hep-ph0605279/Gff.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,prd,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: \newcommand{\vecvar}[1]{\mbox{\boldmath$#1$}}
39: 
40: % Use the \preprint command to place your local institutional report
41: % number in the upper righthand corner of the title page in preprint mode.
42: % Multiple \preprint commands are allowed.
43: % Use the 'preprintnumbers' class option to override journal defaults
44: % to display numbers if necessary
45: %\preprint{}
46: 
47: %Title of paper
48: \title{Generalized form factors, generalized parton distributions\\
49: and the spin contents of the nucleon}
50: 
51: % repeat the \author .. \affiliation  etc. as needed
52: % \email, \thanks, \homepage, \altaffiliation all apply to the current
53: % author. Explanatory text should go in the []'s, actual e-mail
54: % address or url should go in the {}'s for \email and \homepage.
55: % Please use the appropriate macro foreach each type of information
56: 
57: % \affiliation command applies to all authors since the last
58: % \affiliation command. The \affiliation command should follow the
59: % other information
60: % \affiliation can be followed by \email, \homepage, \thanks as well.
61: \author{M.~Wakamatsu and Y.~Nakakoji}
62: %\email[]{wakamatu@phys.sci.osaka-u.ac.jp}
63: %\homepage[]{Your web page}
64: %\thanks{}
65: %\altaffiliation{}
66: \affiliation{Department of Physics, Faculty of Science, \\
67: Osaka University, \\
68: Toyonaka, Osaka 560-0043, JAPAN}
69: 
70: %Collaboration name if desired (requires use of superscriptaddress
71: %option in \documentclass). \noaffiliation is required (may also be
72: %used with the \author command).
73: %\collaboration can be followed by \email, \homepage, \thanks as well.
74: %\collaboration{}
75: %\noaffiliation
76: 
77: %\date{\today}
78: 
79: \begin{abstract}
80: With a special intention of clarifying the underlying spin contents
81: of the nucleon, we investigate the generalized form factors
82: of the nucleon, which are defined as the $n$-th $x$-moments of the
83: generalized parton distribution functions, within the
84: framework of the chiral quark soliton model.
85: A particular emphasis is put on the pion mass dependence
86: of final predictions, which we shall compare with the
87: predictions of lattice QCD simulations carried out in the
88: so-called heavy pion region around $m_\pi \simeq (700 \sim 900)
89: \,\mbox{MeV}$. We find that some observables are very sensitive
90: to the variation of the pion mass.
91: It will be argued that the negligible importance
92: of the quark orbital angular momentum indicated by the LHPC
93: and QCDSF lattice collaborations might be true in the unrealistic
94: heavy pion world, but it is not necessarily the case in our real
95: world close to the chiral limit.
96: \end{abstract}
97: 
98: % insert suggested PACS numbers in braces on next line
99: \pacs{12.39.Fe, 12.39.Ki, 12.38.Lg, 13.15.+g}
100: % insert suggested keywords - APS authors don't need to do this
101: %\keywords{}
102: 
103: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
104: \maketitle
105: 
106: % body of paper here - Use proper section commands
107: % References should be done using the \cite, \ref, and \label commands
108: %\section{}
109: % Put \label in argument of \section for cross-referencing
110: %\section{\label{}}
111: %\subsection{}
112: %\subsubsection{}
113: 
114: \section{Introduction}
115: The so-called ``nucleon spin crisis'' raised by
116: the European Muon Collaboration (EMC) measurement in 1988 is
117: one of the most outstanding findings in the field of hadron
118: physics \cite{EMC88},\cite{EMC89}.
119: The renaissance of the physics of high energy deep inelastic
120: scatterings is greatly indebted to this epoch-making finding.
121: Probably, one of the most outstanding progresses achieved recently in
122: this field of physics is the discovery and the subsequent research
123: of completely new observables called generalized parton distribution
124: functions (GPD). It has been revealed that the GPDs, which can be
125: measured through the so-called deeply-virtual Compton scatterings
126: (DVCS) or the deeply-virtual meson productions (DVMP), contain
127: surprisingly richer information than the standard parton distribution
128: functions \cite{MRGDH94}--\cite{RP02}. 
129: 
130: Roughly speaking, the GPDs are generalization of ordinary parton
131: distributions and the elastic form factors of
132: the nucleon. The GPDs in the most general form are functions of three
133: kinematical variables : the average longitudinal momentum fraction
134: $x$ of the struck parton in the initial and final states,
135: a skewdness parameter $\xi$ which measures the difference between
136: two momentum fractions, and the four-momentum-transfer square $t$ of
137: the initial and final nucleon. In the forward limit $t \rightarrow 0$,
138: some of the GPDs reduce to the usual quark, antiquark and gluon
139: distributions. On the other hand, taking the $n$-th moment of
140: the GPDs with respect to the variable $x$, one obtains the
141: generalizations of the electromagnetic form factors of the nucleon,
142: which are called the generalized form factors of the nucleon.
143: The complex nature of the GPDs, i.e. the fact that they are functions
144: of three variable, makes it quite difficult to grasp their
145: full characteristics both experimentally and theoretically.
146: From the theoretical viewpoint, it may be practical to begin studies
147: with the two limiting cases. The one is the forward limit of
148: zero momentum transfer. We have mentioned that, in this limit, some of
149: the GPDs reduce to the ordinary parton distribution function
150: depending on one variable $x$. However, it turns out that, even
151: in this limit, there appear some completely new distribution
152: functions, which cannot be accessed by the ordinary inclusive
153: deep-inelastic scattering measurements.
154: Very interestingly, it was shown by Ji that one of such
155: distributions contains valuable information on the total angular
156: momentum carried by the quark fields in the
157: nucleon \cite{Ji97}--\cite{JTH96}. This information,
158: combined with the known knowledge on the longitudinal quark
159: polarization, makes it possible to determine the quark orbital
160: angular momentum contribution to the total nucleon spin purely
161: experimentally.
162: 
163: Another relatively-easy-to-handle quantities are the
164: generalized form factors of the nucleon \cite{DFJK05},\cite{Diehl05},
165: which are given as the
166: non-forward nucleon matrix elements of the spin-$n$, twist-two
167: quark and gluon operators. Since these latter quantities are given
168: as the nucleon matrix elements of local operators, they can be
169: objects of lattice QCD simulations.
170: (It should be compared with parton distributions. The direct
171: calculation of parton distributions is beyond the scope of lattice
172: QCD simulations, since it needs to treat the nucleon matrix elements
173: of quark bilinears, which are {\it nonlocal in time}.)
174: In fact, two groups, the LHPC collaboration and the QCDSF collaboration
175: independently investigated the generalized form factors of the nucleon,
176: and gave several interesting predictions, which can in principle be
177: tested by the measurement of GPDs in the
178: near future \cite{LHPC02}--\cite{QCDSF05}.
179: Although interesting, there is no {\it a priori} reason to believe
180: that the predictions of these lattice simulations are realistic
181: enough. The reason is mainly that the above mentioned lattice
182: simulation were carried out in the heavy pion regime around
183: $m_\pi \simeq (700 \sim 900) \,\mbox{MeV}$  with neglect
184: of the so-called disconnected diagrams. Our real world is rather
185: close to the chiral limit with vanishing pion mass, and we know that,
186: in this limit, the Goldstone pion plays very important roles in
187: some intrinsic properties of the nucleon. The lattice simulation
188: carried out in the heavy pion region is in danger of missing some
189: important role of chiral dynamics. 
190: 
191: On the other hand, the chiral quark soliton model (CQSM) is an
192: effective model of baryons, which maximally incorporates the chiral
193: symmetry of QCD and its spontaneous breakdown \cite{DPP88},\cite{WY91}.
194: (See \cite{Wakamatsu92R}--\cite{ARW96} for early reviews.) It was already
195: applied to the physics of ordinary parton distribution functions
196: with remarkable success \cite{WGR96}--\cite{Wakamatsu03b}.
197: For instance, an indispensable role of
198: pion-like quark-antiquark correlation was shown to be essential to
199: understand the famous NMC measurement, which revealed the dominance
200: of the $\bar{d}$-quark over the $\bar{u}$-quark inside the
201: proton \cite{WK98},\cite{PPGWW99},\cite{Wakamatsu92}.
202: Then, it would be interesting to see what predictions the CQSM would
203: give for the quantities mentioned above. Now, the main purpose of the
204: present study is to study the generalized form factors of the
205: nucleon within the framework of the CQSM and compare
206: its predictions with those of the lattice QCD simulations.
207: Of our particular interest  here is to see the change of
208: final theoretical predictions against the variation of the pion mass.
209: Such an analysis is expected to give some hints for judging the
210: reliability of the lattice QCD predictions at the present level
211: for the observables in question.
212: 
213: The plan of the paper is as follows. In sect.II, we shall briefly
214: explain how to introduce the nonzero pion mass into the scheme
215: of the CQSM with Pauli-Villars regularization.
216: In sect.III, we derive the theoretical expressions for the
217: generalized form factors of the nucleon.
218: Sect.IV is devoted to the discussion of the results of the numerical
219: calculations. Some concluding remarks are then given in sect.V.
220: 
221: 
222: 
223: \section{Model lagrangian with pion mass term}
224: 
225: We start with the basic effective lagrangian of the chiral quark soliton 
226: model in the chiral limit given as
227: \begin{equation}
228:  {\cal L}_0 = \bar{\psi} (x) (i \not\!\partial - M U^{\gamma_5} (x) ) 
229:  \psi (x) ,
230: \end{equation}
231: with 
232: \begin{equation}
233:  U^{\gamma_5} (x) \ = \ 
234:  e^{i \gamma_5 \vecvar{\tau} \cdot \vecvar{\pi} (x) / f_{\pi}}
235:  \ = \ \frac{1 + \gamma_5}{2} \,U (x) + 
236:  \frac{1 - \gamma_5}{2} \,U^{\dagger} (x) .
237: \end{equation}
238: which describes the effective quark fields, with a dynamically generated 
239: mass $M$, strongly interacting with pions \cite{DPP88},\cite{WY91}.
240: Since one of the main purposes of the 
241: present study is to see how the relevant observables depend on pion mass, 
242: we add to ${\cal L}_0$ an explicit chiral symmetry breaking term 
243: ${\cal L}^{\prime}$ given by \cite{KWW99}
244: \begin{equation}
245:  {\cal L}^{\prime} = \frac{1}{4} \,f_{\pi}^2 \,m_{\pi}^2 \,\mbox{tr}_f
246:  \,[ U (x) + U^{\dagger} (x) - 2] . \label{eq:LSB}
247: \end{equation}
248: Here the trace in (\ref{eq:LSB}) is to be taken with respect to
249: flavor indices. 
250: The total model lagrangian is therefore given by
251: \begin{equation}
252:  {\cal L}_{CQM} = {\cal L}_0 + {\cal L}^{\prime} .
253: \end{equation}
254: Naturally, one could have taken an alternative choice in which the
255: explicit chiral-symmetry-breaking effect is introduced in the form of
256: current quark mass term as ${\cal L}^{\prime} = -m_0 \bar{\psi} \psi$.
257: We did not do so, because we do not know any consistent regularization
258: of such effective lagrangian with finite current quark mass within the 
259: framework of the Pauli-Villars subtraction scheme, as explained in
260: Appendix of \cite{KWW99}. 
261: The effective action corresponding to the above lagrangian is given as
262: \begin{equation}
263:  S_{eff} [U] = S_F [U] + S_M [U] , \label{eq:EnergSol}
264: \end{equation}
265: with
266: \begin{equation}
267:  S_F [U] = - \,i \,N_c \,\mbox{Sp} \,
268:  \ln ( i \not\!\partial - M U^{\gamma_5} ) ,
269: \end{equation}
270: and
271: \begin{equation}
272:  S_M [U]  = \int \frac{1}{4} \,f_{\pi}^2 \,m_{\pi}^2 \,
273:  \mbox{tr}_f \,[ U (x) + U^{\dagger} (x) - 2 ] \,d^4 x .
274: \end{equation}
275: Here $\mbox{Sp} \hat{O} = \int d^4 x \,\mbox{tr}_\gamma \,\mbox{tr}_f \,
276: \langle x | \hat{O} | x \rangle$ with $\mbox{tr}_\gamma$ and
277: $\mbox{tr}_f$ representing the trace of the Dirac gamma matrices and
278: the flavors (isospins), respectively.
279: The fermion (quark) part of the above action contains ultra-violet 
280: divergences. To remove these divergences, we must introduce physical
281: cutoffs. For the purpose of regularization, here we use the
282: Pauli-Villars subtraction scheme. As explained in \cite{KWW99},
283: we must eliminate not
284: only the logarithmic divergence contained in $S_f [U]$ but also
285: the quadratic and 
286: logarithmic divergence contained in the equation of motion shown below.
287: To get rid of all these troublesome divergence, we need at least two 
288: subtraction terms. The regularized action is thus defined as 
289: \begin{equation}
290:  S_{eff}^{reg} [U] = S_F^{reg} [U] + S_M [U] ,
291: \end{equation}
292: where
293: \begin{equation}
294:  S_F^{reg} [U] = S_F [U] - \sum_{i = 1}^2 \,c_i \,S_F^{\Lambda _i} [U] .
295: \end{equation}
296: Here $S_F^{\Lambda_i}$ is obtained from $S_F [U]$ with $M$ replaced by
297: the Pauli-Villars regulator mass $\Lambda_i$. These parameters are fixed
298: as follows.
299: First, the quadratic and logarithmic divergence contained in the
300: equation of motion (or in the expression of the vacuum quark condensate) 
301: can, respectively, removed if the subtraction constants satisfy the 
302: following two conditions :
303: \begin{eqnarray}
304:  M^2 &-& \sum_{i = 1}^2 c_i \Lambda_i^2 = 0 ,\\
305:  M^4 &-& \sum_{i = 1}^2 c_i \Lambda_i^4 = 0 .
306: \end{eqnarray}
307: (We recall that the condition which removes the logarithmic divergence 
308: in $S_f [U]$ just coincides with the 1st of the above conditions.)
309: By solving the above equations for $c_1$ and $c_2$, we obtain
310: \begin{eqnarray}
311:  c_1 &=& \left(\frac{M}{\Lambda_1} \right)^2 
312:  \frac{\Lambda_2^2 - M^2}{\Lambda_2^2 - \Lambda_1^2} , \\
313:  c_2 &=& - \left(\frac{M}{\Lambda_2} \right)^2 
314:  \frac{\Lambda_1^2 - M^2}{\Lambda_2^2 - \Lambda_1^2} ,
315: \end{eqnarray}
316: which constrains the values of $c_1$ and $c_2$, once $\Lambda_1$ and 
317: $\Lambda_2$ are given. For determining $\Lambda_1$ and $\Lambda_2$, 
318: we use two conditions
319: \begin{equation}
320:  \frac{N_c M^2}{4 \pi^2} \,\sum_{i = 1}^2 \,c_i \,
321:  \left(\frac{\Lambda_i}{M} \right)^2 \,
322:  \ln \left(\frac{\Lambda_i}{M} \right)^2 = f_{\pi}^2 ,
323: \end{equation}
324: and 
325: \begin{equation}
326:  \langle \bar{\psi} \psi \rangle _{vac} = \frac{N_c M^3}{2 \pi^2} \,
327:  \sum_{i = 1}^2 \,c_i \,\left(\frac{\Lambda_i}{M} \right)^4 
328:  \ln \left(\frac{\Lambda_i}{M} \right)^4 ,
329: \end{equation}
330: which amounts to reproducing the correct normalization of the pion
331: kinetic term in the effective meson lagrangian and also the empirical
332: value of the vacuum quark condensate.
333: To derive soliton equation of motion, we must first write down a 
334: regularized expression for the static soliton energy.
335: Under the hedgehog ansatz
336: $\vecvar{\pi} (\vecvar{x}) = f_{\pi} \hat{\vecvar{r}} F (r)$
337: for the background pion fields, it is obtained in the form : 
338: \begin{equation}
339:  E_{static}^{reg} [F (r)] = E_F^{reg} [F (r)] + E_M [F (r)] ,
340: \end{equation}
341: where the meson (pion) part is given by
342: \begin{equation}
343:  E_M [F (r)] = - \,f_{\pi}^2 \,m_{\pi}^2 \int d^3 x \,[\cos F (r) - 1] ,
344: \end{equation}
345: while the fermion (quark) part is given as 
346: \begin{equation}
347:  E_F^{reg} [F (r)] = E_{val} + E_{v p}^{reg} , \label{eq:Estatic}
348: \end{equation}
349: with
350: \begin{eqnarray}
351:  E_{val} &=& N_c \,E_0 , \\
352:  E_{v p}^{reg} &=& N_c \,\sum_{n < 0} \,(E_n - E_n^{(0)})
353:  - \sum_{i = 1}^N \,c_i \,N_c \,\sum_{n < 0} \,(E_n^{\Lambda_i} 
354:  - E_n^{(0) \Lambda_i}) .\label{eq:regenerg}
355: \end{eqnarray}
356: Here $E_n$ are the quark single-particle energies, given as the 
357: eigenvalues of the static Dirac Hamiltonian in the background pion
358: fields : 
359: \begin{equation}
360:  H \,| n \rangle = E_n | n \rangle ,
361: \end{equation}
362: with 
363: \begin{equation}
364:  H = \frac{\vecvar{\alpha} \cdot \nabla}{i} + \beta M \,
365:  [ \cos F(r) + i \gamma_5 \vecvar{\tau} \cdot 
366:  \hat{\vecvar{r}} \sin F(r)] , \label{eq:DiracH}
367: \end{equation}
368: while $E_n^{(0)}$ denote energy eigenvalues of the vacuum Hamiltonian
369: given by (\ref{eq:DiracH}) with $F (r) = 0$ (or $U = 1$).
370: The particular state $| n = 0 \rangle$, which is a discrete
371: bound-state orbital coming from the upper Dirac continuum under the
372: influence of the hedgehog mean field, is called the valence level.
373: The symbol $\sum_{n < 0}$ in (\ref{eq:regenerg}) denotes the
374: summation over all the negative energy eigenstates of $H$, i.e.
375: the negative energy Dirac continuum.
376: The soliton equation of motion is 
377: obtained from the stationary condition of $E_{static}^{reg} [F (r)]$
378: with respect to the variation of the profile function $F (r)$ : 
379: \begin{eqnarray}
380:  0 &=& \frac{\delta}{\delta F (r)} \,E_{static} [F (r)] \nonumber \\
381:  &=& 4 \pi r^2 \,\left\{ - \,M \,[S (r) \sin F (r) - P (r) \cos F (r)]
382:  + f_{\pi}^2 m_{\pi}^2 \sin F (r) \right\} ,
383: \end{eqnarray}
384: which gives 
385: \begin{equation}
386:  F (r) = \arctan 
387:  \left(\frac{P (r)}{S (r) - f_{\pi}^2 m_{\pi}^2 / M} \right) .
388:  \label{eq:profile}
389: \end{equation}
390: Here $S (r)$ and $P (r)$ are regularized scalar and pseudoscalar quark 
391: densities given as 
392: \begin{eqnarray}
393:  S (r) &=& S_{val} (r) + \sum_{n < 0} S_n (r) 
394:  - \sum_{i = 1}^2 \,c_i \,\frac{\Lambda_i}{M} \,\sum_{n < 0} \,
395:  S_n^{\Lambda_i} (r) , \\
396:  P (r) &=& P_{val} (r) + \sum_{n < 0} P_n (r) 
397:  - \sum_{i = 1}^2 \,c_i \,\frac{\Lambda_i}{M} \,\sum_{n < 0} \,
398:  P_n^{\Lambda_i} (r) ,
399: \end{eqnarray}
400: with
401: \begin{eqnarray}
402:  S_n (r) &=& \frac{N_c}{4 \pi} \int d^3 x \,\,
403:  \langle n | \vecvar{x} \rangle \,
404:  \gamma^0 \,\frac{\delta (| \vecvar{x}| - r )}{r^2} \,
405:  \langle \vecvar{x} | n \rangle , \\
406:  P_n (r) &=& \frac{N_c}{4 \pi} \int d^3 x \,
407:  \langle n | \vecvar{x} \rangle \,i \,
408:  \gamma^0 \,\gamma_5 \,\vecvar{\tau} \cdot \hat{\vecvar{r}} \,
409:  \frac{\delta (| \vecvar{x}| - r )}{r^2}
410:  \langle \vecvar{x} | n \rangle ,
411: \end{eqnarray}
412: while $S_n^{\Lambda_i} (r)$ and $P_n^{\Lambda_i} (r)$ are the
413: corresponding densities evaluated with the regulator mass
414: $\Lambda_i$ instead of the dynamical quark mass $M$.
415: We also note that $S_{val} (r) \equiv S_{n=0} (r)$ and
416: $P_{val} (r) \equiv P_{n=0} (r)$.
417: As usual, a self-consistent soliton solution is obtained as follows
418: with use of Kahana and Ripka's discretized momentum
419: basis \cite{KR84},\cite{KRS84}.
420: First by assuming an appropriate (though arbitrary) soliton profile
421: $F (r)$, the eigenvalue problem of the Dirac Hamiltonian is solved.
422: Using the resultant eigenfunctions and their associated eigenenergies,
423: one can calculate the regularized scalar and pseudoscalar
424: densities $S (r)$ and $P (r)$.
425: With use of these $S(r)$ and $P(r)$, Eq.(\ref{eq:profile})
426: can then be used to obtain a new soliton profile 
427: $F (r)$. The whole procedure above is repeated with this new profile 
428: $F (r)$ until the self-consistency is attained.
429: 
430: 
431: \section{Generalized Form Factors in the CQSM}
432: 
433:  Since the generalized form factors of the nucleon are given as moments
434: of generalized parton distributions (GPDs), it is convenient to start
435: with the theoretical expressions of the unpolarized GPDs
436: $H (x, \xi, t)$ and $E (x, \xi, t)$ within the CQSM.
437: Following the notation in \cite{Ossmann05},\cite{WT05},
438: we introduce the quantities
439: \begin{equation}
440:  {\cal M}_{s^{\prime} s}^{(I = 0)} \equiv \int \frac{d \lambda}{2 \pi} \,
441:  e^{i \lambda x} \,\langle \vecvar{p}^{\prime}, s^{\prime} 
442:  | \,\bar{\psi} \left(-\frac{\lambda n}{2} \right) \not\!n 
443:  \psi \left(\frac{\lambda n}{2} \right)
444:  | \vecvar{p}, s \rangle ,
445: \end{equation}
446: and 
447: \begin{equation}
448:  {\cal M}_{s^{\prime} s}^{(I = 1)} \equiv \int \frac{d \lambda}{2 \pi} \,
449:  e^{i \lambda x} \,\langle \vecvar{p}^{\prime}, s^{\prime} 
450:  | \,\bar{\psi} \left(-\frac{\lambda n}{2} \right) \tau_3 \not\!n 
451:  \psi \left(\frac{\lambda n}{2} \right) | \vecvar{p}, s \rangle .
452: \end{equation}
453: Here, the isoscalar and isovector combinations respectively
454: correspond to the sum and the difference of the quark flavors
455: $u$ and $d$.
