Shuryak88},
1530: the bag model is based on the idea that the hadron is a piece of 
1531: a qualitatively different (or ``perturbative") phase of the QCD vacuum.
1532: The physical picture of the CQSM for the baryon and the QCD vacuum is 
1533: fairly different from that of the bag model. According to the words of
1534: Shuryak again, the chiral models (including the CQSM) assume that the 
1535: vacuum is only slightly modified inside the hadron : the relative
1536: orientation of the right- and the left-handed quark fields is somewhat
1537: different. This last statement denotes the fact 
1538: that, in the basic lagrangian of the CQSM, the dynamical quark mass 
1539: parameter $M$ appears as a product with the chiral field
1540: $U^{\gamma_5} (x)$, which is space-time dependent.
1541: It is also the cause of the fact that the product of $M$ and $\beta$
1542: enters the r.h.s. of the second moment sum
1543: rule (\ref{eq:ex2ndmom-cqsm3}).
1544: This supports Schweitzer's viewpoint~\cite{S03} that the quantity
1545: $\beta M$ can be interpreted as an effective mass of quarks bound in
1546: the soliton background at least in the 2nd moment sum rule of
1547: $e^{(T=0)}(x)$. Numerically, we have
1548: %
1549: \begin{equation}
1550:  \beta M \sim 51 \,\mbox{MeV} .
1551:  \label{eq:effmass}
1552: \end{equation}
1553: %
1554: This value is smaller than the one obtained in \cite{S03}, since the
1555: contribution of the Dirac-sea quarks neglected in \cite{S03} works
1556: to reduce the value of $\beta$.
1557: 
1558: In any case, the nonzero value of the second moment of $e^{(T = 0)} (x)$
1559: is nothing contradictory at least within the framework of the CQSM
1560: in which massless quarks are nowhere at the model energy scale of
1561: about $600 \,\mbox{MeV}$.
1562: However, we anticipate that the dynamical quark mass $M$ is generally a
1563: scale dependent quantity which approaches zero in the high energy limit.
1564: The naive QCD sum rule for the second moment of $e^{(T = 0)} (x)$
1565: would be recovered in this limit.
1566: To verify the validity of this idea, what is crucial is experimental 
1567: determination of the second moment sum rule at the relatively low energy 
1568: scale close to the above-mentioned model energy scale.
1569: This may be in principle possible by inversely evoluting high energy
1570: data to low energy scale.
1571: 
1572: 
1573: 
1574: \subsection{isovector distribution $\mbox{\boldmath $e^{(T=1)}(x)$}$}
1575: 
1576: In the case of isovector distribution $e^{(T=1)}(x)$,
1577: no ultraviolet regularization is needed because its first moment,
1578: (\ref{eq:modelsum}), is related to the imaginary part of the
1579: Euclidian effective meson action in the background
1580: soliton field~\cite{GPPSU01} and it is ultraviolet finite.
1581: We have checked that the energy level sum (\ref{eq:nosing}) is
1582: stable enough against the increase of the cutoff momentum $k_{max}$,
1583: above $12M$. The final result for the isovector distribution
1584: $e^{(T=1)}(x)$ is shown in Fig.~\ref{fig:e1(x)}.
1585: %
1586: \begin{figure}[htb] \centering
1587: \begin{center}
1588:  \includegraphics[width=11.0cm]{e1.t.vl.vc.eps}
1589: \end{center}
1590: \vspace*{-0.5cm}
1591: \renewcommand{\baselinestretch}{1.20}
1592: \caption{The theoretical predictions of the CQSM for $e^{(T=1)}(x)$.
1593: The dot-dashed curve stands for the contribution
1594: of $N_c$ valence level quarks, the dashed curve the contribution
1595: of the Dirac-sea quarks, while the solid curves represents
1596: their sum.}
1597: \label{fig:e1(x)}
1598: \end{figure}%
1599: The dashed curve represents the contribution of the $N_c$ valence level
1600: quarks, the dot-dashed curve represents the contribution of
1601: the Dirac-sea quarks, while the solid curve represents their sum.
1602: In contrast to the isosinglet distribution, the Dirac-sea contribution
1603: has no singularity at $x=0$ and it is a smooth function
1604: in the whole region of $x$.
1605: The total contribution is given by the solid curve.
1606: 
1607: %
1608: \begin{figure}[htb] \centering
1609: \begin{center}
1610:  \includegraphics[width=11.0cm]{e0.vs.e1.eps}
1611: \end{center}
1612: \vspace*{-0.5cm}
1613: \renewcommand{\baselinestretch}{1.20}
1614: \caption{The comparison of the theoretical predictions for
1615: $e^{(T=0)}(x)$ and $e^{(T=1)}(x)$ at the model energy scale.}
1616: \label{fig:e0.vs.e1}
1617: \end{figure}%
1618: The first moment or the $x$-integral of this total contribution 
1619: gives the following value
1620: %
1621: \begin{equation}
1622:  \int_{-1}^1 e^{(T = 1)} (x) d x \simeq 0.28 ,
1623: \end{equation}
1624: %
1625: which is order of magnitude consistent with the estimate obtained from 
1626: the analysis of the non-electromagnetic proton-neutron mass difference.
1627: Shown in Fig.\ref{fig:e0.vs.e1} are the comparison of our final
1628: theoretical predictions for $e^{(T = 0)} (x)$ and $e^{(T = 1)} (x)$.
1629: One confirms that the magnitude of $e^{(T = 1)} (x)$ is much smaller
1630: than that of $e^{(T = 0)} (x)$ in conformity 
1631: with the large $N_c$ relation (\ref{eq:N_c-rel}).
1632: Combining these two distributions, we can now give final theoretical
1633: predictions for the chiral-odd twist-3 distribution function $e^a (x)$
1634: of each flavor $a$.
1635: Shown in Fig.\ref{fig:eu.ed}(a) are the distributions for the $u$-quark
1636: and the $\bar{u}$-quark, while Fig.\ref{fig:eu.ed}(b) gives the
1637: distributions for the $d$-quark and the $\bar{d}$-quark.
1638: 
1639: %
1640: \begin{figure}[htb] \centering
1641: \begin{center}
1642:  \includegraphics[width=15.0cm]{eu.ed.eps}
1643: \end{center}
1644: \vspace*{-0.5cm}
1645: \renewcommand{\baselinestretch}{1.20}
1646: \caption{The theoretical predictions for $e^u(x)$, $e^{d}(x)$,
1647: $e^{\bar{u}}(x)$, and $e^{\bar{d}}(x)$ at the model energy scale.}
1648: \label{fig:eu.ed}
1649: \end{figure}%
1650: 
1651: 
1652: 
1653: 
1654: \subsection{comparison with the empirical information from the CLAS
1655: measurements}
1656: 
1657: Here, we make a very preliminary comparison of our theoretical 
1658: predictions for $e(x)$ with the empirical information extracted from the
1659: high-energy semi-inclusive scatterings.
1660: Because of its chiral-odd nature, the distribution function $e(x)$
1661: does not appear in inclusive DIS cross sections.
1662: To extract any information for it, we must therefore carry out 
1663: more specific semi-inclusive type scattering experiments.
1664: Very recently, such an experiment has in fact been done by the CLAS
1665: Collaboration~\cite{CLAS}.
1666: They measured the azimuthal asymmetry $A_{LU}$ in the
1667: electro-production of pions from deeply inelastic scatterings of
1668: longitudinally polarized electrons off unpolarized protons. 
1669: 
1670: The first theoretical analysis of the CLAS data was carried out by
1671: Efremov et al.~\cite{EGS01},\cite{EGS02}.
1672: Their analysis assume that the beam single spin
1673: asymmetry measured by the CLAS group is dominantly generated by
1674: the so-called Collins mechanism~\cite{Collins93}.
1675: Under this assumption together with a particular
1676: parameterization for the Collins fragmentation function, they were
1677: able to extract the first information on the chiral-odd twist-3
1678: distribution function $e(x)$.
1679: Recently, this analysis was criticized by Yuan~\cite{Yuang03}.
1680: According to him, there may be another mechanism which competes
1681: with the Collins mechanism~\cite{Sivers90},\cite{Sivers91}.
1682: It is the leading order transverse momentum dependent parton
1683: distribution $h^{\bot}_1 (x, k_{\bot})$ convoluted with chiral-odd
1684: fragmentation function $\hat{e} (z)$.