456: The relation between these quantities and the generalized parton
457: distribution functions $H(x, \xi, t)$ and $E(x, \xi, t)$
458: are obtained most conveniently 
459: in the so-called Breit frame. They are given by
460: \begin{eqnarray}
461:  {\cal M}_{s^{\prime} s}^{(I = 0)} 
462:  &=& 2 \,\delta_{s^{\prime} s} \,H_E^{(I = 0)} (x, \xi, t)
463:  \ - \ \frac{i \,\varepsilon^{3 k l} \,\Delta^k}{M_N} \,
464:  (\sigma^l)_{s^{\prime} s} \, E_M^{(I = 0)} (x, \xi, t) ,\\
465:  {\cal M}_{s^{\prime} s}^{(I = 1)} 
466:  &=& 2 \,\delta_{s^{\prime} s} \,H_E^{(I = 1)} (x, \xi, t)
467:  \ - \ \frac{i \,\varepsilon^{3 k l} \,\Delta^k}{M_N} \,
468:  (\sigma^l)_{s^{\prime} s} \, E_M^{(I = 1)} (x, \xi, t) .
469: \end{eqnarray}
470: where
471: \begin{eqnarray}
472:  H_E^{(I = 0/1)} (x, \xi, t) &\equiv& H^{(I = 0/1)} (x, \xi, t)
473:  + \frac{t}{4 M_N^2} E^{(I = 0/1)} (x, \xi, t) ,\\
474:  E_M^{(I = 0/1)} (x, \xi, t) &\equiv& H^{(I = 0/1)} (x, \xi, t)
475:  + E^{(I = 0/1)} (x, \xi, t) .
476: \end{eqnarray}
477: These two independent combinations of $H (x, \xi, t)$ and $E (x, \xi, t)$
478: can be extracted through the spin projection of ${\cal M}^{(I = 0/1)}$ as
479: \begin{eqnarray}
480:  H_E^{(I)} (x, \xi, t) &=& \frac{1}{4} \,
481:  \mbox{tr} \{ {\cal M}^{(I)} \} , \label{eq:HEtrace} \\
482:  E_M^{(I)} (x, \xi, t) &=& 
483:  \frac{i \,M_N \,\varepsilon^{3 b m} \,
484:  \Delta^b}{2 \vecvar{\Delta}_{\perp}^2}
485:  \,\mbox{tr} \{ \sigma^m {\cal M}^{(I)} \} , \label{eq:EMtrace}
486: \end{eqnarray}
487: where ``tr" denotes the trace over spin indices, while 
488: $\vecvar{\Delta}_{\perp}^2 = \vecvar{\Delta}^2 - (\Delta^3)^2
489:  = -t - (-2 M_N \xi)^2$. Now, within the CQSM, it is possible to evaluate
490: the right-hand side (rhs) of (\ref{eq:HEtrace}) and (\ref{eq:EMtrace}).
491: Since the answers are already given in several previous
492: papers \cite{Ossmann05}--\cite{Petrov98}, 
493: we do not repeat the derivation. Here we comment only on the following
494: general structure of the theoretical expressions for relevant
495: observables in the CQSM. The leading contribution 
496: just corresponds to the mean field prediction, which is independent of 
497: the collective rotational velocity $\Omega$ of the hedgehog soliton.  
498: The next-to-leading order term takes account of the linear response of 
499: the internal quark motion to the rotational motion as an external 
500: perturbation, and consequently it is proportional to $\Omega$.
501: It is known that the leading-order term contributes to the isoscalar 
502: combination of $H_E (x, \xi, t)$ and to the isovector combination of 
503: $E_M (x, \xi, t)$, while the isoscalar part of $H_E (x, \xi, t)$ and the
504: isovector part of $E_M (x, \xi, t)$ survived only at the next-to-leading 
505: order of $\Omega$ (or of $1/N_c$). The leading-order GPDs are then
506: given as
507: \begin{eqnarray}
508:  H_E^{(I = 0)} (x, \xi, t) &=& M_N N_c \int \frac{d z^0}{2 \pi} \,
509:  \sum_{n \leq 0} \,e^{i z^0 (x M_N - E_n)} \int d^3 \vecvar{x} \,
510:  \Phi_n^{\dagger} (\vecvar{x}) \nonumber \\
511:  &\,& \times \ (1 + \gamma^0 \gamma^3) \,e^{-i(z^0/2) \hat{p}_3} \,
512:  e^{i \vecvar{\Delta} \cdot \vecvar{x}} \,e^{-i (z^0/2) \hat{p}_3} \,
513:  \Phi_n (\vecvar{x}) , \label{eq:GPDHEis} \\
514:  E_M^{(I = 1)} (x, \xi, t) 
515:  &=& \frac{2 \,i \,M_N^2 \,N_c}{3 \,(\vecvar{\Delta}_{\perp})^2} 
516:  \int \frac{d z^0}{2 \pi} \,\sum_{n \leq 0} \,e^{i z^0 (x M_N - E_n)} \,
517:  \int d^3 \vecvar{x} \,\Phi_n^{\dagger} (\vecvar{x}) \nonumber \\
518:  &\,& \times \ (1 + \gamma^0 \gamma^3) \,
519:  (\vecvar{\tau} \times \vecvar{\Delta})^3 \,
520:  e^{-i(z^0/2) \hat{p}_3} \,e^{i \vecvar{\Delta} \cdot \vecvar{x}} 
521:  \,e^{-i (z^0/2) \hat{p}_3} \,\Phi_n (\vecvar{x}) . \label{eq:GPDEMis}
522: \end{eqnarray}
523: Here the symbol $\Sigma_{n \leq 0}$ denotes the summation over the
524: occupied (the valence plus negative-energy Dirac sea) quark
525: orbitals in the hedgehog mean field.
526: On the other hand, the theoretical expressions for the isovector 
527: part of $H_E$ and the isoscalar part of $E_M$, which
528: survive at the next-to-leading order, are a little more complicated.
529: They are given as double sums over the single quark orbitals as
530: \begin{eqnarray}
531:  &\,& H_E^{(I = 1)} (x, \xi, t) \nonumber \\
532:  &=& - \frac{M_N N_c}{12 I} \int \frac{d z^0}{2 \pi} \,
533:  \left[\, \left\{ \sum_{m = all, n \leq 0 (E_m \neq E_n)} \,
534:  e^{-i E_n z^0} \ - \ 
535:  \sum_{n = all, m \leq 0 (E_m \neq E_n)} \,e^{-i E_m z^0} \right\} 
536:  \frac{1}{E_n - E_m} \right. \nonumber \\
537:  &\,& \left. \hspace{50mm} + \  \frac{1}{M_N} \,\frac{d}{d x} \,
538:  \sum_{m = all, n \leq 0 (E_m \neq E_n)} \,e^{-i E_n z^0} \right] \,
539:  e^{i x M_N z^0} \nonumber \\
540:  &\,& \hspace{30mm} \times \ \langle n | \tau^a | m \rangle 
541:  \langle m | \,\tau^a \,(1 + \gamma^0 \gamma^3) \,
542:  e^{-i (z^0/2) \hat{p}_3} \,
543:  e^{i \vecvar{\Delta} \cdot \vecvar{x}} \,e^{-i (z^0/2) \hat{p}_3}
544:  | n \rangle . \label{eq:GPDHEiv}
545: \end{eqnarray}
546: and
547: \begin{eqnarray}
548:  &\,& E_M^{(I = 0)} (x, \xi, t) \nonumber\\
549:  &=& i \,\frac{M_N^2 N_c}{2 I} \int \frac{d z^0}{2 \pi} \,
550:  \left[\, \left\{ \sum_{m = all, n \leq 0 (E_m \neq E_n)} \,
551:  e^{-i E_n z^0} \ - \ 
552:  \sum_{n = all, m \leq 0 (E_m \neq E_n)} e^{-i E_m z^0} \right\} 
553:  \frac{1}{E_n - E_m} \right. \nonumber \\
554:  &\,& \left. \hspace{50mm} + \frac{1}{M_N} \,\frac{d}{d x} 
555:  \sum_{m = all, n \leq 0 (E_m \neq E_n)} e^{-i E_n z^0} \right]
556:  e^{i x M_N z^0} \nonumber \\
557:  &\,& \hspace{25mm} \times \ \langle n | \tau^b | m \rangle 
558:  \langle m | \,(1 + \gamma^0 \gamma^3) \,e^{-i (z^0/2) \hat{p}_3} \,
559:  \frac{\varepsilon^{3 a b} \Delta^a}{\vecvar{\Delta}_{\perp}^2} \,
560:  e^{i \vecvar{\Delta} \cdot \vecvar{x}} \,e^{-i (z^0/2) \hat{p}_3}
561:  | n \rangle  . \label{eq:GPDEMiv}
562: \end{eqnarray}
563: These four expressions for the unpolarized GPDs,
564: i.e. (\ref{eq:GPDHEis}) $\sim$ (\ref{eq:GPDEMiv}),
565: are the basic starting equations for our present study 
566: of the generalized form factors of the nucleon within the CQSM.
567: There are infinite tower of generalized form factors, which are defined
568: as the $n$-th moments of GPDs.
569: In the present study, we confine ourselves to the 
570: 1st and the 2nd moments, which respectively corresponds to the standard 
571: electromagnetic form factors of the nucleon and the so-called
572: gravitational form factors.
573: We are especially interested in the second one, since they are believed
574: to contain valuable information on the spin contents of the nucleon
575: through Ji's angular momentum sum rule \cite{Ji97},\cite{HJL99}.
576: For each isospin channel, the 1st and the 2nd moments of
577: $H_E^{(I)} (x, 0, t)$ define the Sachs-electric and gravito-electric
578: form factors as 
579: \begin{equation}
580:  G_{E, 10}^{(I = 0/1)} (t) \equiv \int_{-1}^1 \,
581:  H_E^{(I =0/1)} (x, 0, t) \,d x , \label{eq:GE10def}
582: \end{equation}
583: and 
584: \begin{equation}
585:  G_{E, 20}^{(I = 0/1)} (t) \equiv \int_{-1}^1 \,x \,
586:  H_E^{(I =0/1)} (x, 0, t) \,d x .
587: \end{equation}
588: On the other hand, the 1st and the 2nd moments of $E_M^{(I)} (x, 0, t)$
589: respectively define the Sachs-magnetic and gravito-magnetic form factors
590: as
591: \begin{equation}
592:  G_{M, 10}^{(I = 0/1)} (t) \equiv \int_{-1}^1 \,
593:  E_M^{(I = 0/1)} (x, 0, t) \,d x ,
594: \end{equation}
595: and
596: \begin{equation}
597:  G_{M, 20}^{(I = 0/1)} (t) \equiv \int_{-1}^1 \,
598:  x \,E_M^{(I = 0/1)} (x, 0, t) \,d x . \label{eq:GM2nd}
599: \end{equation}
600: In the following, we shall explain how we can calculate the generalized
601: form factors based on the theoretical expressions of corresponding GPDs,
602: by taking $G_{E, 10}^{(I = 0)} (t)$ and $G_{E, 20}^{(I = 0)} (t)$
603: as examples. Setting $\xi = 0$ and integrating over $z^0$ in
604: (\ref{eq:GPDHEis}), we obtain
605: \begin{eqnarray}
606:  H_E^{(I = 0)} (x, 0, t) &=& M_N \,N_c \, \int d^3 x \,
607:  e^{i \vecvar{\Delta}_{\perp} \cdot \vecvar{x}} \nonumber \\
608:  &\times& \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
609:  (1 + \gamma^0 \gamma^3) \,\delta (x M_N - E_n - \hat{p}_3) \,
610:  \Phi_n (\vecvar{x}) . \label{eq:HEis}
611: \end{eqnarray}
612: Putting this expression into (\ref{eq:GE10def}), we have
613: \begin{eqnarray}
614:  G_{E, 10}^{(I = 0)} (t) &=& N_c \int d^3 x \,
615:  e^{i \vecvar{\Delta}_{\perp} \cdot \vecvar{x}} \,
616:  \sum_{n \leq 0} \Phi_n^{\dagger} (\vecvar{x})
617:  (1 + \gamma^0 \gamma^3) \Phi_n (\vecvar{x}) . 
618: \end{eqnarray}
619: It is easy to see that, using the generalized spherical symmetry of the 
620: hedgehog configuration, the term containing the factor
621: $\gamma^0 \gamma^3$ identically vanishes, so that $G_{E,10}^{(I=0)}(t)$
622: is reduced to a simple form as follows :
623: \begin{equation}
624:  G_{E, 10}^{(I = 0)} (t) = N_c \int d^3 x 
625:  \sum_{n \leq 0} \Phi_n^{\dagger} (\vecvar{x}) \,j_0 (\Delta_{\perp} x)
626:  \,\Phi_n (\vecvar{x}) .
627: \end{equation}
628: Aside from the factor $N_c \,( = 3)$, this is 
629: nothing but the known expression for the isoscalar Sachs-electric
630: form factor of the nucleon in the CQSM \cite{Christov96}.
631: 
632: A less trivial example is $G_{E, 20}^{(I = 0)} (t)$, which is defined
633: as the 2nd moment of  $H_E^{(I = 0)} (x, 0, t)$.
634: Inserting (\ref{eq:HEis}) into (\ref{eq:GE10def}) and carrying out the 
635: integration over $x$, we obtain
636: \begin{eqnarray}
637:  G_{E, 20}^{(I = 0)} (t) &=& \frac{1}{M_N} \,N_c \,\int d^3 x
638:  \,e^{i \vecvar{\Delta}_{\perp} \cdot \vecvar{x}}
639:  \, \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
640:  (1 + \gamma^0 \gamma^3) \,(E_n + \hat{p}_3) \,\Phi_n (\vecvar{x}) .
641: \end{eqnarray}
642: Using the partial-wave expansion of 
643: $e^{i \vecvar{\Delta}_{\perp} \cdot \vecvar{x}}$, this can be written as
644: \begin{eqnarray}
645:  G_{E, 20}^{(I = 0)} (t) &=& \frac{1}{M_N} \,N_c \,\int d^3 x \,
646:  \sum_{l, m} \,4 \pi \,i^l \,
647:  Y_{l m}^* (\hat{\Delta}_{\perp}) \nonumber \\
648:  &\times& \sum_{n \leq 0} \Phi_n^{\dagger} (\vecvar{x}) \,
649:  j_l (\Delta_{\perp} x) \,Y_{l m}(\hat{x}) \,(1 + \alpha_3) \,
650:  (E_n + \hat{p}_3) \,\Phi_n (\vecvar{x}) .
651: \end{eqnarray}
652: This can further be divided into four pieces as
653: \begin{equation}
654:  G_{E, 20}^{(I = 0)} (t) = \sum_{i = 1}^4 G_i ,
655: \end{equation}
656: where
657: \begin{equation}
658:  G_i = \frac{1}{M_N} \,N_c \,\int d^3 x \,\sum_{l, m} \,4 \pi \,i^l \,
659:  Y_{l m}^* (\hat{\Delta}_{\perp}) \,M_i ,
660: \end{equation}
661: with
662: \begin{eqnarray}
663:  M_1 &=& \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
664:  j_l (\Delta_{\perp} x) \,Y_{l m} (\hat{x}) \,E_n \,
665:  \Phi_n (\vecvar{x}) , \\
666:  M_2 &=& \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
667:  j_l (\Delta_{\perp} x) \,Y_{l m} (\hat{x}) \,\alpha_3 E_n \,
668:  \Phi_n (\vecvar{x}) , \\
669:  M_3 &=& \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
670:  j_l (\Delta_{\perp} x) \,Y_{l m} (\hat{x}) \,\hat{p}_3 \,
671:  \Phi_n (\vecvar{x}) , \\
672:  M_4 &=& \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
673:  j_l (\Delta_{\perp} x) \,Y_{l m} (\hat{x}) \,\alpha_3 \,\hat{p}_3 \,
674:  \Phi_n (\vecvar{x}) . \label{eq:M4}
675: \end{eqnarray}
676: To proceed further, we first notice that, by using the generalized
677: spherical symmetry, $M_1$ survives only when $l = m = 0$, i.e. 
678: \begin{equation}
679:  M_1 \ \sim \ \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
680:  j_0 (\Delta_{\perp} x) \,\frac{\delta_{l,0} \delta_{m,0}}{\sqrt{4 \pi}}
681:  \,E_n \Phi_n (\vecvar{x}) ,
682: \end{equation}
683: which leads to the result :
684: \begin{equation}
685:  G_1 = \frac{N_c}{M_N} \,\int \,d^3 x \,\sum_{n \leq 0} \,
686:  \Phi_n^{\dagger} (\vecvar{x}) \,
687:  j_0 (\Delta_{\perp} x) \,E_n \,\Phi_n (\vecvar{x}) .
688: \end{equation}
689: To evaluate $G_2$, we first note that 
690: \begin{equation}
691:  Y_{l m} (\hat{x})  \times \alpha_3 
692:  = \sum_{\lambda} \,\langle l m 1 0 | \lambda m \rangle \,
693:  [ Y_l (\hat{x}) \times \vecvar{\alpha} ]^{(\lambda)} .
694: \end{equation}
695: Here, the generalized spherical symmetry dictates that $\lambda$ must
696: be zero, so that the rhs of the above equation is effectively reduced to
697: \begin{equation}
698:  - \,\frac{1}{\sqrt{3}} \,\delta_{l,1} \, \delta_{m,0}
699:  [ Y_1 (\hat{x}) \times \vecvar{\alpha} ]^{(0)} .
700: \end{equation}
701: This then gives
702: \begin{eqnarray}
703:  G_2 &=& \frac{1}{M_N} \,N_c \,\int \,d^3 x \,4 \pi \,i \,
704:  Y_{10}^* (\hat{\Delta}_{\perp}) \,\left(-\frac{1}{\sqrt{3}} \right) \,
705:  \,\sum_{n \leq 0} \Phi_n^{\dagger} (\vecvar{x})
706:  j_1 (\Delta_{\perp} x) \,
707:  [ Y_1 (\hat{x}) \times \vecvar{\alpha} ]^{(0)} \,
708:  \Phi_n (\vecvar{x}) .
709: \end{eqnarray}
710: Owing to the identity
711: \begin{equation}
712:  Y_{10} (\hat{\Delta}_{\perp}) 
713:  \ = \ \sqrt{\frac{3}{4 \pi}} \,P_1 (\cos \frac{\pi}{2}) \ = \ 0 ,
714: \end{equation}
715: we therefore find that
716: \begin{equation}
717:  G_2 \ = \ 0 .
718: \end{equation}
719: Next we investigate the third term $G_3$. Using
720: \begin{eqnarray}
721:  Y_{l m } (\hat{x}) \times \hat{p}_3 &=& \sum_{\lambda} \,
722:  \langle l m 1 0 | \lambda m \rangle \,
723:  [ Y_l (\hat{x}) \times \hat{\vecvar{p}} ]^{(\lambda)}
724:  \ \sim \ - \,\frac{1}{\sqrt{3}} \,\delta_{l,1} \,\delta_{m,0} \,
725:  [ Y_1 (\hat{x}) \times \hat{\vecvar{p}} ]^{(0)} ,
726: \end{eqnarray}
727:  we obtain 
728: \begin{eqnarray}
729:  G_3 &=& \frac{1}{M_N} \,N_c \,\int d^3 x \,4 \pi \,i \,
730:  Y_{10}^* (\hat{\Delta}_{\perp}) \,\left(- \,\frac{1}{\sqrt{3}} \right)
731:  \, \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
732:  j_1 (\Delta_{\perp} x) \,
733:  [ Y_1 (\hat{x}) \times \hat{\vecvar{p}} ]^{(0)} \,
734:  \Phi_n (\vecvar{x}) .
735: \end{eqnarray}
736: This term vanishes by the same reason as $G_2$ does. The last term $G_4$ 
737: is a little more complicated.  We first notice that 
738: \begin{eqnarray}
739:  Y_{l m} (\hat{x}) \,\alpha_3 \,\hat{p}_3
740:  &=& Y_{l m} (\hat{x}) \,\sum_{\lambda} \,
741:  \langle 10 10 | \lambda 0 \rangle \,
742:  [ \vecvar{\alpha} \times \hat{\vecvar{p}} ]^{(\lambda)}_0 \nonumber \\
743:  &\sim& \sum_{\lambda} \,
744:  \langle 1 0 1 0 | \lambda 0 \rangle \,
745:  \langle l m \lambda 0 | 00 \rangle \,
746:  [ Y_l (\hat{x}) \times 
747:  [ \vecvar{\alpha} \times \hat{\vecvar{p}} ]^{(\lambda)} ]^{(0)} 
748:  \nonumber \\
749:  &=& \delta_{m, 0} \,\langle 1 0 1 0 | l 0 \rangle \,
750:  \langle l 0 l 0 | 0 0 \rangle \,
751:  [ Y_l (\hat{x}) \times 
752:  [ \vecvar{
753:  \alpha} \times \hat{\vecvar{p}} ]^{(l)} ] ^{(0)} ,
754: \end{eqnarray}
755: which dictates that $l$ must be 0 or 2. 
756: Inserting the above expression into (\ref{eq:M4}),
757: and using the explicit values of Clebsch-Gordan coefficients, $G_4$ 
758: becomes
759: \begin{eqnarray}
760:  G_4 &=& -\frac{1}{\sqrt{3}} \cdot \frac{N_c}{M_N}
761:  \int d^3 x \,\sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
762:  j_0 (\Delta_{\perp} x) \,
763:  [ \vecvar{\alpha} \times \hat{\vecvar{p}} ]^{(0)} \,
764:  \Phi_n (\vecvar{x}) \nonumber \\
765:  &+& \frac{\sqrt{4 \pi}}{\sqrt{6}} \cdot \frac{N_c}{M_N} 
766:  \int d^3 x \,\sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
767:  j_2 (\Delta_{\perp} x) \,
768:  [ Y_2 (\hat{x}) \times 
769:  [ \vecvar{\alpha} \times \hat{\vecvar{p}} ]^{(2)} ]^{(0)} \,
770:  \Phi_n (\vecvar{x}) .
771: \end{eqnarray}
772: Using the identities
773: \begin{eqnarray}
774:  \,[ \vecvar{\alpha} \times \hat{\vecvar{p}} ]^{(0)}
775:  &=& -\frac{1}{\sqrt{3}} \vecvar{\alpha} \cdot \hat{\vecvar{p}} , \\
776:  \,[ Y_2 (\hat{x}) 
777:  \times [ \vecvar{\alpha} \times \vecvar{p} ]^{(2)} ]^{(0)}
778:  &=& [ [ Y_2 (\hat{x}) \times \hat{\vecvar{p}} ]^{(1)} 
779:  \times \vecvar{\alpha} ]^{(0)} ,
780: \end{eqnarray}
781: $G_4$ can also be written as 
782: \begin{eqnarray}
783:  G_4 &=& \frac{1}{3} \cdot \frac{N_c}{M_N} \,
784:  \int \,d^3 x \,\sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
785:  j_0 (\Delta_{\perp} x) \,
786:  \vecvar{\alpha} \cdot \hat{\vecvar{p}} \,
787:  \Phi_n (\vecvar{x}) \nonumber \\
788:  &+& \frac{\sqrt{4 \pi}}{\sqrt{6}} \cdot \frac{N_c}{M_N} 
789:  \int \,d^3 x \,\sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
790:  j_2 (\Delta_{\perp} x) \,
791:  [ Y_2 (\hat{x}) \times \hat{\vecvar{p}} ]^{(1)}
792:  \times \vecvar{\alpha} ]^{(0)} \,
793:  \Phi_n (\vecvar{x}) .
794: \end{eqnarray}
795: Collecting the answers for $G_1, G_2, G_3$ and $G_4$, we finally obtain
796: \begin{eqnarray}
797:  G_{E, 20}^{(I = 0)} (t)
798:  &=& \frac{N_c}{M_N} \int \,d^3 x \,
799:  \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
800:  j_0 (\Delta_{\perp} x) \,E_n \,\Phi_n (\vecvar{x}) \nonumber \\
801:  &+& \frac{1}{3} \cdot \frac{N_c}{M_N} \,\int \,d^3 x \,
802:  \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
803:  j_0 (\Delta_{\perp} x) \,\vecvar{\alpha} \cdot \hat{\vecvar{p}} \,
804:  \Phi_n (\vecvar{x}) \nonumber \\
805:  &+& \frac{\sqrt{4 \pi}}{\sqrt{6}} \cdot \frac{N_c}{M_N} \,
806:  \int \,d^3 x \,
807:  \sum_{n \leq 0} \,\Phi_n^{\dagger} (\vecvar{x}) \,
808:  j_2 (\Delta_{\perp} x) \,
809:  [ Y_2 (\hat{x}) \times \hat{\vecvar{p}} ]^{(1)} 
810:  \times \vecvar{\alpha} ]^{(0)} \,\Phi_n (\vecvar{x}) .
811: \end{eqnarray}
812: 
813: Up to now, we have obtained the theoretical expressions for the
814: isoscalar combination of the generalized form factors
815: $G_{E, 10}^{(I = 0)} (t)$ and 
816: $G_{E, 20}^{(I = 0)} (t)$. For notational convenience, we summarize
817: these results in a little more compact forms as follows :
818: \begin{eqnarray}
819:  G_{E, 10}^{(I = 0)} (t) &=& \int_{-1}^1 
820:  H_E^{(I = 0)} (x, 0, t) d x
821:  \ = \ N_c \,\sum_{n \leq 0} \,
822:  \langle n | j_0 (\Delta_{\perp} r) | n \rangle , \label{eq:IntHE10is}
823: \end{eqnarray}
824: and 
825: \begin{eqnarray}
826:  G_{E, 20}^{(I = 0)} (t) &=& \int_{-1}^1 \,x 
827:  H_E^{(I = 0)} (x, 0, t) \,d x \nonumber \\
828:  &=& \frac{1}{M_N} \,\left\{ N_c \,\sum_{n \leq 0} \,E_n \,
829:  \langle n | j_0 (\Delta_{\perp} r) | n \rangle \right. \nonumber \\
830:  &+& N_c \,\sum_{n \leq 0} \,
831:  \langle n | j_0 (\Delta_{\perp} r) \,\frac{1}{3} \,
832:  \vecvar{\alpha} \cdot \vecvar{p} \,| n \rangle \nonumber \\
833:  &+& \left. \frac{\sqrt{4 \pi}}{\sqrt{6}} \,N_c \,
834:  \sum_{n \leq 0} \,
835:  \langle n | j_2 (\Delta_{\perp} r) \,[ [Y_2 (\hat{r}) 
836:  \times \vecvar{p} ]^{(1)} \times \vecvar{\alpha} ]^{(0)} \,
837:  | n \rangle \right\} . \label{eq:IntHE20is}
838: \end{eqnarray}
839: As pointed out before, $G_{E, 10}^{(I = 0)} (t)$ is $N_c \,( = 3)$
840: times the isoscalar combination of standard Sachs-electric form factor
841: of the nucleon.
842: Analogously, we may call $G_{E, 20}^{(I = 0)} (t)$ the
843: gravitoelectric form factor of the nucleon (its quark part),
844: since it is related to the nonforward nucleon matrix elements of
845: the quark part of the QCD energy momentum tensor.
846: 
847: The other generalized form factors can be obtained in a similar way.
848: The isovector part of the generalized electric form factors survive only 
849: at the next-to-leading order of $\Omega$. They are given as 
850: \begin{eqnarray}
851:  G_{E, 10}^{(I = 1)} (t) &=& \int_{-1}^1 H_E^{(I = 1)} (x, 0, t) \,
852:  d x \nonumber \\
853:  &=& \frac{1}{3 I} \,\left(\frac{N_c}{2} \right) \,\sum_{m > 0, n \leq 0}
854:  \frac{1}{E_m - E_n} \,\langle m || \vecvar{\tau} || n \rangle
855:  \langle m || j_0 (\Delta_{\perp} r) \vecvar{\tau} || n \rangle ,
856:  \label{eq:IntHE10iv}
857: \end{eqnarray}
858: and 
859: \begin{eqnarray}
860:  G_{E, 20}^{(I = 1)} (t) &=& 
861:  \int_{-1}^1 \,x \,H_E^{(I = 1)} (x, 0, t) \,
862:  d x \nonumber \\
863:  &=& \frac{1}{M_N} \cdot \frac{1}{3 I} \,\left(\frac{N_c}{2} \right) \,
864:  \sum_{m > 0, n \leq 0} \,
865:  \frac{1}{E_m - E_n} \,\langle m || \vecvar{\tau} || n \rangle 
866:  \nonumber \\
867:  &\times& \left\{ \,\frac{E_m + E_n}{2} \,
868:  \langle m || j_0 (\Delta_{\perp} r) \vecvar{\tau} || n \rangle \right.
869:  \nonumber \\
870:  &+& \langle m || \,j_0 (\Delta_{\perp} r) \,\frac{1}{3} \,
871:  (\vecvar{\alpha} \cdot \vecvar{p}) \,\vecvar{\tau} \,|| n \rangle
872:  \nonumber \\
873:  &+& \left. \frac{\sqrt{4 \pi}}{\sqrt{6}} \,
874:  \langle m || \,j_2 (\Delta_{\perp} r) \,
875:  [ [Y_2 (\hat{r}) \times \vecvar{p} ]^{(1)} \times 
876:  \vecvar{\alpha} ]^{(0)} \,\vecvar{\tau} \,|| n \rangle \right\} .
877: \end{eqnarray}
878: The isoscalar combination of the generalized magnetic form factors also 
879: survive only at the next-to-leading order of $\Omega$, so that they are 
880: given as double sums over the single-quark orbitals in the hedgehog mean 
881: field as 
882: \begin{eqnarray}
883:  G_{M, 10}^{(I = 0)} (t) &=& \int_{-1}^1 \,
884:  E_M^{(I = 0)} (x, 0, t) \,d x \nonumber \\
885:  &=& - \,\frac{M_N}{I} \,\left(\frac{N_c}{2} \right) \,
886:  \sum_{m > 0, n \leq 0} \,
887:  \frac{1}{E_m - E_n} \,\langle m || \vecvar{\tau} || n \rangle \,
888:  \langle m || \,\frac{j_1 (\Delta_{\perp} r)}{\Delta_{\perp} r} \,
889:  (\vecvar{r} \times \vecvar{\alpha}) \,|| n \rangle ,
890: \end{eqnarray}
891: and
892: \begin{eqnarray}
893:  G_{M, 20}^{(I = 0)} (t)
894:  &=& \int_{-1}^1 \,x E_M^{(I = 0)} (x, 0, t) \,d x \nonumber \\
895:  &=& - \,\frac{1}{I} \,\left(\frac{N_c}{2} \right) \,
896:  \sum_{m > 0, n \leq 0} \,\frac{1}{E_m - E_n} \,
897:  \langle m || \vecvar{\tau} || n \rangle \nonumber \\
898:  &\times& \left\{\, \frac{E_m + E_n}{2} \,
899:  \langle m || \,\frac{j_1 (\Delta_{\perp} r)}{\Delta_{\perp} r} \,
900:  (\vecvar{r} \times \vecvar{\alpha}) \,|| n \rangle \ + \ 
901:  \langle m || \,\frac{j_1 (\Delta_{\perp} r)}{\Delta_{\perp} r} \,
902:  \vecvar{L} \,|| n \rangle \right\} .
903: \end{eqnarray}
904: We recall that $G_{M, 10}^{(I = 0)} (t)$ just coincides with the known 
905: expression of the isoscalar Sachs-magnetic form factor of the nucleon
906: in the CQSM \cite{Christov95}.
907: On the other hand, $G_{M, 20}^{(I = 0)} (t)$ is sometimes called the 
908: gravitomagnetic form factor of the nucleon (its isoscalar part), which 
909: we can evaluate within the QCSM based on the above theoretical
910: expression. Finally, the leading-order contribution to the isovector
911: part of the generalized magnetic form factors are given as 
912: \begin{eqnarray}
913:  G_{M, 10}^{(I = 1)} (t) 
914:  &=& \int_{-1}^1  E_M^{(I = 1)} (x, 0, t) \,d x
915:  \ = \ - \,\frac{M_N}{3} \cdot N_c \,\sum_{n \leq 0} \,
916:  \langle n || \,\frac{j_1 (\Delta_{\perp} r)}{\Delta_{\perp} r} \,
917:  \vecvar{\tau} \cdot (\vecvar{r} \times \vecvar{\alpha})\,|| n \rangle ,
918:  \ \ \ \label{eq:MGivform}
919: \end{eqnarray}
920: and
921: \begin{eqnarray}
922:  G_{M, 20}^{(I = 1)} (t) 
923:  &=& \int_{-1}^1 x E_M^{(I = 1)} (x, 0, t) \,d x \nonumber \\
924:  &=& - \frac{1}{3} \cdot N_c \,\sum_{n \leq 0} \,
925:  \left\{\, E_n \,\langle n || \,
926:  \frac{j_1 (\Delta_{\perp} r)}{\Delta_{\perp} r} \,
927:  \vecvar{\tau} \cdot (\vecvar{r} \times \vecvar{\alpha})\,|| n \rangle 
928:  \ + \ \langle n || \,\frac{j_1 (\Delta_{\perp} r)}{\Delta_{\perp} r} 
929:  \vecvar{\tau} \cdot \vecvar{L} \,|| n \rangle \right\}. 
930:  \label{eq:MG2ivform} \ \ \ \ \ 
931: \end{eqnarray}
932: Especially interesting to us are the values of the generalized form 
933: factors in the forward limit $t \rightarrow 0$.
934: The consideration of this 
935: limit is also useful for verifying consistency of our theoretical
936: analyses, since it leads to fundamental sum rules discussed below.
937: We first consider the forward limit of $G_{E, 10}^{(I = 0)} (t)$.
938: From (\ref{eq:IntHE10is}), we find that
939: \begin{eqnarray}
940:  G_{E, 10}^{(I = 0)} (t = 0) &=& \int_{-1}^1 
941:  H_E^{(I = 0)} (x, 0, 0) \,d x
942:  \ = \ N_c \,\sum_{n \leq 0} \,1 .
943: \end{eqnarray}
944: Subtracting the corresponding vacuum contribution, this reduces to 
945: $N_c \,(= 3)$. If we remember the relation 
946: \begin{eqnarray}
947:  \int_{-1}^1 H_E^{(I = 0)} (x, 0, 0) d x 
948:  &=& \int_{-1}^1 H^{u + d} (x, 0, 0) d x
949:  \ = \ \int_{-1}^1 f^{u+ d} (x) d x 
950:  \ = \ N^u + N^d ,
951: \end{eqnarray}
952: the forward limit of (\ref{eq:IntHE10is}) just leads to the sum rule :
953: \begin{equation}
954:  G_{E, 10}^{(I = 0)} (t = 0) = N^u + N^d = 3 , \label{eq:udpSR}
955: \end{equation}
956: which denotes that the sum of the $u$-quark and $d$-quark numbers
957: in the proton is three.
958: 
959: Next we turn to the forward limit of $G_{E, 20}^{(I = 0)} (t)$,
960: which gives 
961: \begin{eqnarray}
962:  G_{E, 20}^{(I = 0)} (t = 0) &=& \frac{1}{M_N} \,
963:  \left\{\, N_c \,
964:  \sum_{n \leq 0} \,E_n \ + \ 
965:  \frac{1}{3} \,N_c \,\sum_{n \leq 0} \,
966:  \langle n | \vecvar{\alpha} \cdot \vecvar{p} | n \rangle \right\} .
967:  \label{eq:MomSum}
968: \end{eqnarray}
969: It is easy to see that, after regularization and vacuum subtraction,
970: the first term of the rhs of the 
971: above equation reduces to the fermion (quark) part of the soliton energy,
972: i.e. $E_F^{reg}$ in (\ref{eq:Estatic}).
973: It was proved in \cite{DPPPW96} that, in the CQSM 
974: with vanishing pion mass, the following identity holds :
975: \begin{equation}
976:  \sum_{n \leq 0} \,
977:  \langle n | \vecvar{\alpha} \cdot \vecvar{p} | n \rangle
978:  = 0
979: \end{equation}
980: In the case of finite pion mass, which we are handling, this identity
981: does not hold. Instead, we can prove (see Appendix) that
982: \begin{equation}
983:  \frac{1}{3} \,N_c \,\sum_{n \leq 0} \,
984:  \langle n | \vecvar{\alpha} \cdot \vecvar{p} | n \rangle = E_M .
985:  \label{eq:alfp}
986: \end{equation}
987: That is, the second term in the parenthesis of rhs of
988: eq.(\ref{eq:MomSum}) just coincides with the pion part of the soliton
989: energy (or mass).
990: Since the sum of the quark and pion part give the total soliton mass
991: $M_N$, we then find that
992: \begin{equation}
993:  G_{E, 20}^{(I = 0)} (t = 0) 
994:  = \frac{1}{M_N} \cdot M_N = 1 .
995: \end{equation}
996: In consideration of eq.(\ref{eq:IntHE20is}), this relation can also
997: be expressed as
998: \begin{eqnarray}
999:  \int_{-1}^1 x H_E^{(I = 0)} (x, 0, 0) \,d x 
1000:  &=& \int_{-1}^1 x H^{u + d} (x, 0, 0) \,d x
1001:  \ = \ \int_{-1}^1 \,x f^{u + d} (x) \,dx \ = \ 
1002:  \langle x \rangle^{u + d} = 1 , \ \ \ \label{eq:momSR}
1003: \end{eqnarray}
1004: which means that the total momentum fraction carried by quark fields 
1005: (the $u$- and $d$-quarks) is just unity. 
1006: This is an expected result, since the CQSM contains quark fields only
1007: (note that the pion is not an independent field of quarks), so that the
1008: total nucleon momentum should be saturated by the quark fields 
1009: alone.
1010: 
1011: Taking the forward limit of $G_{E, 10}^{(I = 1)} (t)$, we are again
1012: led to a trivial sum rule, constrained by the conservation low.
1013: In fact, we have
1014: \begin{eqnarray}
1015:  G_{E, 10}^{(I = 1)} (t= 0) &=& \frac{1}{I} \,
1016:  \left(\frac{N_c}{6} \right)
1017:  \sum_{m > 0, n \leq 0} \,\frac{1}{E_m - E_n} \,
1018:  \langle m || \vecvar{\tau} || n \rangle^2
1019:  \ = \ \frac{1}{I} \cdot I = 1 ,
1020: \end{eqnarray}
1021: thereby leading to
1022: \begin{eqnarray}
1023:  \int_{-1}^1 H_E^{(I = 1)} (x, 0, 0) \,d x 
1024:  &=& \int_{-1}^1 H^{u-d} (x, 0, 0) \,d x
1025:  \ = \ \int_{-1}^1 f^{u - d} (x) \,d x 
1026:  \ = \ N^u - N^d = 1 , \ \ \ \label{eq:udmSR}
1027: \end{eqnarray}
1028: which denotes that the difference of the $u$-quark and the
1029: $d$-quark numbers in the proton is just unity.
1030: On the other hand, the forward
1031: limit of $G_{E, 20}^{(I = 1)} (t)$ leads to the first nontrivial sum
1032: rule as
1033: \begin{eqnarray}
1034:  G_{E, 20}^{(I = 1)} (t = 0) 
1035:  &=& \int_{-1}^1 \,x H_E^{(I = 1)} (x, 0, 0) \,d x
1036:  \ = \ \int_{-1}^1 \,x f^{u - d} (x) \,d x
1037:  \ = \ \langle x \rangle^{u - d} \nonumber \\
1038:  &=& \frac{1}{M_N} \,\frac{1}{I} \,
1039:  \left(\frac{N_c}{6} \right) \,\sum_{m > 0, n \leq 0} \,
1040:  \frac{1}{E_m - E_n} \,\langle m || \vecvar{\tau} || n \rangle \nonumber \\
1041:  &\times& \left\{ \frac{E_m + E_n}{2} \,
1042:  \langle m || \vecvar{\tau} || n \rangle \ + \ 
1043:  \langle m || \,\frac{1}{3} \,(\vecvar{\alpha} \cdot \vecvar{p}) \,
1044:  \vecvar{\tau} \,|| n \rangle \right\} .
1045: \end{eqnarray}
1046: Since this quantity, which represents the difference of momentum 
1047: fraction carried by the $u$-quark and the $d$-quark in the proton,
1048: is not constrained by any conservation law, its actual value can be
1049: estimated only numerically.
1050: 
1051: Next we turn to the discussion of the forward limit of the generalized 
1052: magnetic form factors. First, the forward limit of 
1053:  $G_{M, 10}^{(I = 0)} (t)$ gives
1054: \begin{eqnarray}
1055:  &\,& G_{M, 10}^{(I = 0)} (t = 0)
1056:  \ = \ - \,\frac{M_N}{I} \left(\frac{N_c}{6} \right)
1057:  \sum_{m > 0, n \leq 0} \,\frac{1}{E_m - E_n} \,
1058:  \langle m || \vecvar{\tau} || n \rangle
1059:  \langle m || \vecvar{r} \times \vecvar{\alpha} || n \rangle ,
1060: \end{eqnarray}
1061: which reproduces the known expression of the isoscalar magnetic moment of
1062: the nucleon in the CQSM \cite{Christov95}.
1063: On the other hand, the forward limit of 
1064: $G_{M, 20}^{(I = 0)} (t)$ gives 
1065: \begin{eqnarray}
1066:  G_{M, 20}^{(I = 0)} (t = 0) 
1067:  &=& - \,\frac{1}{I} \,\left(\frac{N_c}{6} \right) \,
1068:  \sum_{m > 0, n \leq 0} \,\frac{1}{E_m - E_n} \,
1069:  \langle m || \vecvar{\tau} || n \rangle \nonumber \\
1070:  &\times& \left\{ \frac{E_m + E_n}{2} \,
1071:  \langle m || \vecvar{r} \times \vecvar{\alpha} || n \rangle
1072:  \ + \ \langle m || \vecvar{L} || n \rangle \right\} .
1073:  \label{eq:GM20isFL}
1074: \end{eqnarray}
1075: It was shown in \cite{Ossmann05} that the rhs of the
1076: above equation is just unity, i.e.
1077: \begin{equation}
1078:  G_{M,20}^{(I=0)} (t=0) \ = \ 1 . \label{eq:GM20ist0}
1079: \end{equation}
1080: %The proof goes as follows. Since $| n \rangle$ and $| m \rangle$ are the
1081: %eigenstates of the static Dirac hamiltonian $H$, it holds that 
1082: %\begin{equation}
1083: % \frac{E_m + E_n}{2} \,\langle m || \vecvar{r} 
1084: % \times \vecvar{\alpha} || n \rangle 
1085: % \ = \ \frac{1}{2} \,\langle m || \{ H, \vecvar{r} 
1086: % \times \vecvar{\alpha} \} || n \rangle .
1087: %\end{equation}
1088: %Here we proceed as 
1089: %\begin{eqnarray}
1090: % &\,& \{ H, (\vecvar{r} \times \vecvar{\alpha})_i \} \nonumber \\
1091: % &=& \varepsilon_{i j k} \,\{ \alpha_l p_l + \beta M (\cos F (r) 
1092: % + i \gamma_5 \vecvar{\tau} \cdot \hat{\vecvar{r}} 
1093: % \sin F (r) ), \,x_j \alpha_k \} \nonumber \\
1094: % &=& \varepsilon_{i j k} \{ \alpha_l p_l, x_j \alpha_k \}
1095: % \ + \ M \,\varepsilon_{i j k} \,x_j \,\{ \gamma^0 (\cos F (r) 
1096: % + i \gamma_5 \vecvar{\tau} \cdot \hat{\vecvar{r}} 
1097: % \sin F (r) ), \gamma^0 \gamma^k \} . \label{eq:AntiCR}
1098: %\end{eqnarray}
1099: %Using the identities
1100: %\begin{equation}
1101: % \{ \gamma^0, \gamma^0 \gamma^k \} = 0, \ \ \ 
1102: % \{ \gamma^0 \gamma_5, \gamma^0 \gamma^k \} = 0 ,
1103: %\end{equation}
1104: %we find that the 2nd term of (\ref{eq:AntiCR}) vanishes.
1105: %Next, by utilizing the identity,
1106: %\begin{equation}
1107: % \{ \alpha_l p_l, x_j \alpha_k \} 
1108: % = - \,i \,\alpha_j \,\alpha_k + 2 x_j p_k ,
1109: %\end{equation}
1110: %the 1st term of (\ref{eq:AntiCR}) can be rewritten as 
1111: %\begin{eqnarray}
1112: % \varepsilon_{i j k} \,\{ \alpha_l p_l, x_j \alpha_k \}
1113: % &=& - \,i \,\varepsilon_{i j k} \,\alpha_j \cdot \alpha_k 
1114: % + 2 \,(\vecvar{r} \times \vecvar{p})_i \nonumber \\
1115: % &=& 2 \,
1116: % \left(
1117: % \begin{array}{cc}
1118: % \sigma_i & 0 \\
1119: % 0 & \sigma_i
1120: % \end{array}
1121: % \right)
1122: % + 2 \,L_i
1123: % \ = \ 2 \,\Sigma_i + 2 \,L_i .
1124: %\end{eqnarray}
1125: %The consequence of the above consideration is a rather nontrivial 
1126: %identity given as 
1127: %\begin{equation}
1128: % \frac{E_m + E_n}{2} \,\langle m || \vecvar{r} 
1129: % \times \vecvar{\alpha} || n \rangle
1130: % \ = \ \langle m || \vecvar{\Sigma} + \vecvar{L} || n \rangle .
1131: %\end{equation}
1132: %Now putting this relation into (\ref{eq:GM20isFL}), we get
1133: %\begin{eqnarray}
1134: % 3 \,G_{M, 20}^{(I = 0)} (t = 0)
1135: % &=& - \,\frac{1}{I} \,\left(\frac{N_c}{6} \right) \,
1136: % \sum_{m > 0, n \leq 0} \,
1137: % \frac{1}{E_m - E_n} \,
1138: % \langle m || \vecvar{\tau} || n \rangle
1139: % \langle m || \vecvar{\Sigma} + 2 \vecvar{L} || n \rangle \nonumber \\
1140: % &=& - \frac{1}{I} \,\left(\frac{N_c}{6} \right) \,
1141: % \sum_{m > 0, n \leq 0} \,
1142: % \frac{1}{E_m - E_n} \,
1143: % \langle m || \vecvar{\tau} || n \rangle
1144: % \langle m || 2 \vecvar{J} || n \rangle .
1145: %\end{eqnarray}
1146: %Here we recall the definition of the grand spin operator 
1147: %$\vecvar{K} = \vecvar{J} + \frac{1}{2} \vecvar{\tau}$ and also the fact 
1148: %that $\langle m || K || n \rangle = 0$ for $m \neq n$. We thus obtain
1149: %\begin{eqnarray}
1150: % 3 \,G_{M, 20}^{(I = 0)} (t = 0)
1151: % &=& \frac{1}{I} \,\left(\frac{N_c}{6} \right) \,\sum_{m > 0, n \leq 0} \%, \frac{1}{E_m - E_n} \,
1152: % \langle m || \vecvar{\tau} || n \rangle^2
1153: % \ = \ \frac{1}{I} \cdot I \ = \ 1 . \label{eq:GM20isSR}
1154: %\end{eqnarray}
1155: In consideration of (\ref{eq:GM2nd}), this identity can be recast
1156: into a little different form as
1157: \begin{eqnarray}
1158:  1 &=& \int_{-1}^1 x E_M^{(I = 0)} (x, 0, 0) \,d x \nonumber \\
1159:  &=& \int_{-1}^1 x E_M^{u + d} (x, 0, 0) \,d x
1160:  \ = \ \int_{-1}^1 x \,
1161:  [  H^{u + d} (x, 0, 0) + E^{u + d} (x, 0, 0) ] \,d x .