1685: After all, the fact is that we still have poor knowledge about the
1686: mechanism that generates the beam single spin asymmetry in
1687: semi-inclusive deep inelastic scatterings. We must understand the
1688: mechanism of parton fragmentation processes into hadrons, especially
1689: the physics of time-reversal-odd fragmentation
1690: functions~\cite{Collins93},\cite{MT96}.
1691: We must also clarify the dynamics of transverse momentum dependent
1692: parton distribution functions in combination with the physics of
1693: chiral-odd fragmentation
1694: functions~\cite{Sivers90}\nocite{Sivers91}-\cite{MT96}.
1695: A truly reliable extraction of the chiral-odd twist-3
1696: distribution function $e(x)$,
1697: which is of our primary concern here, can be achieved only after
1698: more complete understanding of the above-mentioned mechanisms of
1699: semi-inclusive DIS processes.
1700: 
1701: Keeping this fact in mind, we shall proceed here by assuming the 
1702: dominance of the Collins mechanism. Under this assumption, the
1703: asymmetry measured by the CLAS experiment is interpreted to be
1704: proportional to 
1705: %
1706: \begin{equation}
1707:  A^{\sin \phi}_{LU} \sim -\frac{4 \pi \alpha^2 s}{Q^4} \,\lambda_e \,
1708:  2 y \sqrt{1 - y} \sum_a e^2_a x^2 e^a (x) H^{\bot a}_1 (z) ,
1709: \end{equation}
1710: %
1711: with $y = (P \cdot q) / (P \cdot l), z = (P \cdot p_h) / (P \cdot q)$
1712: and $s$ is the invariant mass squared of the photon-hadron system
1713: in the notation of \cite{EGS01}. $\lambda_e$ denotes the beam helicity.
1714: The chiral- and T-odd twist-2 ``Collins" fragmentation function
1715: $H^{\bot a}_1 (z)$ gives the probability of a spinless or unpolarized
1716: hadron to be created from a transversely polarized scattered quark.
1717: Using informations on $H^{\bot a}_1 (z)$ from HERMES
1718: data~\cite{herm01},\cite{herm02},
1719: one can then get direct information on the distribution function
1720: $e(x)$~\cite{EGS01},\cite{EGS02}.
1721: In the CLAS experiment, the azimuthal asymmetries
1722: $A^{\sin \phi}_{LU}$ for the process
1723: $\vec{e} p \rightarrow e^\prime \pi^+ X$ were measured at
1724: $Q^2 \sim 1.5 \,\mbox{GeV}^2$.
1725: Under the dominant-flavor-only approximation for the fragmentation
1726: functions, the semi-inclusive $\pi^+$ production measures the
1727: following combination of the distributions,
1728: %
1729: \begin{equation}
1730:  e^u (x) + \frac{1}{4} e^{\bar{d}} (x) .
1731: \end{equation}
1732: %
1733: In Fig.~\ref{fig:with-clas}, we make a comparison between the
1734: predictions of the CQSM for the above combinations of the distributions
1735: and the corresponding empirical information
1736: extracted from the CLAS data by Efremov et al.~\cite{EGS01},\cite{EGS02}
1737: under the assumption of Collins mechanism dominance.
1738: The theoretical distribution here corresponds to the energy scale of
1739: $Q^2 = 1.5 \,\mbox{GeV}^2$.
1740: The scale dependence of the distribution is taken
1741: into account by solving the leading-order DGLAP type equation
1742: obtained in the large $N_c$ limit~\cite{KN97}.
1743: (The starting energy scale of
1744: this evolution is taken to be $Q^2_{ini} \simeq 0.30 \,\mbox{GeV}^2$.)
1745: The distribution $e^u (x) + \frac{1}{4} e^{\bar{d}} (x)$ extracted
1746: from the CLAS data contains large errors mainly due to the large
1747: uncertainties of $H^{\bot}_1 (z)$ from the HERMES
1748: data~\cite{herm01},\cite{herm02}.