1162: \end{eqnarray}
1163: Assuming the familiar angular momentum sum rule due to Ji
1164: \begin{equation}
1165:  \frac{1}{2} \int x [ H^{u + d} (x, 0, 0) + E^{u + d} (x, 0, 0) ] \,d x
1166:  = J^{u + d} ,
1167: \end{equation}
1168: the above identity claims that
1169: \begin{equation}
1170:  J^{u + d} = \frac{1}{2} , \label{eq:TotalSpin}
1171: \end{equation}
1172: which means that the nucleon spin is saturated by the quark fields alone.
1173: This is again a reasonable result, because the CQSM is an effective
1174: quark model which contains no explicit gluon fields.
1175: The derived identity (\ref{eq:GM20ist0}) has still another
1176: interpretation. Remembering the fact that 
1177: $G_{M, 20}^{(I = 0)} (t)$ consists of two parts as
1178: \begin{equation}
1179:  G_{M, 20}^{(I = 0)} (t) = A_{20}^{u+d} (t) + B_{20}^{u+d} (t) ,
1180: \end{equation}
1181: Eq.(\ref{eq:GM20ist0}) dictates that 
1182: \begin{equation}
1183:  A_{20}^{u + d} (0) + B_{20}^{u + d} (0) = 1 .
1184: \end{equation}
1185: Since it also holds that (the momentum sum rule)
1186: \begin{equation}
1187:  G_{M,20}^{(I = 0)} (0) = A_{20}^{u + d} (0) = 1 ,
1188: \end{equation}
1189: it immediately follows that
1190: \begin{equation}
1191:  B_{20}^{u + d} (0) = 0 ,
1192: \end{equation}
1193: which is interpreted as showing the absence of the {\it net quark
1194: contribution to the anomalous gravitomagnetic moment of the nucleon}.
1195: 
1196: Finally, we investigate the forward limit of the isovector combination
1197: of the generalized magnetic form factors. From eq. (\ref{eq:MGivform}),
1198: we get
1199: \begin{equation}
1200:  G_{M, 10}^{(I = 1)} (t = 0) = 
1201:  - \,\frac{M_N}{9} \,N_c \,\sum_{n \leq 0} \,
1202:  \langle n || \vecvar{\tau} \cdot (\vecvar{r} \times \vecvar{\alpha} ) 
1203:  || n \rangle ,
1204: \end{equation}
1205: which reproduces the known expression of the isovector magnetic form
1206: factor of the nucleon in the CQSM. On the other hand, letting
1207: $t \rightarrow 0$ in (\ref{eq:MG2ivform}), we have 
1208: \begin{eqnarray}
1209:  G_{M, 20}^{(I = 1)} (t = 0) &=& - \,\frac{1}{9} \,
1210:  N_c \,\sum_{n \leq 0} \,
1211:  \{ \,E_n \,\langle n || \vecvar{\tau} \cdot 
1212:  (\vecvar{r} \times \vecvar{\alpha}) || n \rangle
1213:  \ + \ \langle n || \vecvar{\tau} \cdot \vecvar{L} || n \rangle \} 
1214:  \nonumber \\
1215:  &=& \int_{-1}^1 \,x E_M^{(I = 1)} (x, 0, 0) \,d x .
1216: \end{eqnarray}
1217: As shown in \cite{WT05}, this sum rule can be recast into the form : 
1218: \begin{equation}
1219:  \frac{1}{2} \int_{-1}^1 x E_M^{(I = 1)} (x, 0, 0) \,d x 
1220:  = J^{(I = 1)} ,
1221: \end{equation}
1222: where $J^{(I = 1)}$ consists of two parts as 
1223: \begin{equation}
1224:  J^{(I = 1)} = J_f^{(I = 1)} + \delta J^{(I = 1)} .
1225: \end{equation}
1226: Here, the first part is given as a proton matrix element of the
1227: {\it free field expression} for the isovector total angular momentum
1228: operator of quark fields as 
1229: \begin{equation}
1230:  J_f^{(I = 1)} = \langle p \uparrow | \hat{J}_f^{(I = 1)} | 
1231:  p \uparrow \rangle ,
1232: \end{equation}
1233: with
1234: \begin{eqnarray}
1235:  \hat{J}_f^{(I = 1)} &=& \int \,\psi^{\dagger} (\vecvar{x}) \,
1236:  \tau_3 \,
1237:  \left[ (\vecvar{x} \times \hat{\vecvar{p}} )_3 + 
1238:  \frac{1}{2} \Sigma_3 \right] \,
1239:  \psi (\vecvar{x}) \,d^3 x \nonumber \\
1240:  &=& \hat{L}_f^{(I = 1)} + \frac{1}{2} \,\hat{\Sigma}^{(I = 1)} .
1241: \end{eqnarray}
1242: On the other hand, the second term is given as
1243: \begin{equation}
1244:  \delta J^{(I = 1)} \ = \ - M \left(\frac{N_c}{18} \right) \,
1245:  \sum_{n \leq 0} \,
1246:  \langle n | r \sin F (r) \gamma^0 
1247:  [ \vecvar{\Sigma} \cdot \hat{\vecvar{r}} \vecvar{\tau} 
1248:  \cdot \hat{\vecvar{r}} 
1249:  - \vecvar{\Sigma} \cdot \vecvar{\tau} ] | n \rangle .
1250: \end{equation}
1251: 
1252: 
1253: \section{Numerical Results and Discussions}
1254: 
1255: The model in the chiral limit contains two parameters, the weak pion 
1256: decay constant $f_{\pi}$ and the dynamical quark mass $M$. As usual, 
1257: $f_{\pi}$ is fixed to its physical value, i.e. 
1258: $f_{\pi} = 93 \,\mbox{MeV}$.
1259: For the mass parameter $M$, there is some argument based on the
1260: instanton liquid picture of the QCD vacuum that it is not extremely
1261: far from $350 \,\mbox{MeV}$ \cite{DPP88}.
1262: The previous phenomenological analysis of various static 
1263: baryon observables based on this model prefer a slightly larger value of 
1264: $M$ between $350 \,\mbox{MeV}$ and $425 \,\mbox{MeV}$
1265: \cite{Wakamatsu92R}--\cite{ARW96}.
1266: In the present analysis, we use the value $M = 400 \,\mbox{MeV}$.
1267: With this value of $M = 400 \,\mbox{MeV}$, we prepare self-consistent
1268: soliton solutions for seven values of $m_{\pi}$, i.e. $m_{\pi} = 0, 100,
1269: 200, 300, 400, 500,$ and $600 \,\mbox{MeV}$, in order to see the pion
1270: mass dependence of the generalized form factors etc.
1271: Favorable physical predictions of the model will be obtained by using
1272: the value of $M = 400 \,\mbox{MeV}$ and
1273: $m_\pi = 100 \,\mbox{MeV}$, since this set gives a self-consistent
1274: soliton solution close to the phenomenologically successful one
1275: obtained with $M = \,375 \,\mbox{MeV}$ and $m_\pi = 0 \,\mbox{MeV}$
1276: in the single-subtraction Pauli-Villars regularization scheme
1277: \cite{WK98}--\cite{PPGWW99}.
1278: 
1279: We first show in Fig.\ref{fig:profile}
1280: the soliton profile functions $F (r)$ obtained with several values of 
1281: $m_{\pi}$, i.e. $m_{\pi} = 0, 200, 400$, and $600 \,\mbox{MeV}$.
1282: One sees that the spatial size of the soliton profile becomes more
1283: and more compact as the pion mass increases.
1284: 
1285: \begin{figure}[htb] \centering
1286: \begin{center}
1287:  \includegraphics[width=9.0cm,height=7.0cm]{profile.eps}
1288: \end{center}
1289: \vspace*{-0.5cm}
1290: \renewcommand{\baselinestretch}{1.20}
1291: \caption{The self-consistent soliton profile functions obtained
1292: with $M = 400 \,\mbox{MeV}$ and $m_\pi = 0, 200, 400$,
1293: and $600 \,\mbox{MeV}$.}
1294: \label{fig:profile}
1295: \end{figure}%
1296: 
1297: We are now ready to show the theoretical predictions of the CQSM for the 
1298: generalized form factors. Since the corresponding lattice predictions
1299: are given for the generalized form factors $A_{n 0}^{u \pm d} (Q^2)$ and 
1300: $B_{n 0}^{u \pm d} (Q^2)$, which are the generalization of the standard 
1301: Dirac and Pauli form factors, we first write down the relations between 
1302: these form factors and the generalized Sachs-type factors, which we have 
1303: calculated in the CQSM. They are given by
1304: \begin{eqnarray}
1305:  A_{10}^{u + d} (t) &=& 
1306:  \left[\, G_{E, 10}^{(I = 0)} (t) + \tau \,
1307:  G_{M, 10}^{(I = 0)} (t) \right] \,/\, (1 + \tau) , \\
1308:  A_{20}^{u + d} (t) &=& 
1309:  \left[\, G_{E, 20}^{(I = 0)} (t) + \tau \,
1310:  G_{M, 20}^{(I = 0)} (t) \right] \,/\, (1 + \tau) , \\
1311:  A_{10}^{u - d} (t) &=& 
1312:  \left[\, G_{E, 10}^{(I = 1)} (t) + \tau \,
1313:  G_{M, 10}^{(I = 1)} (t) \right] \,/\, (1 + \tau) , \\
1314:  A_{20}^{u - d} (t) &=& 
1315:  \left[\, G_{E, 20}^{(I = 1)} (t) + \tau \,
1316:  G_{M, 20}^{(I = 1)} (t) \right] \,/\, (1 + \tau) ,
1317: \end{eqnarray}
1318: and
1319: \begin{eqnarray}
1320:  B_{10}^{u + d} (t) &=& 
1321:  \left[\, G_{M, 10}^{(I = 0)} (t) - G_{E, 10}^{(I = 0)} (t) \right] 
1322:  \,/\, (1 + \tau) , \\
1323:  B_{20}^{u + d} (t) &=& 
1324:  \left[\, G_{M, 10}^{(I = 0)} (t) - G_{E, 20}^{(I = 0)} (t) \right] 
1325:  \,/\, (1 + \tau) , \\
1326:  B_{10}^{u - d} (t) &=& 
1327:  \left[\, G_{M, 10}^{(I = 1)} (t) - G_{E, 10}^{(I = 1)} (t) \right] 
1328:  \,/\, (1 + \tau) , \\
1329:  B_{20}^{u - d} (t) &=& 
1330:  \left[\, G_{M, 20}^{(I = 1)} (t) - G_{E, 20}^{(I = 1)} (t) \right] 
1331:  \,/\, (1 + \tau) ,
1332: \end{eqnarray}
1333: where $\tau = - t / 4 M_N^2$.
1334: We recall that $A_{10} (t)$ and $B_{10}(t)$ are nothing but the
1335: standard Dirac and Pauli form factors of the nucleon :
1336: \begin{eqnarray}
1337:  A_{10}^{u+d} (t) &=& F_1^{(I = 0)} (t) 
1338:  =  F_1^p (t) + F_1^n (t) , \\
1339:  B_{10}^{u+d} (t) &=& F_2^{(I = 0)} (t) 
1340:  =  F_2^p (t) + F_2^n (t) , \\
1341:  A_{10}^{u-d} (t) &=& F_1^{(I = 1)} (t) 
1342:  =  F_1^p (t) - F_1^n (t)  , \\
1343:  B_{10}^{u-d} (t) &=& F_2^{(I = 1)} (t) 
1344:  =  F_2^p (t) - F_2^n (t) .
1345: \end{eqnarray}
1346: 
1347: Since the lattice simulations by the LHPC and QCDSF collaborations were 
1348: carried out in the heavy pion region around 
1349: $m_{\pi} \simeq (700 \sim 900) \,\mbox{MeV}$ and since the simulation
1350: in the small pion mass region is hard to perform, we think it interesting
1351: to investigate the pion mass dependence of the generalized form factors
1352: within the framework of the CQSM.
1353: For simplicity, we shall show the pion mass 
1354: dependence of the generalized form factors at the zero momentum transfer 
1355: only. We think it enough for our purpose because the generalized
1356: form factors at the zero 
1357: momentum transfer contain the most important information for clarifying
1358: the underlying spin structure of the nucleon.
1359: At zero momentum transfer, the 
1360: relations between the generalized Dirac and Pauli form factors and the 
1361: generalized Sachs-type form factors are simplified to become
1362: \begin{eqnarray}
1363:  A_{10}^{u + d} (0) &=& G_{E, 10}^{(I = 0)} (0) , \\ 
1364:  A_{20}^{u + d} (0) &=& G_{E, 20}^{(I = 0)} (0) , \\ 
1365:  A_{10}^{u - d} (0) &=& G_{E, 10}^{(I = 1)} (0) , \\ 
1366:  A_{20}^{u - d} (0) &=& G_{E, 20}^{(I = 1)} (0) ,
1367: \end{eqnarray}
1368: and
1369: \begin{eqnarray}
1370:  B_{10}^{u + d} (0) &=& G_{M, 10}^{(I = 0)} (0)
1371:  - G_{E, 10}^{(I = 0)} (0)  , \\ 
1372:  B_{20}^{u + d} (0) &=& G_{M, 20}^{(I = 0)} (0)
1373:  - G_{E, 20}^{(I = 0)} (0)  , \\ 
1374:  B_{10}^{u - d} (0) &=& G_{M, 10}^{(I = 1)} (0)
1375:  - G_{E, 10}^{(I = 1)} (0)  , \\ 
1376:  B_{20}^{u - d} (0) &=& G_{M, 20}^{(I = 1)} (0)
1377:  - G_{E, 20}^{(I = 1)} (0)  .
1378: \end{eqnarray}
1379: 
1380: \vspace{4mm}
1381: \begin{figure}[htb] \centering
1382: \begin{center}
1383:  \includegraphics[width=16.0cm,height=7.0cm]{A1020is.eps}
1384: \end{center}
1385: \vspace*{-0.5cm}
1386: \renewcommand{\baselinestretch}{1.20}
1387: \caption{The predictions of the CQSM for 
1388: $A_{10}^{u + d} (0)$ and $A_{20}^{u + d} (0)$ as functions of
1389: $m_{\pi}$ (the filled diamonds), together with the corresponding
1390: lattice predictions. Here, the open triangles correspond to
1391: the predictions of the LHPC group \cite{LHPC04}, while the open
1392: squares to those of the QCDSF collaboration \cite{QCDSF04a}.}
1393: \label{fig:a1020is}
1394: \end{figure}%
1395: \vspace{2mm}
1396: 
1397: Fig.\ref{fig:a1020is} shows the predictions of the CQSM for 
1398: $A_{10}^{u + d} (0)$ and $A_{20}^{u + d} (0)$ as functions of
1399: $m_{\pi}$, together with the corresponding lattice predictions.
1400: As for $A_{10}^{u + d} (0)$, the CQSM predictions and the lattice QCD
1401: predictions are both independent of $m_{\pi}$ and consistent with
1402: the constraint of the quark number sum rule :
1403: \begin{equation}
1404:  A_{10}^{u + d} (0) = N^u + N^d = 3 ,
1405: \end{equation}
1406: with high numerical precision. Turning to $A_{20}^{u + d} (0)$, one finds
1407: a sizable difference between the predictions of the CQSM and of the 
1408: lattice QCD. The lattice QCD predicts that 
1409: \begin{equation}
1410:  A_{20}^{u + d} (0) = \langle x \rangle^u + \langle x \rangle^d 
1411:  \simeq (0.6 \sim 0.7) ,
1412: \end{equation}
1413: which means that only about $(60 \sim 70)\%$ of the total nucleon 
1414: momentum is carried by the quark fields, while the rest is borne by the
1415: gluon fields. On the other hand, the CQSM predictions for the same
1416: quantity is 
1417: \begin{equation}
1418:  A_{20} ^{u + d} (0) = \langle x \rangle ^u + \langle x \rangle^d = 1 ,
1419: \end{equation}
1420: which means that the quark fields saturates the total nucleon momentum. 
1421: This may certainly be a limitation of an effective quark model, which
1422: contains no explicit gluon fields. Note, however, that the total quark
1423: momentum fraction $A_{20}^{u + d} (0)$ is a scale dependent quantity.
1424: The lattice result corresponds to the energy scale of
1425: $Q^2 = (2 \,\mbox{GeV})^2$ \cite{LHPC04},
1426: while the CQSM prediction should be taken as that of the model energy
1427: scale around $Q^2 = 0.30 \,\mbox{GeV}^2 \simeq (560 \,\mbox{MeV})^2$
1428: \cite{WK99}. We shall later make more meaningful comparison by taking
1429: care of the scale dependencies of relevant observables.
1430: 
1431: \vspace{4mm}
1432: \begin{figure}[htb] \centering
1433: \begin{center}
1434:  \includegraphics[width=16.0cm,height=7.0cm]{A1020iv.eps}
1435: \end{center}
1436: \vspace*{-0.5cm}
1437: \renewcommand{\baselinestretch}{1.20}
1438: \caption{The predictions of the CQSM for 
1439: $A_{10}^{u - d} (0)$ and $A_{20}^{u - d} (0)$ as functions of
1440: $m_{\pi}$, together with the corresponding lattice
1441: predictions \cite{LHPC04},\cite{QCDSF04a}.
1442: The meaning of the symbols are the same as in Fig.\ref{fig:a1020is}.}
1443: \label{fig:a1020iv}
1444: \end{figure}%
1445: \vspace{2mm}
1446: 
1447: Next, in Fig.\ref{fig:a1020iv}, we show the isovector combination
1448: of the generalized form 
1449: factors $A_{10}^{u - d} (0)$ and $A_{20}^{u - d} (0)$.
1450: The meaning of the 
1451: symbols are the same as in Fig.\ref{fig:a1020is}.
1452:  As for $A_{10}^{u - d} (0)$, both the 
1453: CQSM and the lattice simulation reproduce the quark number sum rule
1454: \begin{equation}
1455:  A_{10}^{u - d} (0) = N^u - N^d = 1 ,
1456: \end{equation}
1457: with good prediction.
1458: Turning to $A_{20}^{u - d} (0)$, one observes that the prediction of
1459: the CQSM shows somewhat peculiar dependence on the pion mass.
1460: Starting from a fairly small value in the chiral limit ($m_{\pi} = 0$), 
1461: it first increases as $m_{\pi}$ increases, but as $m_{\pi}$ further 
1462: increases it begins to decrease, thereby showing a tendency to match the 
1463: lattice prediction in the heavy pion region.
1464: Very interestingly, letting put aside the absolute value,
1465: a similar $m_{\pi}$ dependence is also observed in the chiral
1466: extrapolation of the lattice prediction for the momentum fraction
1467: $\langle x \rangle^u - \langle x \rangle^d$ shown in Fig.25 of 
1468: \cite{LHPC02}.
1469: Physically, the quantity $A_{20}^{u - d} (0)$ has a meaning of the 
1470: difference of the momentum fractions carried by the $u$-quark and
1471: the $d$-quark. The empirical value for it is 
1472: $A_{20}^{u - d} (0) = \langle x \rangle ^{u - d} = 0.154 \pm 0.003$
1473: \cite{LHPC02}. 
1474: One sees that the prediction of the CQSM in the chiral limit is not far 
1475: from this empirical information, although more serious comparison
1476: must take account of the scale dependence of $\langle x \rangle^{u-d}$.
1477: 
1478: \vspace{4mm}
1479: \begin{figure}[htb] \centering
1480: \begin{center}
1481:  \includegraphics[width=16.0cm,height=7.0cm]{B1020is.eps}
1482: \end{center}
1483: \vspace*{-0.5cm}
1484: \renewcommand{\baselinestretch}{1.20}
1485: \caption{The predictions of the CQSM for 
1486: $B_{10}^{u + d} (0)$ and $B_{20}^{u + d} (0)$ as functions of
1487: $m_{\pi}$, together with the corresponding lattice
1488: predictions \cite{LHPC05},\cite{QCDSF04b}.
1489: The meaning of the symbols are the same as in Fig.\ref{fig:a1020is}.}
1490: \label{fig:b1020is}
1491: \end{figure}%
1492: 
1493: Next, shown in Fig.\ref{fig:b1020is} are the CQSM predictions for
1494: $B_{10}^{u + d} (0)$ and $B_{20}^{u + d} (0)$.
1495: The former quantity is related to the isoscalar 
1496: combination of the nucleon anomalous magnetic moment as 
1497: $B_{10}^{u + d} (0) = \kappa^u + \kappa^d = 3 \,
1498: (\kappa^p + \kappa^n) \equiv 3 \kappa^{(T = 0)}$.
1499: (We recall that its 
1500: empirical value is $B_{10}^{u + d} (0) \simeq -0.36$.) We find that this 
1501: quantity is very sensitive to the variation of the pion mass. It appears 
1502: that the CQSM prediction $B_{10}^{u + d} (0) \simeq -1.5$ corresponding
1503: to chiral limit underestimates the observation significantly.
1504: However, the difference 
1505: is exaggerated too much in this comparison. In fact, if we carry out a 
1506: comparison in the total isoscalar magnetic moment of the nucleon 
1507: $\frac{1}{3} \,G_{M, 10}^{(I = 0)} (0) = \mu^p + \mu^n \equiv 
1508: \mu^{(T = 0)}$, the CQSM in the chiral limit 
1509: gives $\mu_{CQSM}^{(T = 0)} \simeq 0.5$ in comparison with the observed 
1510: value $\mu_{exp}^{(T = 0)} \simeq 0.88$.
1511: To our knowledge, no theoretical 
1512: predictions are given for this quantity by either of the LHPC or QCDSF 
1513: collaborations. The right panel of Fig.4 shows the predictions for
1514: $B_{20}^{u + d} (0)$, 
1515: which is sometimes called the isoscalar part of the nucleon anomalous 
1516: gravitomagnetic moment, or alternatively the {\it net quark contribution
1517: to the nucleon anomalous gravitomagnetic moment}.
1518: As already pointed out, the prediction of the CQSM for this quantity
1519: is exactly zero :
1520: i.e.
1521: \begin{equation}
1522:  B_{20}^{u + d} (0) = 0 ,
1523: \end{equation}
1524: The explicit numerical calculation also confirms it.
1525: It should be recognized that the 
1526: above result $B_{20}^{u + d} (0) = 0$ obtained in the CQSM is just
1527: a necessary consequence of the {\it momentum sum rule} and the
1528: {\it total nucleon spin sum rule}, both of which are saturated
1529: by the quark field only in the CQSM as
1530: \begin{equation}
1531:  A_{20}^{u + d} (0) = \langle x \rangle^{u + d} = 1 ,
1532: \end{equation}
1533: and 
1534: \begin{equation}
1535:  \frac{1}{2} \,[ A_{20}^{u + d} (0) + B_{20}^{u + d} (0) ]
1536:  \ = \ \langle J \rangle^{u + d} \ = \ \frac{1}{2} .
1537: \end{equation}
1538: In real QCD, the gluon also contributes to
1539: these sum rules, thereby leading to more general identities :
1540: \begin{eqnarray}
1541:  &\,& A_{20}^{u + d} (0) + A_{20}^g (0) = 1 , \\
1542:  &\,& [ A_{20}^{u + d} (0) + B_{20}^{u +d} (0) ] 
1543:  + [ A_{20}^g (0) + B_{20}^g (0) ] = 1 ,
1544: \end{eqnarray}
1545: which constrains that only the sum of 
1546: $B_{20}^{u + d} (0)$ and $B_{20}^g (0)$ is forced to vanish as
1547: \begin{equation}
1548:  B_{20}^{u + d} (0) + B_{20}^g (0) = 0 .
1549: \end{equation}
1550: (While we neglect here the contributions of other quarks than the
1551: $u$- and $d$-quarks, it loses no generality in our discussion below. 
1552: In fact, to include them, we have only to replace the combination 
1553: $u + d$ by $u + d + s + \cdots$.)
1554: The above nontrivial identity claims that the net contributions of quark
1555: and gluon fields to the anomalous gravitomagnetic moment of the nucleon
1556: must be zero. An interesting question is whether the quark and gluon 
1557: contribution to the anomalous gravitomagnetic moment vanishes separately
1558: or they are both large with opposite sign.
1559: A perturbative analysis based on a very simple toy model indicates
1560: the latter possibility \cite{BHMS01}.
1561: On the other hand, a nonperturbative analysis within the framework of
1562: the lattice QCD indicates that the net quark contribution to the
1563: anomalous gravitomagnetic moment is small or nearly zero,
1564: $B_{20}^{u+d}(0) = 0$ \cite{LHPC05},\cite{QCDSF04b}.
1565: (To be more precise, we sees that the prediction of the LHPC
1566: collaboration for $B_{20}^{u + d} (0)$ is slightly
1567: negative \cite{LHPC05}, while that of the QCDSF group is slightly
1568: positive \cite{QCDSF04b}.) This strongly indicates
1569: a surprising possibility that the quark and gluon contribution to the
1570: anomalous gravitomagnetic moment of the nucleon may separately vanish.
1571: Worthy of special mention here is an interesting argument given by
1572: Teryaev some years ago, claiming that the vanishing net quark
1573: contributions to the anomalous gravitomagnetic moment of the nucleon,
1574: violated in perturbation theory, is expected to be restored in
1575: full nonperturbative QCD due to the confinement
1576: \cite{Teryaev99},\cite{Teryaev98},\cite{Teryaev03}.
1577: Very interestingly, once it
1578: actually happens, it leads to a surprisingly
1579: simple result, i.e. the proportionality of the quark momentum and
1580: angular momentum fraction
1581: \begin{equation}
1582:  J^{u + d} = \frac{1}{2} \,\langle x \rangle^{u + d} ,
1583: \end{equation}
1584: as advocated by Teryaev \cite{Teryaev99},\cite{Teryaev98},\cite{Teryaev03}.
1585: A far reaching physical consequence resulting
1586: from this observation was extensively discussed 
1587: in our recent report \cite{Wakamatsu05}.
1588: (See also the discussion at the end of this section.)
1589: 
1590: \vspace{4mm}
1591: \begin{figure}[htb] \centering
1592: \begin{center}
1593:  \includegraphics[width=16.0cm,height=7.0cm]{B1020iv.eps}
1594: \end{center}
1595: \vspace*{-0.5cm}
1596: \renewcommand{\baselinestretch}{1.20}
1597: \caption{The predictions of the CQSM for 
1598: $B_{10}^{u - d} (0)$ and $B_{20}^{u - d} (0)$ as functions of
1599: $m_{\pi}$, together with the corresponding lattice
1600: predictions \cite{LHPC05},\cite{QCDSF04b}.
1601: The meaning of the symbols are the same as in Fig.\ref{fig:a1020is}.}
1602: \label{fig:b1020iv}
1603: \end{figure}%
1604: 
1605: Next, we show in Fig.\ref{fig:b1020iv} the predictions for the
1606: isovector case, i.e. $B_{10}^{u - d} (0)$ and $B_{20}^{u - d} (0)$.
1607: We recall first that the quantity $B_{10}^{u - d} (0)$ represents
1608: the isovector combination of the nucleon anomalous 
1609: magnetic moment $\kappa^{(T = 1)} \equiv \kappa^p - \kappa^n \,
1610: (\, = \kappa^u - \kappa^d \equiv \kappa^{(I=1)} )$,
1611: the empirical value of which is 
1612: known to be $\kappa^{(T = 1)} = 3.706$.
1613: One find that this quantity is extremely sensitive to the variation
1614: of the pion mass especially near $m_\pi = 0$.
1615: This is only natural if one remembers the important role of the pion
1616: cloud in the isovector magnetic moment of the nucleon.
1617: (One may notice that the prediction of the CQSM for $\kappa^{(T=1)}$
1618: underestimates a little its empirical value even in the chiral limit.
1619: We recall, however, that, within the framework of the CQSM, there is
1620: an important $1 / N_c$ correction or the 1st order rotational
1621: correction to some kind of isovector quantities
1622: like the isovector magnetic moment of the nucleon in question or
1623: the axial-vector coupling constant of the 
1624: nucleon \cite{WW93}--\cite{Wakamatsu96}. This next-to-leading
1625: correction in $1 / N_c$ should also be taken into account in more
1626: advanced investigations.)
1627: %Although the predictions by the LHPC and QCDSF collaborations are not
1628: %given in their papers, we conjecture that the 
1629: %lattice simulation carried out in the heavy pion region would
1630: %underestimate the magnitude of this quantity.
1631: Shown in the right panel of Fig.\ref{fig:b1020iv} is the theoretical
1632: predictions for $B_{20}^{u - d} (0)$,
1633: the half of which can be interpreted as the 
1634: difference of the total angular momentum carried by the $u$-quark and
1635: the $d$-quark fields according to Ji's angular momentum
1636: sum rule \cite{Ji97}.
1637: The CQSM predicts fairly small value for this quantity,
1638: in contrast to the lattice predictions of sizable magnitude.
1639: It seems that the pion mass dependence 
1640: rescues this discrepancy only partially.
1641: Here we argue that, the reason why the CQSM (in the chiral limit)
1642: gives rather small prediction for this quantity is intimately
1643: connected with the characteristic $x$ dependence of the quantity
1644: $E_M^{(I = 1)} (x, 0, 0)$, 
1645: the forward limit of the isovector unpolarized spin-flip GPD of the
1646: nucleon. To show it, we first recall that, within the theoretical
1647: frame work of the CQSM, $B_{10}^{u - d} (0)$ as well as
1648: $B_{20}^{u -d}(0)$ are calculated as 
1649: difference of $G_{M, 10}^{(I=1)}(0)$ and $G_{E, 10}^{(I=1)}(0)$ and of 
1650: $G_{M, 20}^{(I=1)} (0)$ and $G_{E, 20}^{(I=1)}(0)$, respectively, as 
1651: \begin{eqnarray}
1652:  B_{10}^{u - d} (0) &=& G_{M, 10}^{(I=1)} (0) - G_{E, 10}^{(I=1)} (0) ,\\
1653:  B_{20}^{u - d} (0) &=& G_{M, 20}^{(I=1)} (0) - G_{E, 20}^{(I=1)} (0) .
1654: \end{eqnarray}
1655: Although the quantities of the rhs can be calculated directly without 
1656: recourse to any distribution functions, they can also be evaluated as 
1657: $x$-weighted integrals of the corresponding GPDs as 
1658: \begin{eqnarray}
1659:  G_{M, 10}^{(I=1)} (0) &=& \int_{-1}^1 \,
1660:  E_M^{(I = 1)} (x, 0, 0) \,d x ,\\
1661:  G_{M, 20}^{(I=1)} (0) &=& \int_{-1}^1 \,
1662:  x E_M^{(I = 1)} (x, 0, 0) \,d x ,\\
1663:  G_{E, 10}^{(I=1)} (0) &=& \int_{-1}^1 \,H_E^{(I = 1)} (x, 0, 0) \,d x
1664:  \ = \ \int_{-1}^1 \,f^{u - d} (x) \,d x = N^u - N^d = 1 , \\
1665:  G_{E, 20}^{(I=1)} (0) &=& \int_{-1}^1 \,x H_E^{(I = 1)} (x, 0, 0) \,d x
1666:  \ = \ \int_{-1}^1 \,x f^{u - d} (x) \,d x 
1667:  = \langle x \rangle^u - \langle x \rangle^d .
1668: \end{eqnarray}
1669: The distribution function $E_M^{(I = 1)} (x, 0, 0)$ has already been 
1670: calculated within the CQSM in our recent paper \cite{WT05}.
1671: As shown there, the Dirac 
1672: sea contribution to this quantity has a sizably large peak
1673: around $x = 0$.
1674: Since this significant peak due to the deformed Dirac-sea quarks is
1675: approximately symmetric with respect to the reflection
1676: $x \longrightarrow -x$, it hardly contributes to the second moment 
1677: $G_{M, 20}^{(I = 1)} (0)$, whereas
1678: it gives a sizable contribution to the first moment
1679: $G_{M, 10}^{(I = 1)} (0)$.
1680: The predicted significant peak of $E_M^{(I = 1)} (x, 0, 0)$ around
1681: $x = 0$ can physically be interpreted as the effects of pion cloud.
1682: It can be convinced in several ways.
1683: First, we investigate how this behavior of $E_M^{(I = 1)} (x, 0, 0)$ 
1684: changes as the pion mass is varied.
1685: 
1686: \begin{figure}[htb] \centering
1687: \begin{center}
1688:  \includegraphics[width=9.0cm,height=7.0cm]{eiv_pi0.eps}
1689: \end{center}
1690: \vspace*{-0.5cm}
1691: \renewcommand{\baselinestretch}{1.20}
1692: \caption{The prediction of the CQSM for $E_M^{u-d}(x,0,0) = 
1693: E_M^{(I=1)}(x,0,0)$ obtained with $M = 400 \,\mbox{MeV}$ and
1694: $m_\pi = 0$.}
1695: \label{fig:eiv_pi0}
1696: \end{figure}%
1697: 
1698: \vspace{6mm}
1699: \begin{figure}[htb] \centering
1700: \begin{center}
1701:  \includegraphics[width=16.0cm,height=7.0cm]{eiv_pi24.eps}
1702: \end{center}
1703: \vspace*{-0.5cm}
1704: \renewcommand{\baselinestretch}{1.20}
1705: \caption{The $m_\pi$ dependence of $E_M^{u-d}(x,0,0)$.}
1706: \label{fig:eiv_pi24}
1707: \end{figure}%
1708: 
1709: Shown in Fig.\ref{fig:eiv_pi0} and in Fig.\ref{fig:eiv_pi24} are
1710: the CQSM predictions for $E_M^{(I = 1)} (x, 0, 0)$ with several values
1711: of $m_{\pi}$. i.e. $m_{\pi} = 0, 200$, and $400 \,\mbox{MeV}$.
1712: One clearly sees that the height of the peak around $x = 0$, due to the 
1713: deformed Dirac-sea quarks, decreases rapidly as $m_{\pi}$ increases.
1714: This supports our interpretation of this peak as the effects of pion
1715: clouds. On the other hand, one also observes that the magnitude of
1716: the valence quark contribution, peaked around
1717: $x \sim 1/3$, gradually increases as 
1718: $m_{\pi}$ becomes large. This behavior of $E_M^{(I = 1)} (x, 0, 0)$
1719: turns out to cause a somewhat unexpected $m_{\pi}$ dependence of 
1720: $G_{M, 10}^{(I = 1)} (0)$ and $G_{M, 20}^{(I = 1)} (0)$.
1721: As a function of $m_{\pi}$, the Dirac sea contribution to
1722: $G_{M, 10}^{(I = 1)} (0)$ decreases fast,
1723: whereas the valence quark contribution to it increases slowly, so 
1724: that the total $G_{M, 10}^{(I = 1)} (0)$ becomes a decreasing function
1725: of $m_{\pi}$.
1726: On the other hand, owing to the approximate odd-function nature of
1727: the Dirac sea contribution to $x E_M^{(I = 1)} (x, 0, 0)$ with respect
1728: to $x$, it hardly contributes to $G_{M, 20}^{(I = 0)} (0)$ independent
1729: of the pion mass, while the valence quark contribution to 
1730: $x E_M^{(I = 1)} (x, 0, 0)$ is an increasing function of $m_{\pi}$, 
1731: thereby leading to the result that the net $G_{M, 20}^{(I = 0)} (0)$ is 
1732: a increasing function of $m_{\pi}$. 
1733: 
1734: \begin{figure}[htb] \centering
1735: \begin{center}
1736:  \includegraphics[width=9.0cm,height=7.0cm]{fiv_pi0.eps}
1737: \end{center}
1738: \vspace*{-0.5cm}
1739: \renewcommand{\baselinestretch}{1.20}
1740: \caption{The prediction of the CQSM for $f^{u-d}(x,0,0)$
1741: obtained with $M = 400 \,\mbox{MeV}$ and $m_\pi = 0$.}
1742: \label{fig:fiv_pi0}
1743: \end{figure}%
1744: 
1745: \vspace{6mm}
1746: \begin{figure}[htb] \centering
1747: \begin{center}
1748:  \includegraphics[width=16.0cm,height=7.0cm]{fiv_pi24.eps}
1749: \end{center}
1750: \vspace*{-0.5cm}
1751: \renewcommand{\baselinestretch}{1.20}
1752: \caption{The $m_\pi$ dependence of $f^{u-d}(x,0,0)$.}
1753: \label{fig:fiv_pi24}
1754: \end{figure}%
1755: \vspace{2mm}
1756: 
1757: We can give still another support to the above-mentioned interpretation
1758: of the contribution of the Dirac-sea quarks. To see it, we first recall 
1759: that the theoretical unpolarized distribution function $f^{u - d} (x)$
1760: appearing in the decomposition
1761: \begin{equation}
1762:  E_M^{u-d} (x,0,0) = f^{u-d}(x) + E^{u-d}(x,0,0), 
1763: \end{equation}
1764: also has a sizable peak around $x = 0$ due to the deformed Dirac-sea
1765: quarks.
1766: As shown in Fig.\ref{fig:fiv_pi0} and in Fig.\ref{fig:fiv_pi24},
1767: this peak is again a rapidly decreasing function 
1768: of $m_{\pi}$, supporting our interpretation of it as the effects of pion 
1769: clouds. Here, we can say more. We point out that this small-$x$ behavior of $f^{u - d} (x)$ is just what is required by the famous
1770: NMC measurement \cite{NMC91}.
1771: To confirm it, first remember that the distribution $f^{u - d}(x)$ in the
1772: negative $x$ region should actually be interpreted as the distribution of
1773: antiquarks. To be explicit, it holds that
1774: \begin{equation}
1775:  \bar{u} (x) - \bar{d} (x) = -f^{u - d} (-x) \ \ \ (x \geq 0) .
1776: \end{equation}
1777: The large and positive value of $f^{u - d} (x)$ in the negative $x$
1778: region close to $x = 0$ means that $\bar{u}(x) - \bar{d} (x)$ is
1779: negative, i.e. the dominance of the $\bar{d}$-quark over
1780: the $\bar{u}$-quark inside the proton, which has been established
1781: by the NMC measurement \cite{NMC91}.
1782: 
1783: \vspace{4mm}
1784: \begin{figure}[htb] \centering
1785: \begin{center}
1786:  \includegraphics[width=9.0cm,height=7.5cm]{dbub.eps}
1787: \end{center}
1788: \vspace*{-0.5cm}
1789: \renewcommand{\baselinestretch}{1.20}
1790: \caption{The prediction of the CQSM for $\bar{d}(x) - \bar{u}(x)$
1791: evolved to $Q^2 = 2.3 \,\mbox{GeV}^2$ and $Q^2 = 54 \,\mbox{GeV}^2$
1792: in comparison with the Hermes and NuTeV data at the corresponding
1793: energy scales \cite{HERMES98},\cite{E866}.}
1794: \label{fig:dbdu}
1795: \end{figure}%
1796: \vspace{3mm}
1797: 
1798: Shown in Fig.\ref{fig:dbdu} are the predictions of the CQSM
1799: for $\bar{d} (x) - \bar{u} (x)$ evolved to the high energy scales
1800: corresponding to the experimental observation \cite{Saga96}.
1801: (The theoretical predictions here were obtained with
1802: $M = 400 \,\mbox{MeV}$ and $m_{\pi} = 100 \,\mbox{MeV}$.)
1803: The model reproduces well the observed behavior of 
1804: $\bar{d} (x) - \bar{u} (x)$, although the magnitude of the flavor
1805: asymmetry in smaller $x$ region seems to be slightly overestimated.
1806: It is a widely accepted fact that this flavor 
1807: asymmetry of the sea quark distribution in the proton can physically be 
1808: understood as the effects of pion cloud at least qualitatively
1809: \cite{HM91}--\cite{Wakamatsu92}.
1810: This then supports our interpretation of the effects of the deformed
1811: Dirac-sea quarks in $E_M^{(I = 1)} (x, 0, 0)$ and $f^{u - d} (x)$
1812: as the effects of pion clouds.
1813: 
1814: \begin{table}[htbp]
1815: \begin{center}
1816: \renewcommand{\baselinestretch}{1.20}
1817: \caption{The $m_\pi$ dependencies of $G_{M,20}^{(I=1)}(0)$,
1818: $G_{E,20}^{(I=1)}(0)$, and $B_{20}^{u-d} (0)$ in the CQSM with
1819: $M = 400 \,\mbox{MeV}$. \label{tabivb20}}
1820: \vspace{10mm}
1821: \begin{tabular}{cccc} \hline
1822:  \ \ $m_\pi \, (\mbox{MeV})$ \ \ & \ \ \ $G_{M,20}^{(I=1)}(0)$ \ \ \ & 
1823:  \ \ \ $G_{E,20}^{(I=1)}(0)$ \ \ \ & \ \ \ $B_{20}^{u-d} (0)$ \ \ \ 
1824:  \\ \hline\hline
1825:  \   0 \ & 0.361 & 0.228 & 0.133 \\ \hline
1826:  \ 100 \ & 0.392 & 0.276 & 0.116 \\ \hline
1827:  \ 200 \ & 0.452 & 0.327 & 0.125 \\ \hline
1828:  \ 300 \ & 0.519 & 0.350 & 0.169 \\ \hline
1829:  \ 400 \ & 0.579 & 0.354 & 0.225 \\ \hline
1830:  \ 500 \ & 0.640 & 0.347 & 0.293 \\ \hline
1831:  \ 600 \ & 0.716 & 0.328 & 0.388 \\ \hline
1832: \end{tabular}
1833: \end{center}
1834: \renewcommand{\baselinestretch}{1.20}
1835: \end{table}
1836: 
1837: We show in Table \ref{tabivb20} the model predictions for 
1838: $G_{M, 20}^{(I = 1)} (0), G_{E, 20}^{(I = 1)} (0)$, and 
1839: $B_{20}^{u - d} (0) = G_{M, 20}^{(I = 1)} (0) - G_{E, 20}^{(I = 1)} (0)$ 
1840: as functions of $m_{\pi}$.
1841: One sees that the value of $G_{M, 20}^{(I = 1)} (0)$ with $m_{\pi} = 0$
1842: is an order of $0.3 \sim 0.4$.
1843: As already pointed out, it gradually increases as $m_{\pi}$ becomes
1844: large. This is also the case with $G_{E, 20}^{(I = 1)} (0)$.
1845: As a consequence, the isovector combination of the nucleon anomalous 
1846: gravitomagnetic moment $B_{20}^{u - d} (0)$, which is obtained as a 
1847: difference of the above two quantities, is also an increasing function
1848: of $m_{\pi}$, thereby having a tendency to come closer to the lattice
1849: prediction given in the heavy pion region. Still, the CQSM predictions 
1850: $B_{20}^{u - d} (0) \simeq 0.3$ around $m_{\pi} = 500 \,\mbox{MeV}$ is a factor of two smaller than the corresponding lattice prediction 
1851: $B_{20}^{u - d} (0) \simeq 0.6$.
1852: Now we summarize the reason why the CQSM gives fairly 
1853: small prediction for $B_{20}^{u -d} (0)$. It is due to two types of 
1854: cancellations. The first is the cancellation of the potentially large 
1855: contribution of Dirac-sea quarks arising from the approximately 
1856: antisymmetric behavior of the Dirac sea contribution to 
1857: $x E_M^{(I = 1)} (x, 0, 0)$ as well as $x f^{u - d} (x)$.
1858: The second is the cancellation between the total gravitomagnetic moment 
1859: $G_{M, 20}^{(I = 1)} (0)$ and its canonical part
1860: $G_{E, 20}^{(I = 1)} (0)$.
1861: We are not sure whether the lattice simulation carried out in the heavy 
1862: pion region with neglect of the so-called disconnected diagrams can 
1863: efficiently take account of such effects of chiral dynamics as discussed 
1864: above.
1865: 
1866: So far, we have investigated the pion mass dependence of the
1867: forward limit of the generalized form factors of the nucleon.
1868: Here, we investigate the momentum-transfer dependencies of some
1869: form factors of the nucleon.
1870: For the reason explained before,
1871: all the physical predictions given hereafter will be 
1872: obtained with use of the mass parameter $M = 400 \,\mbox{MeV}$
1873: and $m_{\pi} = 100 \,\mbox{MeV}$. 
1874: We show in Fig.\ref{fig:GEis1020}, Fig.\ref{fig:GEiv1020},
1875: Fig.\ref{fig:GMis1020}, Fig.\ref{fig:GMiv1020} the predicted
1876: momentum-transfer dependencies of the generalized Sachs form factors,
1877: $G_{E, 10}^{(I = 0)} (t)$, $G_{E, 20}^{(I = 0)} (t)$,
1878: $G_{E, 10}^{(I = 1)} (t)$, $G_{E, 20}^{(I = 1)} (t)$,
1879: $G_{M, 10}^{(I = 0)} (t)$, $G_{M, 20}^{(I = 0)} (t)$,
1880: $G_{M, 10}^{(I = 1)} (t)$, and $G_{M, 20}^{(I = 1)} (t)$.
1881: 
1882: %Shown in Fig.\ref{fig:GEis1020} are
1883: %the momentum-transfer dependencies of the isoscalar 
1884: %generalized electric form factors $G_{E, 10}^{(I = 0)} (t)$ and 
1885: %$G_{E, 20}^{(I = 0)} (t)$.
1886: %One sees that the isoscalar Sachs form factor $G_{E, 10}^{(I = 0)} (t)$ 
1887: %is saturated by the valence quark contribution only, and there is no 
1888: %contribution from the polarized Dirac-sea quarks. 
1889: %One also confirms that, in the forward limit $t = 0$, the quark 
1890: %number sum rule (\ref{eq:udpSR}) is satisfied.
1891: %Turning to the generalized electric form factor $G_{E, 20}^{(I = 0)} (t)$, the %Dirac-sea contribution turns out to be around $25\%$ in the forward 
1892: %limit. As the momentum transfer increases, this Dirac-sea contribution 
1893: %decreases faster that the valence quark contribution, and beyond 
1894: %$|t| \simeq 0.4 \,\mbox{GeV}^2$, it becomes to play a minor role.
1895: %One can also convince that, at 
1896: %zero momentum transfer, the sum rule (\ref{eq:momSR}) is satisfied,
1897: %which means that the total nucleon momentum is saturated by the quark
1898: %field alone in the CQSM.