1749: Still, it was emphasized in \cite{EGS01},\cite{EGS02} that
1750: the extracted distribution
1751: is definitely larger than the ``twist-3 bound" and about two times
1752: smaller than the corresponding unpolarized distribution $f^a_1 (x)$
1753: at the same energy scale. One sees that our theoretical prediction for
1754: $e^u (x) + \frac{1}{4} e^{\bar{d}} (x)$ is in fairly good agreement
1755: with the extracted behavior from the CLAS data. The relatively small
1756: magnitude of the extracted $e(x)$ indicates that there must be a
1757: significant contribution to the $\pi N$ sigma-term sum rule from
1758: the small $x$ region. Whether this is due to the indicated
1759: $\delta$-function singularity in $e(x)$ or it is due to the
1760: yet-unresolved Regge behavior in the small $x$ region is difficult
1761: to judge at the present stage of study.
1762: It is highly desirable to extend the region of measurements to
1763: smaller $x$ region. This is important, because the unambiguous
1764: establishment of the violation of the $\pi N$ sigma-term sum rule
1765: would indirectly prove the existence of a novel $\delta$-function
1766: singularity in the distribution function $e(x)$ of the nucleon,
1767: which in turn may be interpteted as a manifestation of the nontrivial
1768: structure of QCD vacuum in an observable of a localized QCD excitation,
1769: i.e. the nucleon.
1770: 
1771: %
1772: \begin{figure}[htb] \centering
1773: \begin{center}
1774:  \includegraphics[width=11.0cm]{compclas_log.eps}
1775: \end{center}
1776: \vspace*{-0.5cm}
1777: \renewcommand{\baselinestretch}{1.20}
1778: \caption{The theoretical prediction for
1779: $e(x)=e^u(x)+\frac14e^{\bar{d}}(x)$ in comparison with the
1780: corresponding empirical information extracted from the CLAS data at
1781: $\langle Q^2\rangle =1.5\mbox{GeV}^2$ under the assumption of
1782: Collins mechanism dominance.}
1783: \label{fig:with-clas}
1784: \end{figure}
1785: %
1786: 
1787: Finally, we want to make some comments on the prediction for $e(x)$ 
1788: based on the MIT bag model. As mentioned in \cite{EGS02}, the bag model
1789: prediction of \cite{Signal97} evolved to the comparable energy
1790: scale of $Q^2 = 1 \,\mbox{GeV}^2$ is in qualitative agreement with the
1791: extracted $e(x)$ from the CLAS data in \cite{EGS02}.
1792: In our opinion, this agreement should be
1793: taken as fortuitous by the following reason. First, as already pointed
1794: out, the isosinglet scalar charge of the nucleon predicted by the
1795: MIT bag model is only about $15 \%$ of the value expected from the
1796: phenomenological knowledge of the $\pi N$ sigma-term.
1797: The fact is that the nucleon isoscalar charge is a quantity of 
1798: order 1 (or order $N_c$, more precisely) in the MIT bag model
1799: or in any other models which contains
1800: three valence-quark degrees of freedom only. 
1801: The situation is totally different in the CQSM. Although the
1802: contribution of the $N_c$ valance level quarks is of the same order
1803: as that of the MIT bag model, the vacuum polarization effect or the
1804: contribution of the Dirac-sea quarks give nearly seven times larger
1805: contribution as compared with that of the valence quarks, thereby
1806: reproducing the correct magnitude of the nucleon scalar charge or
1807: the $\pi N$ sigma-term. Unfortunately, this crucial difference
1808: between the two models is not reflected in the observable distribution
1809: function $e(x)$. Since the Dirac-sea contribution in the CQSM is nearly
1810: saturated by the $\delta$-function singularity, it happens that the
1811: distributions $e(x)$ at $x \neq 0$ predicted by the two models are not
1812: extremely different with each other. This is the reason why the
1813: naive MIT bag model, which fails to explain the magnitude of the
1814: $\pi N$ sigma-term, can reproduce the empirical distribution $e(x)$
1815: extracted from the CLAS data at least qualitatively.