1899: 
1900: \vspace{4mm}
1901: \begin{figure}[htb] \centering
1902: \begin{center}
1903:  \includegraphics[width=16.0cm,height=7.5cm]{GEis1020.eps}
1904: \end{center}
1905: \vspace*{-0.5cm}
1906: \renewcommand{\baselinestretch}{1.20}
1907: \caption{The predictions of the CQSM for the isoscalar generalized electric
1908: form factors $G^{(I=0)}_{E,10}(t)$ and $G^{(I=0)}_{E,20}(t)$.}
1909: \label{fig:GEis1020}
1910: \end{figure}%
1911: 
1912: %Next, we discuss the isovector generalized electric form factors shown 
1913: %in Fig.\ref{fig:GEiv1020}.
1914: %First, we observe that, in the forward limit, the Dirac-sea 
1915: %contribution to the isovector Sachs electric form factor 
1916: %$G_{E, 10}^{(I = 1)} (t)$ is around $30\%$. It decreases as the momentum 
1917: %transfer increases, and beyond $|t| \simeq 0.4 \,\mbox{GeV}^2$
1918: %it becomes negligible as 
1919: %compared with the valence quark contribution. In the forward limit, the 
1920: %sum rule (\ref{eq:udmSR}) holds,
1921: %which assures that the number difference of the 
1922: %$u$-quark and the $d$-quark in the proton is just unity.
1923: %The isovector generalized electric form factor $G_{E, 20}^{(I = 1)} (t)$ 
1924: %is shown in the right panel of Fig.\ref{fig:GEiv1020}.
1925: %It is interesting to see that this 
1926: %form factors is nearly saturated by the contribution of the Dirac-sea 
1927: %quark, while the contribution of the Dirac-sea quarks vanishes almost
1928: %perfectly. In the particular case of zero-momentum transfer, we have
1929: %already explained the possible cause of this peculiar observation.
1930: %It is due to the approximate even-function nature of the 
1931: %Dirac-sea contribution to $f^{u - d} (x)$ as a function of $x$.
1932: %The result shown in Fig.\ref{fig:GEiv1020} then indicates that the
1933: %Dirac-sea contribution to $H_E^{u - d} (x, 0, t)$ remains to be an
1934: %approximate even-function of $x$ irrespective of the value
1935: %of the momentum transfer $t$. 
1936: 
1937: \vspace{4mm}
1938: \begin{figure}[htb] \centering
1939: \begin{center}
1940:  \includegraphics[width=16.0cm,height=7.5cm]{GEiv1020.eps}
1941: \end{center}
1942: \vspace*{-0.5cm}
1943: \renewcommand{\baselinestretch}{1.20}
1944: \caption{The predictions of the CQSM for the isovector generalized electric
1945: form factors $G^{(I=1)}_{E,10}(t)$ and $G^{(I=1)}_{E,20}(t)$.}
1946: \label{fig:GEiv1020}
1947: \end{figure}%
1948: 
1949: %Next, we turn our discussion to the generalized magnetic form factors.
1950: %Shown in Fig.\ref{fig:GMis1020} are the isoscalar magnetic form factor 
1951: %$G_{M, 10}^{(I = 0)} (t)$ and the corresponding generalized magnetic
1952: %(gravitomagnetic) form factors $G_{M, 20}^{(I = 0)} (t)$.
1953: %First, one finds that $G_{M, 10}^{(I = 0)} (t)$ is nearly saturated
1954: %by the valence quark contribution and the Dirac sea quark contribution
1955: %is almost negligible.
1956: %On the other hand, the contribution of the sea quarks to 
1957: %$G_{M, 20}^{(I = 0)} (t)$ turns out to be sizable. Since the reason of 
1958: %this observation has already been explained, we do not repeat it here.
1959: 
1960: %\vspace{4mm}
1961: \begin{figure}[htb] \centering
1962: \begin{center}
1963:  \includegraphics[width=16.0cm,height=7.5cm]{GMis1020.eps}
1964: \end{center}
1965: \vspace*{-0.5cm}
1966: \renewcommand{\baselinestretch}{1.20}
1967: \caption{The predictions of the CQSM for the isoscalar generalized
1968: magnetic form factors $G^{(I=0)}_{M,10}(t)$ and $G^{(I=0)}_{M,20}(t)$.}
1969: \label{fig:GMis1020}
1970: \end{figure}%
1971: 
1972: %Next, we show in Fig.\ref{fig:GMiv1020} the isovector magnetic form
1973: %factor $G_{M, 10}^{(I = 1)} (t)$ and the corresponding generalized
1974: %form factor $G_{M, 20}^{(I = 1)} (t)$.
1975: %We find that the sea quark contribution to 
1976: %$G_{M, 10}^{(I = 1)} (t)$ is around $42\%$ in the forward limit,
1977: %while the contributions to $G_{M, 20}^{(I = 1)} (t)$ is close to zero.
1978: %The reason of this difference can again be traced back to the behavior
1979: %of the corresponding generalized parton distribution function 
1980: %$E_M^{(I = 1)} (x, 0, t)$ as a function of $x$.
1981: 
1982: %\vspace{4mm}
1983: \begin{figure}[htb] \centering
1984: \begin{center}
1985:  \includegraphics[width=16.0cm,height=7.5cm]{GMiv1020.eps}
1986: \end{center}
1987: \vspace*{-0.5cm}
1988: \renewcommand{\baselinestretch}{1.20}
1989: \caption{The predictions of the CQSM for the isovector generalized magnetic
1990: form factors $G^{(I=1)}_{M,10}(t)$ and $G^{(I=1)}_{M,20}(t)$.}
1991: \label{fig:GMiv1020}
1992: \end{figure}%
1993: 
1994: To get some feeling about the momentum-transfer dependencies of the 
1995: predicted form factors, we shall compare them with the existing empirical
1996: data.
1997: At present, only the lowest moment of the GPDs, i.e. the standard 
1998: electromagnetic form factors of the nucleon are experimentally known.
1999: Shown in Fig.\ref{fig:F1pn} are the predictions of the CQSM for the
2000: Dirac form factors of the proton and the neutron,
2001: in comparison with the empirical data \cite{Arnold86}
2002: together with the corresponding predictions at the LHPC lattice
2003: simulation \cite{LHPC04},\cite{LHPC05}.
2004: One observes that the $t$-dependence of the CQSM prediction
2005: for $F_1^p (t)$ is a little too stronger than the empirical one,
2006: while the t-dependence of the lattice predictions is too much weaker
2007: than the empirical one.
2008: A little too fast falloff of the CQSM predictions means that it 
2009: slightly overestimates the electromagnetic size of the proton.
2010: On the contrary, the electromagnetic proton size predicted by the LHPC 
2011: simulation is too small as compared with the empirically known size.
2012: As is well known, the Dirac form factor of the neutron is not well 
2013: determined experimentally. Both of the CQSM prediction and the lattice 
2014: QCD predictions are very small in magnitude in qualitatively consistent 
2015: with the empirical information, although the former is slightly
2016: positive, while the latter is slightly negative.
2017: 
2018: \vspace{4mm}
2019: \begin{figure}[htb] \centering
2020: \begin{center}
2021:  \includegraphics[width=16.0cm,height=7.5cm]{F1pn.eps}
2022: \end{center}
2023: \vspace*{-0.5cm}
2024: \renewcommand{\baselinestretch}{1.20}
2025: \caption{The prediction of the CQSM for the Dirac form factors
2026: of the proton and the neutron in comparison with the empirical data
2027: \cite{Arnold86} together with the LHPC lattice predictions.
2028: The LHPC lattice predictions are based on their dipole fits
2029: \cite{LHPC04},\cite{LHPC05}.}
2030: \label{fig:F1pn}
2031: \end{figure}%
2032: 
2033: Fig.\ref{fig:F2pn} shows the predictions of the 
2034: CQSM for the Pauli form factors of the proton and the neutron, in 
2035: comparison with the empirical data together with the corresponding 
2036: predictions of the LHPC lattice simulation. (Here, both of the CQSM
2037: predictions and the lattice QCD predictions are normalized to the
2038: observed anomalous magnetic moments of the proton and the neutron
2039: at the zero momentum transfer.)
2040: The solid curves represent the predictions of the CQSM, whereas the
2041: dashed curves show the corresponding lattice predictions.
2042: One can see that the 
2043: predictions of the CQSM reproduce the empirical Pauli form factors
2044: fairly well.
2045: On the other hand, the lattice QCD predicts too slow falloff of the 
2046: Pauli form factors, which means that the magnetic sizes of the nucleon are
2047: largely underestimated by the lattice QCD simulations. The underestimate 
2048: of the nucleon electromagnetic sizes seems to be a general tendency of
2049: the lattice QCD simulations by the LHPC and QCDSF collaborations.
2050: It is not clear yet whether the origin of discrepancy can be traced
2051: back to the fact that these lattice simulations were carried out in
2052: the region with unrealistically heavy pion mass.
2053: 
2054: \vspace{4mm}
2055: \begin{figure}[htb] \centering
2056: \begin{center}
2057:  \includegraphics[width=10.0cm,height=7.5cm]{F2pn.eps}
2058: \end{center}
2059: \vspace*{-0.5cm}
2060: \renewcommand{\baselinestretch}{1.20}
2061: \caption{The predictions of the CQSM for the Pauli form factors
2062: of the proton and the neutron in comparison with the empirical data
2063: \cite{Arnold86} together with the LHPC lattice predictions based on
2064: the dipole fits \cite{LHPC04},\cite{LHPC05}.}
2065: \label{fig:F2pn}
2066: \end{figure}%
2067: 
2068: Concerning the genuine generalized form factors with $n \geq 2$,
2069: we have no experimental information yet. Among them, of our
2070: particular interest is the generalized form factor $B_{20}^{u+d}(t)$
2071: appearing in Ji's angular momentum sum rule. We show in 
2072: Fig.\ref{fig:B20t} the prediction of the CQSM for $B_{20}^{u+d}(t)$
2073: obtained with $M = 400 \,\mbox{MeV}$ and $m_\pi = 100 \,\mbox{MeV}$,
2074: in comparison with the corresponding predictions of the LHPC group.
2075: One sees that, for arbitrary value of $t$, $B_{20}^{u+d}(t)$ does not
2076: vanish in the CQSM, which is an indication of the fact that the shapes
2077: of the quark momentum and the total angular momentum distributions
2078: are not completely the same. Still, the magnitude of $B_{20}^{u+d}(t)$
2079: turns out to be very small, which seems qualitatively compatible
2080: with the prediction of the LHPC group, although one should not
2081: forget about large uncertainties in the lattice simulation at the
2082: present level.
2083: %Especially interesting here is
2084: %the value of $B_{20}^{u+d}(t)$ at $t=0$,
2085: %since, if it holds that $B_{20}^{u+d}(0) = 0$, the proportionality of
2086: %the total momentum fraction and the total angular momentum carries
2087: %by the quark fields is maintained.
2088: 
2089: \vspace{4mm}
2090: \begin{figure}[htb] \centering
2091: \begin{center}
2092:  \includegraphics[width=10.0cm,height=7.5cm]{budp.eps}
2093: \end{center}
2094: \vspace*{-0.5cm}
2095: \renewcommand{\baselinestretch}{1.20}
2096: \caption{The predictions of the CQSM for the generalized form factor
2097: $B_{20}^{u+d} (t)$ of the nucleon, in comparison with the
2098: corresponding prediction of the lattice QCD by LHPC
2099: collaboration \cite{LHPC05}.}
2100: \label{fig:B20t}
2101: \end{figure}%
2102: 
2103: We are now ready to discuss what we can say about the spin contents of
2104: the nucleon from our investigation on the generalized form factors of
2105: the nucleon.
2106: The quantities of our interest are all obtained from the forward limit of
2107: the generalized form factors, which are defined as the second moments
2108: of the relevant GPDs : 
2109: \begin{eqnarray}
2110:  \langle x \rangle^{u + d} &=& G_{E, 20}^{(I = 0)} (0), \ \ \ \ \ 
2111:  \langle x \rangle^{u - d} \ = \ G_{E, 20}^{(I = 1)} (0), \\
2112:  2 \,J^{u + d} &=& G_{M, 20}^{(I = 0)} (0), \ \ \ \ \ 
2113:  2 \,J^{u - d} \ = \ G_{M, 20}^{(I = 1)} (0) .
2114: \end{eqnarray}
2115: Summarized below are the predictions of the CQSM model for these quantities obtained with $M = 400 \,\mbox{MeV}$ and $m_{\pi} = 100 \,\mbox{MeV}$ :
2116: \begin{eqnarray}
2117:  \langle x \rangle^{u + d} &=& 1.00, \ \ \ \ \ 
2118:  \langle x \rangle^{u - d} = 0.276, \\
2119:  2 J^{u + d} &=& 1.00, \ \ \ \ \ 
2120:  2 J^{u - d} = 0.406 .
2121: \end{eqnarray}
2122: As pointed out before, these second moments of GPDs are generally scale 
2123: dependent. Our viewpoint is that the predictions of the CQSM correspond
2124: to those at the low energy scale where the validity of the model is
2125: ensured. (This energy may typically be characterized by the Pauli-Villars
2126: mass $\Lambda_1 \simeq 600 \,\mbox{MeV}$.) 
2127: We shall take account of the scale 
2128: dependence of the above quantities by solving the QCD evolution
2129: equation at the next-to-leading order (NLO) with the predictions of
2130: the CQSM as the initial conditions \cite{FRS77}--\cite{ABY97}. 
2131: For simplicity, let us assume that, at this low energy 
2132: scale, there is no contribution of gluon fields as well as of the
2133: strange quarks, which dictates that
2134: \begin{equation}
2135:  \langle x \rangle^s = 0.0, \ \ \ 2 \,J^s = 0.0, \ \ \ \ 
2136:  \langle x \rangle^g = 0.0, \ \ \ 2 \,J^g = 0.0 .
2137: \end{equation}
2138: The starting energy of the evolution is taken to be 
2139: $Q_{ini}^2 = 0.30 \,\mbox{GeV}^2 \simeq (550 \,\mbox{MeV})^2$,
2140: because it is favored from the previous successful application of
2141: the model to high energy deep-inelastic-scattering
2142: observables \cite{WK98}--\cite{Wakamatsu03b}.
2143: Taking $N_f = 3$ and $\Lambda_{QCD} = 0.248 \,\mbox{GeV}$,
2144: we find that, at $Q^2 = 4 \,\mbox{GeV}^2$, 
2145: \begin{eqnarray}
2146:  \langle x \rangle^{u + d + s} &=& 0.676, \ \ \ 
2147:  \langle x \rangle^{u - d} = 0.171, \ \ \ 
2148:  \langle x \rangle^g = 0.324, \\
2149:  2 J^{u + d + s} &=& 0.676, \ \ \ 
2150:  2 J^{u - d} = 0.257, \ \ \ \,
2151:  2 J^g = 0.324 .
2152: \end{eqnarray}
2153: One may notice that the values of $\langle x \rangle^{u+d+s}$ and
2154: $2 \,J^{u+d+s}$ at $Q^2 = 4 \,\mbox{GeV}^2$ precisely coincide.
2155: Actually, the equality of these two quantities holds at any energy
2156: scale. The reason is because these two quantities obey exactly the
2157: same evolution equation and because they are equal at the initial
2158: energy scale according to the CQSM \cite{JTH96}.
2159: (We emphasize that the latter is
2160: also the case for the LHPC and QCDSF lattice predictions at least
2161: approximately \cite{LHPC05},\cite{QCDSF04b}.)
2162: In table \ref{tabxandj}, we compare these predictions of the CQSM
2163: with those of the lattice QCD and also with the empirical values for 
2164: $\langle x \rangle^{u + d + s}$ and $\langle x \rangle^{u - d}$ obtained
2165: from the phenomenological PDF fits \cite{MRST2004}.
2166: As one sees, the total momentum
2167: fraction $\langle x \rangle^{u + d + s}$
2168: carried by the quarks decreases rapidly as 
2169: $Q^2$ increases. The evolved value $\langle x \rangle ^{u + d} 
2170: \simeq 0.68$ at $Q^2 = 4 \,\mbox{GeV}^2$ is not extremely far from the
2171: lattice QCD prediction at the same normalization scale,
2172: although it is a little larger than the empirical value
2173: $\langle x \rangle^{u + d + s}_{empirical} \simeq 0.57$. This difference 
2174: may be an indication of the fact that, even at the low energy scale
2175: around $Q^2 \simeq 0.30 \,\mbox{GeV}^2$,
2176: the gluons may carry some portion of the
2177: nucleon momentum. As far as the difference of the momentum fractions
2178: carried by the $u$-quark and the $d$-quark, the CQSM well reproduces
2179: the empirical value $\langle x \rangle^{u - d}_{empirical} \simeq 0.16$
2180: at $Q^2 = 4 \, \mbox{GeV}^2$, whereas the lattice QCD overestimates it
2181: a little.
2182: 
2183: \begin{table}[htbp]
2184: \begin{center}
2185: \renewcommand{\baselinestretch}{1.20}
2186: \caption{The predictions of the CQSM for ${\langle x \rangle}^{u+d}$,
2187: ${\langle x \rangle}^{u-d}$, $2 \,J^{u+d}$, and $2 \,J^{u-d}$
2188: in comparison with the predictions of the lattice QCD
2189: simulations \cite{LHPC04},\cite{LHPC05},\cite{QCDSF04b},\cite{QCDSF05}
2190: as well as with the empirical 
2191: information \cite{MRST2004}. \label{tabxandj}}
2192: \vspace{10mm}
2193: \begin{tabular}{cccccc} \hline
2194:  \, & \ \ CQSM (model scale) \ \ & \ \ 
2195:  CQSM \,($Q^2 = 4 \,\mbox{GeV}^2)$ \ \ & 
2196:  \ \ LHPC \ \ & \ \ QCDSF \ \ & \ \ empirical \ \ 
2197:  \\ \hline\hline
2198:  \ ${\langle x \rangle}^{u+d+s}$ \ & 1.000 & 0.676 & 0.61 & 0.59 & 0.57
2199:  \\ \hline
2200:  \ ${\langle x \rangle}^{u-d}$ \ & 0.276 & 0.171 & 0.269 & 0.24 & 0.157
2201:  \\ \hline
2202:  \ $2 \,J^{u+d+s}$ \ & 1.000 & 0.676 & 0.58 & 0.66 & --- 
2203:  \\ \hline
2204:  \ $2 \,J^{u-d}$ \ & 0.406 & 0.257 & 0.93 & 0.82 & --- 
2205:  \\ \hline
2206: \end{tabular}
2207: \end{center}
2208: \renewcommand{\baselinestretch}{1.20}
2209: \end{table}
2210: 
2211: Next we turn to the discussion of the total angular momenta
2212: $J^u$ and $J^d$ carried by the quark fields, on which we do not have
2213: any empirical information yet.
2214: One can see that, as far as the total angular momentum fraction 
2215: $J^{u + d + s}$ carried by the quark fields, is concerned, the prediction
2216: of the CQSM is qualitatively consistent with that of the lattice QCD.
2217: However, a big discrepancy is observed for the difference $J^{u - d}$
2218: of the angular momentum carried by the $u$-quark and the $d$-quark.
2219: The cause of this discrepancy can be traced back to that of the isovector
2220: gravitomagnetic moment of the nucleon $G_{M, 20}^{(I = 1)} (0)$,
2221: which we have already discussed.
2222: Fairly small prediction of the CQSM for $J^{u - d}$ also appears to be 
2223: incompatible with the semi-theoretical (or semi-phenomenological)
2224: estimate carried out in \cite{GPV01} with partial use of the predictions
2225: of the CQSM.
2226: As we shall discuss below, however, their estimate for
2227: $J^u$ and $J^d$ shown in Table 4 and Table 5 of \cite{GPV01}
2228: should be taken with care. In fact, it was
2229: obtained based on the valence-like approximation, i.e. by neglecting the
2230: sizable Dirac sea contributions to $E^u (x, 0, 0)$ and $E^d (x, 0, 0)$.
2231: To be more concrete, they start with a simple guess for these
2232: distributions functions as 
2233: \begin{eqnarray}
2234:  E^u (x, 0, 0) &=& \frac{1}{2} \,\kappa^u \,f^{u_{val}} (x), \\
2235:  E^d (x, 0, 0) &=& \ \ \kappa^d \,f^{d_{val}} (x) ,
2236: \end{eqnarray}
2237: with 
2238: \begin{eqnarray}
2239:  \kappa^u &=& 2 \,\kappa^p + \kappa^u = \ 1.673, \\
2240:  \kappa^d &=& \kappa^p + 2 \,\kappa^u = - \,2.033 .
2241: \end{eqnarray}
2242: This parameterization trivially satisfies the 1st moment sum rule
2243: \begin{equation}
2244:  \int_{-1}^1 \,E^q (x, 0, 0) \,d x = \kappa^q .
2245: \end{equation}
2246: On the other hand, by using the 2nd moment sum rule or Ji's angular 
2247: momentum sum rule, they obtain
2248: \begin{eqnarray}
2249:  J^u &=& \frac{1}{2} \,\left[\, {\langle x \rangle}^u + 
2250:  \kappa^u {\langle x \rangle}^{u_{val}} \right], \label{eq:Juval} \\
2251:  J^d &=& \frac{1}{2} \,\left[\, {\langle x \rangle}^d + 
2252:  \kappa^d {\langle x \rangle}^{d_{val}} \right] , \label{eq:Jdval}
2253: \end{eqnarray}
2254: where ${\langle x \rangle}^q$ is the momentum fraction carried by the
2255: quark of flavor $q$, while ${\langle x \rangle}^{q_{val}}$ is the
2256: corresponding contribution of the valence 
2257: quark in the sense of parton model.
2258: Using the MRST98 parameterization for the unpolarized
2259: PDFs \cite{MRST98}, they could thus obtain at $Q^2 \simeq 1 \,
2260: \mbox{GeV}^2$
2261: \begin{equation}
2262:  {\langle x \rangle}^{u_{val}} \simeq 0.34, \ \ \ 
2263:  {\langle x \rangle}^u \simeq 0.40, \ \ \ 
2264:  2 \,J^u \simeq 0.69 , \label{eq:xJu}
2265: \end{equation}
2266: and 
2267: \begin{equation}
2268:  {\langle x \rangle}^{d_{val}} \simeq 0.14, \ \ \ 
2269:  {\langle x \rangle}^d \simeq 0.22, \ \ \ 
2270:  2 \,J^d \simeq -0.07 , \label{eq:xJd}
2271: \end{equation}
2272: which appears to be qualitatively consistent with the lattice
2273: predictions.
2274: (Here, we have discarded the small contribution of the $s$-quark, for 
2275: simplicity.) However, after this simple estimate for $J^u$ and $J^d$,
2276: they next try to take account of the sizable Dirac sea contribution to
2277: the distributions $E^u (x, 0, 0)$ and $E^d (x, 0, 0)$.
2278: As already shown in our exact model calculation,
2279: and as shown in Fig.8 of \cite{GPV01},
2280: which is obtained based on the
2281: derivative-expansions-type approximation within the 
2282: CQSM, the Dirac sea contribution to $E^{u - d} (x, 0, 0)$ has a narrow
2283: and positive peak around $x \simeq 0$. To simulate this narrowly peaked 
2284: behavior of the Dirac-sea contribution, they propose to parameterize it
2285: by a $\delta$ function in $x$. The new and improved parameterization for
2286: $E^u (x, 0, 0)$ and $E^d (x, 0, 0)$ are then given as  
2287: \begin{eqnarray}
2288:  E^u (x, 0, 0) &=& A^u \,f^{u_{val}} (x) + B^u \,\delta (x) , \\
2289:  E^d (x, 0, 0) &=& A^d \,f^{d_{val}} (x) + B^d \,\delta (x) .
2290: \end{eqnarray}
2291: From the 1st and 2nd sum rules for $E^u (x, 0, 0)$ and $E^d (x, 0, 0)$, 
2292: they obtain
2293: \begin{eqnarray}
2294:  A^u &=& \frac{2 \,J^u - {\langle x \rangle}^u}
2295:  {{\langle x \rangle}^{u_{val}}}, \\ \label{eq:Au}
2296:  A^d &=& \frac{2 \,J^d - {\langle x \rangle}^d}
2297:  {{\langle x \rangle}^{d_{val}}}, \\
2298:  B^u &=& 2 \,\left[\, \frac{1}{2} \,\kappa^u \ - \ 
2299:  \frac{2 \,J^u - {\langle x \rangle}^u}
2300:  {{\langle x \rangle}^{u_{val}}} \,\right], \\
2301:  B^d &=& \kappa^d - \frac{2 J^d - {\langle x \rangle}^d}
2302:  {{\langle x \rangle}^{d_{val}}} . \label{eq:Bd}
2303: \end{eqnarray}
2304: As pointed out in \cite{GPV01}, the total angular momentum carried by
2305: $u$- and $d$-quarks, $J^u$ and $J^d$, now enter as fit parameters in the
2306: parameterization of $E^u (x, 0, 0)$ and $E^d (x, 0, 0)$.
2307: If so, there is no compelling reason to believe that they are close to 
2308: the estimate given in (\ref{eq:Juval}) and (\ref{eq:Jdval}),
2309: obtained within the valence-type 
2310: parameterization for $E^u (x, 0, 0)$ and $E^d (x, 0, 0)$.
2311: In fact, if one puts the estimate given in (\ref{eq:xJu}) and
2312: (\ref{eq:xJd}) into the above relations (\ref{eq:Au}) 
2313: $\sim$ (\ref{eq:Bd}), one obtains
2314: \begin{equation}
2315:  A^u \simeq 0.853, \ \ \ 
2316:  A^d \simeq -2.071, \ \ \ 
2317:  B^u \simeq -0.033, \ \ \ 
2318:  B^d \simeq 0.038 ,
2319: \end{equation}
2320: which dictates that the coefficients of the $\delta$ functions are small
2321: or nearly zero. This is only natural, since the used value of $J^u$ and
2322: $J^d$ are just estimated based on the  valence-type parameterization for
2323: $E^u (x, 0, 0)$ and $E^d (x, 0, 0)$.
2324: Conversely speaking, the values of $J^u$ and $J^d$ 
2325: quoted in Table 4 and table 5 of \cite{GPV01} need a revision, because
2326: they are incompatible with the existence of sharp peak of 
2327: $E^{u - d} (x, 0, 0)$ with sizable positive magnitude around
2328: $x \simeq 0$ observed in Fig.8 of the same paper. 
2329: The inseparable relation between the magnitude of $J^u - J^d$ and the 
2330: sharp of $E^{u - d} (x, 0, 0)$ can be  made more transparent by slightly 
2331: modifying their schematic analysis.
2332: Instead of $E^{u - d} (x, 0, 0)$, we chose here to parameterize
2333: $E_M^{u - d} (x, 0, 0) \equiv H^{u - d} (x, 0, 0) + E^{u - d} (x, 0, 0)$ 
2334: as 
2335: \begin{equation}
2336:  E_M^{u - d} (x, 0, 0) \ = \ c_1 \,
2337:  \left[\,f^{u_{val}} (x) - f^{d_{val}} (x) \,\right]
2338:  \ + \ c_2 \,\delta (x) .
2339: \end{equation}
2340: Here the $\delta$ function term is thought to simulate the sizable sharp 
2341: peak of $E_M^{u - d} (x, 0, 0)$ predicted by the CQSM. Actually, it need 
2342: not be a $\delta$ function. It can be  an any function $g$ of $x$, 
2343: as far as it satisfies the following two conditions :
2344: 
2345: \begin{itemize}
2346: \item $g(x)$ is an even function of $x$ at least approximately,
2347: \item the integral of $g (x)$ over $x$ gives a (positive) number $c_2$.
2348: \end{itemize}
2349: 
2350: Assuming that these conditions are satisfied (as is the case for the 
2351: predictions of the CQSM), the 1st and the 2nd moment sum rule of 
2352: $E^{u - d}_M (x, 0, 0)$ leads to the identities :
2353: \begin{eqnarray}
2354:  1 + \kappa^u - \kappa^d &=& \int_{-1}^1 E_M^{u - d} (x, 0, 0) \,d x 
2355:  \ = \ c_1 \ + \ c_2 , \\
2356:  2 \,(J^u - J^d) &=& \int_{-1}^1 \,E_M^{u - d} (x, 0, 0) \,d x
2357:  \ = \ c_1 \,\left(\, \langle x \rangle^{u_{val}} - 
2358:  \langle x \rangle^{d_{val}} \,\right) .
2359: \end{eqnarray}
2360: Combining these relations, we have
2361: \begin{equation}
2362:  2 \,(J^u - J^d) = \frac{1 + \kappa^u - \kappa^d}{1 + r} \,
2363:  \left(\,\langle x \rangle^{u_{val}} - 
2364:  \langle x \rangle^{d_{val}} \,\right) . \label{eq:Judm}
2365: \end{equation}
2366: Here we have introduced a parameter $r \equiv c_2 / c_1$.
2367: This relation shows that, except for one parameter $r$, the difference 
2368: of the total angular momenta, carried by the $u$- and $d$-quarks,
2369: $J^u - J^d$, is given by $\kappa^u, \kappa^d, \langle x 
2370: \rangle^{u_{val}}$ and $\langle x \rangle^{d_{val}}$, which are all
2371: observables. What is the physical meaning of the parameter $r$, then ?
2372: To understand it, 
2373: we first recall that the quantity $E_M^{u - d} (x, 0, 0)$ represents the 
2374: {\it distribution of the isovector magnetic moment} of the nucleon
2375: {\it in Feynman $x$-space} not in ordinary coordinate space.
2376: The contribution of its valence-like 
2377: distribution to the magnetic moment gives $c_1$, whereas that of its 
2378: sea-like distribution gives $c_2$. The CQSM indicates that they are of
2379: approximately equal magnitude, i.e. $c_1 \simeq c_2$, which means that 
2380: $r \simeq 1$. On the other hand, roughly speaking, the lattice
2381: simulations carried out in the heavy pion region with neglect of the
2382: disconnected diagrams correspond to $r \simeq 0$.
2383: 
2384: \begin{table}[htbp]
2385: \begin{center}
2386: \renewcommand{\baselinestretch}{1.20}
2387: \caption{The value of $2 \,(J^u - J^d)$ as a functions of the ratio
2388: of the parameters $c_2$ and $c_1$ defined in (\ref{eq:Judm}).
2389: \label{tabju-jd}}
2390: \vspace{10mm}
2391: \begin{tabular}{cc} \hline
2392:  \ \ \ \ \ \ $r \equiv c_2 / c_1$ \ \ \ \ \ \ & 
2393:  \ \ \ \ \ \ $2 \,(J^u - J^d)$ \ \ \ \ \ \ 
2394:  \\ \hline\hline
2395:  0 & 0.941 \\ \hline
2396:  0.5 & 0.627 \\ \hline
2397:  1 & 0.471 \\ \hline
2398: \end{tabular}
2399: \end{center}
2400: \renewcommand{\baselinestretch}{1.20}
2401: \end{table}
2402: 
2403: Table \ref{tabju-jd} show the values of $2 (J^u - J^d)$ obtained
2404: from (\ref{eq:Judm}) for several typical values of the ratio $r$.
2405: The quantity $\langle x \rangle^{u_{val}} - \langle x \rangle^{d_{val}}$
2406: are actually scale dependent. For simplicity, here we have used the 
2407: empirical value $\langle x \rangle^{u_{val}} - \langle x 
2408: \rangle^{d_{val}} \simeq 0.2$ corresponding to $Q^2 = 1 \,
2409: \mbox{GeV}^2$ quoted in \cite{GPV01}.
2410: One sees that the two cases, i.e. $r = 0$ case and $r = 1$ case,
2411: lead to a factor of two 
2412: difference for $J^u - J^d$. What is indicated by this observation is an 
2413: inseparable connection between the {\it angular momentum carried by the
2414: quark fields} in the nucleon and the magnetic moment of the nucleon,
2415: or more precisely the {\it distribution of magnetic moments in
2416: Feynman $x$-space}.
2417: As a matter of course, the relation between the quark angular momentum
2418: and the nucleon magnetic moment could be anticipated 
2419: from more general ground of Ji's angular momentum sum rule. 
2420: However, an advantage of our explicit model analysis is that we can get 
2421: more deep and concrete insight into the possible behavior of the
2422: relevant distribution $E^{u - d} (x, 0, 0)$,
2423: on which we have no experimental information yet.
2424: 
2425: Returning to the isoscalar (flavor-singlet) combination or the net quark 
2426: contribution to the total angular momentum of the nucleon, $J^{u+d+s}$, 
2427: we have pointed out that the prediction of the CQSM for it is not so far 
2428: from that of the lattice QCD. However, both of LHPC and QCDSF collaborations also estimated the net orbital angular momentum carried by the quark
2429: fields, thereby being led to the conclusion that the {\it total orbital
2430: angular momentum of the quarks} is very {\it small} or consistent
2431: with {\it zero}.
2432: As pointed out in our recent paper, this conclusion contradicts not only
2433: the prediction of the CQSM but also the famous EMC observation.
2434: Although the possible reason of this discrepancy 
2435: was already pointed out in that paper \cite{Wakamatsu05},
2436: here we discuss it in more detail,
2437: especially by taking care of the scale dependencies of the relevant 
2438: observables.
2439: The quark orbital angular momentum can be obtained by subtracting the 
2440: intrinsic quark spin term from the total quark angular momentum 
2441: $J^{u+d}$ as
2442: \begin{equation}
2443:  L^{u + d} = J^{u + d} - \frac{1}{2} \,\Delta \Sigma^{u + d} ,
2444: \end{equation}
2445: where $J^{u + d}$ is given by
2446: \begin{equation}
2447:  J^{u + d} = \frac{1}{2} \,\left( \,
2448:  \langle x \rangle^{u + d} + B_{20}^{u + d} (0) \right) ,
2449: \end{equation}
2450: Using the results of the dipole fits for the generalized form
2451: factors, the LHPC collaboration obtain
2452: \begin{eqnarray}
2453:  \langle x \rangle^{u+d} &=& A_{20}^{u+d} (0) \ = \ 0.666 \pm 0.009, \\
2454:  B_{20}^u (0) &=& 0.29 \pm 0.04, \ \ \ 
2455:  B_{20}^d (0) \ = \ - \,0.38 \pm 0.02, \\
2456:  \Delta \Sigma^u &=& 0.860 \pm 0.069, \ \ \ 
2457:  \Delta \Sigma^d \ = \ - \,0.171 \pm 0.043,
2458: \end{eqnarray}
2459: which in turn gives
2460: \begin{equation}
2461:  L^u \ = \ - \,0.088 \pm 0.019, \ \ \ 
2462:  L^d \ = \ 0.036 \pm 0.013 ,
2463: \end{equation}
2464: or
2465: \begin{equation}
2466:  L^{u+d} \ = \ - \,0.052 \pm 0.019 .
2467: \end{equation}
2468: On the other hand, the QCDSF collaboration obtain
2469: \begin{eqnarray}
2470:  \langle x \rangle^{u} &=& A_{20}^{u} (0) \ = \ 0.400 \pm 0.022, \ \ \
2471:  \langle x \rangle^{d} \ = \ A_{20}^{d} (0) \ = \ 0.147 \pm 0.011, \\
2472:  B_{20}^u (0) &=& 0.334 \pm 0.113, \ \ \ 
2473:  B_{20}^d (0) \ = \ - \,0.232 \pm 0.077, \\
2474:  \Delta \Sigma^u &=& 0.84 \pm 0.02, \ \ \ 
2475:  \Delta \Sigma^d \ = \ - \,0.24 \pm 0.02,
2476: \end{eqnarray}
2477: which gives
2478: \begin{equation}
2479:  L^{u+d} \ = \ 0.03 \pm 0.07 .
2480: \end{equation}
2481: As one sees, a common conclusion of the two groups is that the total 
2482: orbital angular momentum of quarks is very small or consistent
2483: with zero.
2484: 
2485: Since these lattice predictions corresponds to the energy scale of 
2486: $Q^2 = 4 \,\mbox{GeV}^2$ in the $\overline{MS}$ scheme, we try to
2487: evolve the corresponding predictions of the CQSM to the same energy.
2488: To know the scale dependence of the total quark orbital angular
2489: momentum $L^Q$ with $Q$ denoting the sum of all quark flavors,
2490: we need to know the scale dependence of the
2491: total quark angular momentum $J^Q$ and that of the total quark
2492: longitudinal polarization $\Delta \Sigma^Q$.
2493: We recall again the fact that the angular momentum fractions 
2494: carried by the quark and the gluon field, i.e. $J^Q$ and $J^g$,
2495: obey exactly the same evolution equation as the total momentum
2496: fractions carried by the quark and gluon fields
2497: $\langle x \rangle^Q$ and $\langle x \rangle^g$ \cite{JTH96}.
2498: The evolution equation of $\Delta \Sigma^Q$,
2499: which is coupled with the evolution of the gluon polarization, is
2500: also well known \cite{GRSV96}.
2501: As initial conditions of the evolution, we use the predictions of the 
2502: CQSM (the flavor SU(2) version) :
2503: \begin{eqnarray}
2504:  2 J^{u + d} = 1.0,
2505: % \mbox{at} \ \ Q^2_{ini} = 0.30 \,\mbox{GeV}^2 .
2506: \end{eqnarray}
2507: supplemented with the assumption
2508: \begin{eqnarray}
2509:  \ \ 2 \,J^s = 0.0, \ \ \ \Delta \Sigma^s = 0.0, \ \ \ \Delta g = 0.0,
2510: % \mbox{at} \ \ Q^2_{ini} = 0.30 \,\mbox{GeV}^2 .
2511: \end{eqnarray}
2512: at $Q^2_{ini} = 0.30 \,\mbox{GeV}^2$. (There also exists the
2513: flavor SU(3) version of the CQSM \cite{Wakamatsu03a},\cite{Wakamatsu03b}.
2514: It predicts that $\Delta \Sigma^s$
2515: is a negative quantity of the order of $(5 \sim 10) \,\%$.
2516: However, the flavor singlet combination or the net quark contribution
2517: to the total quark longitudinal polarization $\Delta \Sigma^Q$
2518: takes almost the same value in both version
2519: of the CQSM, so that the following discussion will receive no
2520: modification.)
2521: 
2522: \vspace{3mm}
2523: \begin{figure}[htb] \centering
2524: \begin{center}
2525:  \includegraphics[width=16.0cm,height=7.0cm]{JDelSigevol.eps}
2526: \end{center}
2527: \vspace*{-0.5cm}
2528: \renewcommand{\baselinestretch}{1.20}
2529: \caption{The left panel shows the scale dependencies of the
2530: quark and gluon total angular momenta $J^Q$ and $J^g$ obtained by
2531: solving the evolution equation at the NLO with the predictions of
2532: the CQSM as initial conditions, while the right panel
2533: shows the scale dependence of $\Delta \Sigma^Q$ and $\Delta g$.}
2534: \label{fig:JDelSigevol}
2535: \end{figure}%
2536: 
2537: The left panel of Fig.\ref{fig:JDelSigevol}
2538: shows the scale dependence of $J^Q$ obtained by solving the
2539: evolution equation at the NLO in the $\overline{M S}$ 
2540: scheme. (Here we set $N_c = 3$ and $\Lambda_{QCD} = 0.248 \,\mbox{GeV}$.)
2541: One observers that, especially at low energy scales, $J^Q$ is a
2542: rapidly decreasing function, while $J^g$ is a rapidly increasing function
2543: of $Q^2$.
2544: The right panel of Fig.\ref{fig:JDelSigevol} shows the scale dependence
2545: of $\Delta \Sigma^Q$ and $\Delta g$ obtained by solving the NLO
2546: evolution equation in the same renormalization scheme \cite{GRSV96}.
2547: As one sees, $\Delta g$ is a rapidly-increasing function of $Q^2$.
2548: On the other hand, $\Delta \Sigma^{u+d+s}$ has a fairly weak scale
2549: dependence. Its scale dependence is restricted to the very low
2550: energy region below $Q^2 \leq 0.6\,\mbox{GeV}^2$ and beyond that scale
2551: it changes very slowly. (We recall that, at the leading order (LO),
2552: $\Delta \Sigma^Q$ in the $\overline{MS}$ is exactly scale
2553: independent.
2554: One may also remember the fact that $\Delta \Sigma^Q$ in the
2555: chiral-invariant renormalization scheme is scale independent by
2556: definition \cite{LSS2000}--\cite{ABFS97}.) 
2557: Now, combining the results for the scale dependencies
2558: of $J^Q, J^g, \Delta \Sigma^Q$, and $\Delta g$, one can
2559: predict the scale dependencies of $L^Q$ and $L^g$ at the NLO
2560: from
2561: \begin{eqnarray}
2562:  L^Q &=& J^Q - \frac{1}{2} \,\Delta \Sigma^Q, \\
2563:  L^g &=& J^g - \Delta g .
2564: \end{eqnarray}
2565: Fig \ref{fig:Lqgevol} shows the scale dependencies of quark and
2566: gluon orbital angular momenta obtained in this manner.
2567: One sees that the total quark orbital angular momentum is a rapidly
2568: decreasing functions of the energy scale, especially at the low energy
2569: scales. Since the longitudinal quark polarization is only weakly scale
2570: dependent, this feature comes from the scale dependence of
2571: the total quark angular momentum, which has the same scale
2572: dependence as the total quark momentum fraction. One also sees
2573: that the gluon orbital angular momentum is a decreasing
2574: function of the energy scale.
2575: 
2576: 
2577: \vspace{6mm}
2578: \begin{figure}[htb] \centering
2579: \begin{center}
2580:  \includegraphics[width=9.0cm,height=7.0cm]{Lqgevol.eps}
2581: \end{center}
2582: \vspace*{-0.5cm}
2583: \renewcommand{\baselinestretch}{1.20}
2584: \caption{The scale dependencies of the quark and gluon orbital
2585: angular momenta $L^Q$ and $L^g$ at the NLO obtained with
2586: combined use of the predictions of the CQSM and the QCD evolution
2587: equations at the NLO.}
2588: \label{fig:Lqgevol}
2589: \end{figure}%
2590: 
2591: For the sake of comparison with the lattice QCD predictions
2592: corresponding the the energy scale of $Q^2 = 4 \,\mbox{GeV}^2$,
2593: we summarize in Table \ref{tabspincontents} the predictions of
2594: the CQSM for the nucleon spin contents at the same energy scale.
2595: One confirms that the total quark orbital angular momentum is a rapidly
2596: decreasing function of the energy scale and 
2597: its value at $Q^2 = 4 \,\mbox{GeV}^2$ is nearly half of that at the
2598: low energy model scale around $Q_{ini}^2 = 0.30 \,\mbox{GeV}^2$.
2599: Nonetheless, it still bears a sizable amount of the total nucleon spin
2600: even at the scale $Q^2 = 4 \,\mbox{GeV}^2$, in contrast to the
2601: lattice predictions. 
2602: We point out that, after taking account of the scale dependence, the 
2603: predictions for $J^Q$ are not so different between the CQSM and the 
2604: lattice QCD. What is remarkably different is the predictions of the two 
2605: theories for the net longitudinal quark polarization or the contribution 
2606: of intrinsic quark spin. It is clear that the lattice QCD simulations by 
2607: the LHPC and the QCDSF collaborations for $\Delta \Sigma^Q$ 
2608: considerably overestimate the empirically known value of 
2609: $\Delta \Sigma^Q$, which is known to be quite small as  
2610: \begin{equation}
2611:  \Delta \Sigma^Q_{empirical} = (0.2 \sim 0.35),
2612: \end{equation}
2613: while the prediction of the CQSM is qualitatively consistent with this 
2614: empirical information. 
2615: A plausible reason why the lattice simulations
2616: by the LHPC and the QCDSF collaboration predicts fairly large 
2617: $\Delta \Sigma^Q$ around $0.6$ was pointed out in \cite{WT05}.
2618: In that paper, we investigated the pion mass dependence of
2619: $\Delta \Sigma^Q$ within the framework of the CQSM and found that
2620: it is very sensitive to the variation of $m_{\pi}$, especially in the
2621: region close to the chiral limit $m_{\pi} = 0$.
2622: (The magnitude of $\Delta \Sigma^Q$ decreases rapidly as $m_\pi$
2623: approaches $0$.)
2624: This indicates that the lattice estimates carried out in 
2625: the heavy pion region around $m_{\pi} = (700 \sim 900) \,\mbox{MeV}$ may
2626: not give reliable prediction for the particular observable
2627: $\Delta \Sigma^Q$. As a consequence,
2628: their conclusion that the orbital angular momentum carried 
2629: by the quark fields in the nucleon is negligible, must also be taken
2630: with care. It may be justified in the heavy pion world, but whether
2631: it is also the case in our chiral world is a different question,
2632: which should be answered by the lattice QCD studies in the future.
2633: 
2634: Also noteworthy is the fact that the large values of
2635: $\Delta \Sigma^{u+d}$ obtained by the LHPC and QCDSF lattice
2636: collaborations seem to contradict the results of earlier lattice QCD
2637: studies \cite{MDLMM00}--\cite{DLL95}, which predict fairly small
2638: $\Delta \Sigma^Q$ around $(0.2 \sim 0.3)$.
2639: Among them, Mathur et. al. \cite{MDLMM00} also estimated the total
2640: quark angular momentum $J^Q$ from the quark energy-momentum tensor
2641: form factors on the lattice with the quenched approximation,
2642: and deduced that the quark orbital angular momentum carries about
2643: $34 \,\%$
2644: of the total proton spin, which is compatible or even dominant
2645: over the contribution of intrinsic quark spin around $26 \,\%$
2646: obtained by their simulation.
2647: 
2648: \begin{table}[htbp]
2649: \begin{center}
2650: \renewcommand{\baselinestretch}{1.20}
2651: \caption{The predictions of the CQSM for the spin contents
2652: of the nucleon in comparison with the corresponding predictions of
2653: the LHPC and QCDSF lattice QCD simulations \cite{LHPC05},\cite{QCDSF04b}.
2654: \label{tabspincontents}}
2655: \vspace{10mm}
2656: \begin{tabular}{ccccc} \hline
2657:  \, & \ \ CQSM (model scale) \ \ & \ \ 
2658:  CQSM \,($Q^2 = 4 \,\mbox{GeV}^2)$ \ \ & 
2659:  \ \ LHPC \ \ & \ \ QCDSF \ \ \\ \hline\hline
2660:  \ $ 2 \,J^{u+d+s}$ \ & 1.000 & 0.676 & 0.56 & 0.66
2661:  \\ \hline
2662:  \ $\Delta \Sigma^{u+d+s}$ \ & 0.350 & 0.318 & 0.69 & 0.60
2663:  \\ \hline
2664:  \ $2 \,L^{u+d+s}$ \ & 0.650 & 0.358 & -0.11 & 0.06
2665:  \\ \hline
2666: \end{tabular}
2667: \end{center}
2668: \renewcommand{\baselinestretch}{1.20}
2669: \end{table}
2670: 
2671: So far, we have given several interesting theoretical predictions
2672: on the basis of an effective theory of QCD, i.e. the CQSM.