1816: 
1817: %
1818: \begin{figure}[htb] \centering
1819: \begin{center}
1820:  \includegraphics[width=11.0cm]{pi+.pi-.1.5.eps}
1821: \end{center}
1822: \vspace*{-0.5cm}
1823: \renewcommand{\baselinestretch}{1.20}
1824: \caption{The predictions of the CQSM for $e^u(x)+(1/4)e^{\bar{d}}(x)$
1825: and $e^d(x)+(1/4)e^{\bar{u}}(x)$ evolved to the energy scale
1826: $Q^2 \simeq 1.5 \,\mbox{GeV}^2$ of the CLAS data from the initial scale
1827: of the model $Q^2_{ini} \simeq 0.30 \,\mbox{GeV}^2$. Also shown for
1828: comparison is the prediction of the MIT bag model evolved to the
1829: same scale from somewhat lower energy scale of $Q^2_{ini} \simeq
1830: 0.16 \,\mbox{GeV}^2$.}
1831: \label{fig:pi+.pi-}
1832: \end{figure}
1833: %
1834: 
1835: Still, we will show that there are some qualitative and observable
1836: differences between the predictions of the CQSM and the MIT bag model.
1837: The key observation here is that, for the spin-independent chiral-odd  
1838: twist-3 distribution functions, the MIT bag model predicts no flavor 
1839: dependence. That is, within the framework of the naive MIT bag model, 
1840: we have
1841: \begin{equation}
1842:  e^u (x) = e^d (x), \ \ 
1843:  e^{\bar{u}} (x) = e^{\bar{d}} (x) ,
1844: \end{equation}
1845: or more specifically
1846: \begin{equation} 
1847:  e^u (x) + \frac{1}{4} e^{\bar{d}} (x) 
1848:  = e^d (x) + \frac{1}{4} e^{\bar{u}} (x) .
1849: \end{equation}
1850: Such equalities can be expected to hold only in the fictitious limit
1851: of $N_c \rightarrow \infty$. As is in fact the case with the CQSM,
1852: for a finite value of $N_c$, the isovector distribution
1853: $e^{(T = 1)} (x) = e^u (x) - e^d (x)$ does not vanish, so that we
1854: definitely expect that
1855: \begin{equation}
1856:  e^u (x) + \frac{1}{4} e^{\bar{d}} (x) 
1857:  \neq e^d (x) + \frac{1}{4} e^{\bar{u}} (x) .
1858: \end{equation}
1859: %
1860: Fig.~\ref{fig:pi+.pi-} shows the comparison of the predictions of
1861: the two models for the distributions
1862: $e^u (x) + \frac{1}{4} e^{\bar{d}} (x)$ and
1863: $e^d (x) + \frac{1}{4} e^{\bar{u}} (x)$ evolved to the energy scale
1864: of CLAS experiment, i.e. $Q^2 \simeq 1.5 \,\mbox{GeV}^2$ from the
1865: initial energy scale of the model $Q^2_{ini} \simeq 0.30 \,\mbox{GeV}^2$.
1866: The solid and dashed curves here stand for the predictions of the
1867: CQSM, respectively for $e^u (x) + \frac{1}{4} e^{\bar{d}} (x)$ and
1868: $e^d (x) + \frac{1}{4} e^{\bar{u}} (x)$. On the other hand,
1869: the dot-dashed curve represents the prediction of the MIT bag model,
1870: which gives an identical answer for both these combinations of
1871: distributions. One sees that the CQSM predicts a sizably large
1872: difference between the two distributions
1873: $e^u (x) + \frac{1}{4} e^{\bar{d}} (x)$ and 
1874: $e^d (x) + \frac{1}{4} e^{\bar{u}} (x)$, in sharp contrast to
1875: the MIT bag model.
1876: In principle, the possible differences of these two distributions
1877: can be detected by performing a comparative analysis of the
1878: semi-inclusive $\pi^+$ and $\pi^-$ productions.
1879: 
1880: 
1881: 
1882: \section{Summary and Conclusion}
1883: 
1884: In summary, we have given theoretical predictions for the chiral-odd 
1885: twist-3 distribution function $e^a (x)$ of the nucleon with each flavor
1886: $a$ on the basis of the chiral quark soliton model.
1887: A prominent feature of the isosinglet combination of the distributions,
1888: $e^u (x) + e^d (x)$, is that its first moment is proportional to the
1889: familiar $\pi N$ sigma-term and that it 
1890: contains a delta-function singularity at $x = 0$. In the previous study 
1891: based on the derivative expansion technique, we demonstrated that the
1892: physical origin of this singularity can be traced back to the long-range
1893: quark-quark correlation of scalar type, which signals the
1894: spontaneous chiral symmetry breaking of the QCD vacuum.