2673: Although we believe that the reliability of these predictions
2674: is guaranteed by the phenomenological success of the CQSM achieved
2675: in the nucleon structure function physics, it would be nicer if
2676: one can extract some predictions, which do not depend on a specific
2677: model of the nucleon. It is in fact possible, if one accepts the
2678: following two theoretical postulates. They are
2679: \begin{itemize}
2680:  \item Ji's angular momentum sum rule \ : \ 
2681:  $J^{Q} = \frac{1}{2} \,[ \langle x \rangle^{Q} + 
2682:  B_{20}^{Q} (0) ]$ ,
2683:  \item absence of the {\it net quark contribution} to the
2684:  {\it anomalous gravitomagnetic moment} of the nucleon \ : \
2685:  $B_{20}^{Q} (0) = 0$ .
2686: \end{itemize}
2687: It is reasonable to accept the first postulate. Otherwise, we would 
2688: lose a only clue to experimentally access the quark angular momentum in
2689: the nucleon. What is crucial in the following argument is therefore
2690: the second postulate. As already mentioned, the identity
2691: $B_{20}^Q (0) = 0$, that holds within the CQSM, just follows from the
2692: total momentum and the total angular momentum sum rules, both of
2693: which are saturated by the quark fields alone in this effective quark
2694: theory. It can therefore be an artifact of the model.
2695: However, an interesting observation is that the smallness of
2696: $B_{20}^Q (0)$ is also the predictions of the LHPC and the QCDSF
2697: lattice QCD simulations, which take account of full quark-gluon dynamics.
2698: Naturally, one should not forget about large uncertainties in the
2699: lattice simulations at the present stage. One should also worry about
2700: the $m_\pi$-dependence of $B_{20}^Q (0)$, although we conjecture from
2701: our analyses in the CQSM a weak $m_\pi$-dependence of this quantity.
2702: 
2703: %It is in fact possible, if one accepts only one theoretical postulate,
2704: %i.e. the {\it absence of the net quark contribution to the
2705: %gravitomagnetic moment of the nucleon}, indicated by the LHPC and
2706: %the QCDSF collaborations.
2707: 
2708: Here, we shall proceed by assuming that $B_{20}^Q (0)$ vanishes
2709: exactly or at least very small. As already pointed out, the identity
2710: $B_{20}^Q (0)$ leads to remarkable relations, i.e. the proportionality
2711: of the total momentum and total angular momentum carried by the quark
2712: fields and also by the gluon fields as
2713: \begin{equation}
2714:  J^Q = \frac{1}{2} \,\langle x \rangle^Q, \ \ \ 
2715:  J^g = \frac{1}{2} \,\langle x \rangle^g .
2716: \end{equation}
2717: An important fact here is that the quark and gluon momentum
2718: fractions, i.e. $\langle x \rangle^Q$ and $\langle x \rangle^g$
2719: are empirically known with fairly good precision.
2720: For instance, the two popular PDF fits, i.e. MRST2004 and CTEQ5,
2721: give almost the same answer for $\langle x \rangle^Q$ and
2722: $\langle x \rangle^g$ at least within the energy range
2723: $Q^2 \leq 10 \,\mbox{GeV}^2$. 
2724: There also exist phenomenological fits for the longitudinally
2725: polarized PDFs, which contains the information on $\Delta \Sigma^Q$
2726: and $\Delta g$, although with larger uncertainties compared with
2727: the unpolarized case. These phenomenological PDFs can therefore be used
2728: for estimating the orbital angular momenta carries by the quark
2729: fields and the gluon fields through the relations :
2730: \begin{eqnarray}
2731:  L^Q &=& J^Q - \frac{1}{2} \,\Delta \Sigma^Q, \\
2732:  L^g &=& J^g - \Delta g .
2733: \end{eqnarray}
2734: The values of $L^Q$ and $L^g$ at $Q^2 = 4 \,\mbox{GeV}^2$ estimated
2735: in this way are shown in Table.\ref{fig:modelindep}. Here, we use
2736: MRST2004 PDF fit to estimate $J^Q$ and $J^g$ \cite{MRST2004}.
2737: On the other hand,
2738: $\Delta \Sigma^Q$ and $\Delta g$ are estimated by using three
2739: independent PDF fits, i.e. LSS2005, DNS2005, and GRSV2000
2740: \cite{LSS2005},\cite{DNS2005},\cite{GRSV2000}.
2741: Here, all of the three independent fits we are using correspond
2742: to the $\overline{\rm MS}$ regularization scheme.
2743: As one sees, there are sizable uncertainties for the phenomenological
2744: values of $\Delta \Sigma^Q$ and $\Delta g$.
2745: Still, a common conclusion obtained from all these PDF fits is a
2746: {\it very important role of the quark orbital angular momentum}.
2747: The table \ref{fig:modelindep} shows, at the least, that the magnitude
2748: of the quark orbital angular momentum is comparable with that of
2749: the intrinsic quark spin even at the scale of $Q^2 = 4 \,\mbox{GeV}^2$.
2750: Since the quark orbital angular momentum is a rapidly decreasing
2751: function of the energy scale, while the scale dependence of
2752: $\Delta \Sigma^Q$ is very weak, this means that the former is
2753: dominant over the latter at the scale below $Q^2 \simeq 1 \,\mbox{GeV}^2$
2754: where any low energy models are supposed to hold.
2755: Naturally, the whole argument here is crucially dependent on one
2756: theoretical postulate that $B_{20}^Q (0) \simeq 0$.
2757: Although it is supported by both of the LHPC and QCDSF lattice
2758: simulations, efforts to improve the accuracy of the lattice prediction
2759: should be continued, in consideration of its extremely important
2760: role in determining the quark-gluon contents of the nucleon spin.
2761: Also highly desirable is an analytical proof of it within the
2762: framework of (nonperturbative) QCD.
2763: 
2764: 
2765: \begin{table}[htbp]
2766: \begin{center}
2767: \renewcommand{\arraystretch}{1.00}
2768: \caption{The model independent predictions of the spin
2769: contents of the nucleon at $Q^2 = 4 \,\mbox{GeV}^2$, based
2770: only upon one theoretical postulate $B_{20}^Q (0) = 0$.
2771: Here, all of the three independent fits for the longitudinally
2772: polarized PDFs correspond to the $\overline{\rm MS}$ scheme.
2773: \label{fig:modelindep}}
2774: \vspace{10mm}
2775: \begin{tabular}{cc|cc|cc} \hline
2776:  \multicolumn{2}{c|}{MRST2004} & 
2777:  \multicolumn{2}{c|}{LSS2005} & 
2778:  \multicolumn{2}{c}{MRST2004 + LSS2005} \\ \hline
2779:  \ \ \ \ \ $J^Q$ \ \ \ & \ \ \ $J^g$ \ \ \ & \ \ \ $\Delta \Sigma$ \ \ \ &
2780:  \ \ \ $\Delta g$ \ \ \ & 
2781:  \ \ \ \ \ $L^Q$ \ \ \ & \ \ \ $L^g$ \ \ \ \\ \hline
2782:  0.289 & 0.211 & 0.198 & 0.368 & 0.190 & -0.157 \\ \hline
2783:  \ \\ \hline
2784:  \multicolumn{2}{c|}{MRST2004} & 
2785:  \multicolumn{2}{c|}{DNS2005} & 
2786:  \multicolumn{2}{c}{MRST2004 + DNS2005} \\ \hline
2787:  \ \ \ \ \ $J^Q$ \ \ \ & \ \ \ $J^g$ \ \ \ & \ \ \ $\Delta \Sigma$ \ \ \ &
2788:  \ \ \ $\Delta g$ \ \ \ & 
2789:  \ \ \ \ \ $L^Q$ \ \ \ & \ \ \ $L^g$ \ \ \ \\ \hline
2790:  0.289 & 0.211 & 0.313 & 0.477 & 0.133 & -0.266 \\ \hline
2791:  \ \\ \hline
2792:  \multicolumn{2}{c|}{MRST2004} & 
2793:  \multicolumn{2}{c|}{GRSV2000} & 
2794:  \multicolumn{2}{c}{MRST2004 + GRSV2000} \\ \hline
2795:  \ \ \ \ \ $J^Q$ \ \ \ & \ \ \ $J^g$ \ \ \ & \ \ \ $\Delta \Sigma$ \ \ \ &
2796:  \ \ \ $\Delta g$ \ \ \ & 
2797:  \ \ \ \ \ $L^Q$ \ \ \ & \ \ \ $L^g$ \ \ \ \\ \hline
2798:  0.289 & 0.211 & 0.137 & 0.623 & 0.221 & -0.412 \\ \hline
2799: \end{tabular}
2800: \end{center}
2801: \renewcommand{\arraystretch}{1.20}
2802: \end{table}
2803: 
2804: 
2805: 
2806: \section{Concluding remarks}
2807: 
2808:  In this paper, we have investigated the generalized form factors of the 
2809: nucleon, which will be extracted through near-future measurements of the 
2810: generalized parton distribution functions, within the framework of the 
2811: CQSM. A particular emphasis is put on the pion mass dependence as well
2812: as the scale dependence of the model predictions, which we
2813: compare with the corresponding predictions of the lattice QCD
2814: by the LHPC and the QCDSF collaborations carried out in the heavy
2815: pion regime around $m_{\pi} \simeq (700 \sim 900) \,\mbox{MeV}$.
2816: The generalized form factors contain the ordinary electromagnetic
2817: form factors of the nucleon such as the Dirac and Pauli form factors
2818: of the proton and the neutron.
2819: We have shown that the CQSM with good chiral symmetry reproduces well
2820: the general behaviors of the observed electromagnetic form factors,
2821: while the lattice simulations by the above two groups have a tendency
2822: to underestimate the electromagnetic sizes of the nucleon.
2823: Undoubtedly, this cannot be unrelated to the fact that the above two
2824: lattice simulations were performed with unrealistically heavy pion
2825: mass.
2826: 
2827: We have also tried to figure out the underlying spin contents of the
2828: nucleon through the analysis of the gravitoelectric and 
2829: gravitomagnetic form factors of the nucleon, by taking care of the pion
2830: mass despondencies as well as of the scale dependencies of the
2831: relevant quantities. After taking account of the scale dependencies by
2832: means of the QCD evolution equations at the NLO in the $\overline{MS}$
2833: scheme, the CQSM predicts, at $Q^2 = 4 \,\mbox{GeV}^2$, that
2834: $2 \,J^Q \simeq 0.68, \Delta \Sigma^Q \simeq 0.32$, and
2835: $2 \,L^Q \simeq 0.36$, which means that the quark
2836: orbital angular momentum carries sizable amount of total nucleon spin
2837: even at such a relatively high energy scale.
2838: It contradicts the conclusion of the LHPC and QCDSF
2839: collaborations indicating that the total orbital angular momentum of
2840: quarks is very small or consistent with zero. It should be recognized,
2841: however, that the prediction of the CQSM for the total quark angular
2842: momentum is not extremely far from the corresponding lattice prediction
2843: $2 \,J^Q \simeq 0.6$ at the same renormalization scale.
2844: The cause of discrepancy can therefore be traced back to the LHPC and
2845: QCDSF lattice QCD predictions for the quark spin fraction
2846: $\Delta \Sigma^Q$ around 0.6, which contradicts not only the
2847: prediction of the CQSM but also the
2848: EMC observation. As was shown in our recent paper \cite{Wakamatsu05},
2849: $\Delta \Sigma^Q$ is such a
2850: quantity that is extremely sensitive to the variation of the pion
2851: mass, especially in the region close to the chiral limit.
2852: More serious lattice QCD studies on the $m_\pi$-dependence of
2853: $\Delta \Sigma^Q$ is highly desirable.
2854: 
2855: Worthy of special mention is the fact that, once we accept a
2856: theoretical postulate $B_{20}^Q (0) = 0$, i.e, the
2857: absence of the net quark contribution to the anomalous gravitomagnetic
2858: moment of the nucleon, which is supported by both of the LHPC and
2859: QCDSF lattice simulations, we are necessarily led to a surprisingly
2860: simple relations, $J^Q = \frac{1}{2} \,\langle x \rangle^Q$ and
2861: $J^g = \frac{1}{2} \,\langle x \rangle^g$, i.e. the proportionality
2862: of the linear and angular momentum fractions carried by the
2863: quarks and the gluons. Using these relations,
2864: together with the existing empirical information for the unpolarized
2865: and the longitudinally polarized PDFs,
2866: we can give {\it model-independent predictions} for the
2867: quark and gluon contents of the nucleon spin.
2868: For instance, with combined use of the MRST2004 fit \cite{MRST2004}
2869: and the DNS2005 fit \cite{DNS2005}, we obtain $2 \,J^Q \simeq 0.58$,
2870: $\Delta \Sigma^Q \simeq 0.31$, and $2 \,L^Q \simeq 0.27$ at
2871: $Q^2 = 4 \,\mbox{GeV}^2$.
2872: Since $L^Q$ (as well as $J^Q$) is a rapidly decreasing function of
2873: the energy scale, while the scale dependence of $\Delta \Sigma^Q$
2874: is very weak, we must conclude that the former is even more
2875: dominant over the latter at the scale below $Q^2 \simeq 1 \,\mbox{GeV}^2$ where any low energy models are supposed to hold.
2876: 
2877: The situation is a little more complicated in the flavor-nonsinglet
2878: (or isovector) channel, because $B_{20}^{u-d}(0) \equiv 
2879: G_{M,20}^{(I=1)}(0) - G_{E,20}^{(I=1)}(0) = 2 \,J^{u-d} - 
2880: \langle x \rangle^{u-d} \neq 0$, and also because the CQSM and the
2881: lattice QCD give fairly different predictions for $G_{M,20}^{(I=1)}(0)$.
2882: As compared with the lattice prediction for $G_{M,20}^{(I=1)}(0)$
2883: around $0.8$, the predictions of the CQSM turns out to be around $0.4$.
2884: We have argued that the relatively small value of
2885: $G_{M, 20}^{(I=1)} (0)$ obtained in the CQSM is intimately connected
2886: with the small $x$ enhancement of the generalized parton distribution
2887: $E_M^{(I=1)} (x, 0, 0)$, which is dominated by the clouds of pionic
2888: $q \bar{q}$ excitation around $x \simeq 0$.
2889: (We recall that the 2nd moment of $E_M^{(I=1)} (x, 0, 0)$ gives 
2890: $G_{M, 20}^{(I=1)} (0)$.) Unfortunately, such a $x$-dependent
2891: distribution as $E_M^{(I=1)} (x, 0, 0)$ cannot be accessed within
2892: the framework of lattice QCD. Still, the predicted small $x$ behavior
2893: of $E_M^{u-d} (x,0,0) \equiv E_M^{(I=1)} (x,0,0)$ as well as of
2894: $f^{u-d}(x)$ indicates again
2895: the importance of chiral dynamics in the nucleon structure function
2896: physics, which has not been fully accounted for in the lattice QCD
2897: simulation at the present level.
2898: 
2899: \vspace{3mm}
2900: % If you have acknowledgments, this puts in the proper section head.
2901: \begin{acknowledgments}
2902: This work is supported in part by a Grant-in-Aid for Scientific
2903: Research for Ministry of Education, Culture, Sports, Science
2904: and Technology, Japan (No.~C-16540253)
2905: \end{acknowledgments}
2906: 
2907: 
2908: \vspace{10mm}
2909: \appendix
2910: 
2911: \section{proof of the momentum sum rule}
2912: 
2913: Here we closely follow the proof of the momentum sum rule given in
2914: \cite{DPPPW96}, by taking into account a necessary modification
2915: in the case of
2916: $m_{\pi} \neq 0$. The starting point is the following expression for the 
2917: soliton mass (or the static soliton energy) :
2918: \begin{equation}
2919:  M_N \ = \ N_c \,\mbox{Sp} [\, \theta(E_0 - H + i \varepsilon) \,H \,]
2920:  \ - \ (H \rightarrow H_0) \ + \ E_M ,
2921: \end{equation}
2922: with
2923: \begin{equation}
2924:  E_M [F(r)] \ = \ - \,f_\pi^2 \,m_\pi^2 \,\int \,[
2925:  \cos F(r) - 1] \,d^3 x . \label{eq:Emeson}
2926: \end{equation}
2927: The soliton mass must be stationary with respect to an arbitrary
2928: variation of the chiral field $U$ or equivalently the soliton profile
2929: $F (r)$, which lead to a saddle point equation : 
2930: \begin{equation}
2931:  \mbox{Sp} \,[ \theta ( E_0 - H + i \varepsilon ) \,\delta H ] 
2932:  \ + \ \delta E_m \ = \ 0 .
2933: \end{equation}
2934: Here we consider a particular (dilatational) variation of chiral field 
2935: \begin{equation}
2936:  U (x) \longrightarrow U ( ( 1 + \xi) x ) .
2937: \end{equation}
2938: For infinitesimal $\xi$, we have 
2939: \begin{equation}
2940:  \delta U \equiv U ( ( 1 + \xi) x ) - U (x) \ \simeq \ 
2941:  \xi \,x^k \,\partial_k U (x) ,
2942: \end{equation}
2943: so that 
2944: \begin{eqnarray}
2945:  \delta H &=& M \,\gamma^0 \,\xi \,x^k \,\partial_k U^{\gamma_5} \ = \ 
2946:  \xi \,[ x^k \partial_k, M \gamma^0 U^{\gamma_5} ]
2947:  \ = \ \xi \,( [ x^k \partial_k, H ] - i \gamma^0 \gamma^k \partial_k ) .
2948: \end{eqnarray}
2949: Noting the identity
2950: \begin{eqnarray}
2951:  &\,& \mbox{Sp} (\theta ( E_0 - H + i \varepsilon) 
2952:  [ x^k \partial_k, H ] ) \ = \ 
2953:  \mbox{Sp} ( [ H, \theta (E_0 - H + i \varepsilon) ] x^k \partial_k)
2954:  \ = \ 0 ,
2955: \end{eqnarray}
2956: we therefore obtain a key identity
2957: \begin{equation}
2958:  \xi \,\mbox{Sp} [\theta (E_0 - H + i \varepsilon) (-i) 
2959:  \gamma^0 \gamma^k \partial_k ] = - \,\delta E_M . \label{eq:keyID}
2960: \end{equation}
2961: Now, by using (\ref{eq:Emeson}) together with the relations, 
2962: \begin{eqnarray}
2963:  \delta F (r) &=& \xi \,r \,F^{\prime} (r) , \\
2964:  \delta \cos F (r) &=& - \,\sin F (r) \,\delta F
2965:  = - \,\xi \,r \,\sin F (r) \,F^{\prime} (r) ,
2966: \end{eqnarray}  
2967: we get
2968: \begin{eqnarray}
2969:  \delta E_M &=& \xi \cdot 4 \pi \,f_{\pi}^2 \,m_{\pi}^2 \,
2970:  \int_0^{\infty} \,d r \,r^3 \,\sin F (r) \,F^{\prime} (r) \\
2971:  &=& - \,\xi \cdot 4 \pi \,f_{\pi}^2 \,m_{\pi}^2 \,
2972:  \int_0^{\infty} \,d r \,r^3 \,\frac{d}{d r} \cos F (r) .
2973: \end{eqnarray}
2974: Here, taking account of the boundary condition
2975: \begin{equation}
2976:  F (0) = \pi, \ \ \ F (\infty) = 0 ,
2977: \end{equation}
2978: we can manipulate as 
2979: \begin{eqnarray}
2980:  \int_0^{\infty} \,d r \,r^3 \,\frac{d}{d r} \cos F (r) 
2981:  &=& \int_0^{\infty} \,d r \,r^3 \,\frac{d}{d r} \,(\cos F (r) - 1) \\
2982:  &=& r^3 \,(\cos F (r) - 1) |_0^{\infty}
2983:  \ - \ 3 \,\int_0^{\infty} \,d r \,r^2 \,(\cos F (r) - 1) \\
2984:  &=& - \,3 \,\int_0^{\infty} \,d r \,r^2 \,(\cos F (r) - 1) .
2985: \end{eqnarray}
2986: We thus find an important relation : 
2987: \begin{equation}
2988:  \delta E_M \ = \ - \,3 \,\xi \,E_M .
2989: \end{equation}
2990: Putting this relation into (\ref{eq:keyID}), we have
2991: \begin{equation}
2992:  \xi \,\mbox{Sp} (\theta (E_0 - H + i \varepsilon) 
2993:  \vecvar{\alpha} \cdot \vecvar{p} )
2994:  = 3 \,\xi \,E_M ,
2995: \end{equation}
2996: or 
2997: \begin{equation}
2998:  \frac{1}{3} \,\mbox{Sp} (\theta (E_0 - H + i \varepsilon) 
2999:  \vecvar{\alpha} \cdot \vecvar{p}) = E_M . \label{eq:traceAP}
3000: \end{equation}
3001: If we evaluate the trace sum above by using the eigenstates of the
3002: static Dirac Hamiltonian $H$ as a complete set of basis,
3003: (\ref{eq:traceAP}) can also be written as
3004: \begin{equation}
3005:  \sum_{n \leq 0} \,\langle n | \,\frac{1}{3} \,
3006:  \vecvar{\alpha} \cdot \vecvar{p} \,| n \rangle = E_M ,
3007: \end{equation}
3008: which is the relation quoted in (\ref{eq:alfp}).
3009: We point out that our result has a 
3010: correct chiral limit, since 
3011: $E_M \rightarrow 0$ as $m_{\pi} \rightarrow 0$
3012: and therefore
3013: \begin{equation}
3014:  \lim_{m_\pi \rightarrow 0} \,\sum_{n \leq 0} \,
3015:  \langle n | \,\frac{1}{3} \,
3016:  \vecvar{\alpha} \cdot \vecvar{p} \,| n \rangle \ = \ 0 ,
3017: \end{equation}
3018: in conformity with the proof given in ref.\cite{DPPPW96}.
3019: 
3020: % Create the reference section using BibTeX:
3021: %\bibliographystyle{unsrt}
3022: %\bibliography{GFspin}
3023: 
3024: \begin{thebibliography}{10}
3025: 
3026: \bibitem{EMC88}
3027: EMC Collaboration~: J.~Aschman~et al.
3028: \newblock {\em Phys. Lett.}, B206:364, 1988.
3029: 
3030: \bibitem{EMC89}
3031: EMC Collaboration~: J.~Aschman~et al.
3032: \newblock {\em Nucl. Phys.}, B328:1, 1989.
3033: 
3034: \bibitem{MRGDH94}
3035: D.~Mueller, D.~Robaschik, B.~Geyer, F.M.~Dittes, and J.~Horejsi,
3036: \newblock {\em Fortsch. Phys.}, 42:101, 1994.
3037: 
3038: \bibitem{DMRGH88}
3039: F.M.~Dittes, D.~Mueller, D.~Robaschik, B.~Geyer, and J.~Horejsi,
3040: \newblock {\em Phys. Lett.}, B209:325, 1988.
3041: 
3042: \bibitem{Ji98}
3043: X.~Ji.
3044: \newblock {\em J. Phys.}, G24:1181, 1998.
3045: 
3046: \bibitem{GPV01}
3047: K.~Goeke, M.V.~Polyakov, and M.~Vanderhaeghen
3048: \newblock {\em Prog. Part. Nucl. Phys.}, 47:401, 2001.
3049: 
3050: \bibitem{Diehl03}
3051: M.~Diehl.
3052: \newblock {\em Phys. Rep.}, 388:41, 2003.
3053: 
3054: \bibitem{BR05}
3055: A.V.~Belitsky and A.V.~Radyushkin.
3056: \newblock {\em Phys. Rep.}, 418:1, 2005.
3057: 
3058: \bibitem{Ji97}
3059: X.~Ji.
3060: \newblock {\em Phys. Rev. Lett.}, 78:610, 