1895: The present calculation, without recourse to 
1896: the derivative expansion type approximation, has revealed the following 
1897: facts. The isosinglet distribution $e^u (x) + e^d (x)$ consists of two 
1898: parts, i.e. the contribution of $N_c$ valence level quarks and that of 
1899: the Dirac sea quarks in the hedgehog mean field.
1900: The former takes a familiar shape of distribution 
1901: which has a peak around the value of $x \simeq 1/3$. On the other hand, 
1902: the latter certainly contains a $\delta$-function
1903: type singularity at $x = 0$, but it 
1904: also has nontrivial support for $x \neq 0$. The isovector 
1905: distribution $e^u (x) - e^d (x)$ also consists of the valence and 
1906: Dirac-sea contributions. For this distribution, however, no delta-function 
1907: type singularity is observed, which means that it is a regular function in
1908: all the range of $x$.
1909: 
1910: The moment sum rules of $e(x)$ provide us with 
1911: valuable information concerning the basic dynamical content of the model
1912: in view of the underlying theory, i.e. QCD.
1913: We showed that the first moment sum 
1914: rule for $e^u (x) + e^d (x)$ is satisfied within the model, if and only
1915: if the delta-function singularity is properly taken into account.
1916: Note however that the delta-function term alone does not saturate the
1917: first moment or the $\pi N$ sigma-term sum rule in contrast to the 
1918: previous argument based on the framework of the perturbative QCD.
1919: We also pointed out that the
1920: second moment sum rule for $e^u (x) + e^d (x)$ does not vanish even in
1921: the chiral limit in contrast to the QCD-equation-of-motion argument.
1922: In our opinion, this violation of the second moment sum rule does not
1923: necessarily show a defect of the model.  It is rather to be 
1924: interpreted as showing the limitation of the perturbative analysis
1925: as a tool of handling a bound state problem and/or the problem of
1926: masses nonperturbatively generated by the mechanism of the
1927: spontaneous chiral-symmetry breaking.
1928: We have also shown that the model prediction for the first moment of the
1929: isovector distribution $e^u (x) - e^d (x)$ comes out to be order of 
1930: magnitude consistent with the phenomenological estimate obtained from
1931: the nonelectromagnetic neutron-proton mass difference.
1932: 
1933: It was shown that the theoretical predictions for the
1934: distribution $e^u (x) + \frac{1}{4} e^{\bar{d}} (x)$ are in a good 
1935: agreement with the corresponding empirical information
1936: extracted from the CLAS data for the semi-inclusive $\pi^+$
1937: production under the assumption of the Collins mechanism dominance.
1938: This agreement, combined with our analysis explained in the text,
1939: impiles the existence of $\delta$-function singularity at $x = 0$
1940: in the isosinglet distribution $e^u (x) + e^d (x)$,
1941: although a definite conclusion must awaits for
1942: more complete measurements and more thorough 
1943: understanding of the reaction mechanism that generates the beam single 
1944: spin asymmetry in the semi-inclusive pion productions.
1945: 
1946:  Finally, we compare our theoretical predictions with those of the MIT 
1947: bag model. As shown in the body of the paper, the two models give
1948: accidentally close predictions for the distribution function 
1949: $e^u (x) + \frac{1}{4} e^{\bar{d}} (x)$ at $x \neq 0$. We have shown,
1950: however, that the CQSM predicts a sizably large difference between the 
1951: two distributions $e^u (x) + \frac{1}{4} e^{\bar{d}} (x)$ and $e^d (x) 
1952: + \frac{1}{4} e^{\bar{u}} (x)$, for which the MIT bag model makes no 
1953: difference. The predicted sizable difference between 
1954: the two combinations of distributions will be detected by performing 
1955: a comparative experimental analysis of the semi-inclusive 
1956: $\pi^+$ and $\pi^-$ productions.
1957: 
1958: 
1959: \newpage
1960: % Create the reference section using BibTeX:
1961: \bibliographystyle{unsrt}
1962: \bibliography{ex3}
1963: 
1964: \end{document}
1965: %
1966: % ****** End of file template.aps ******
1967: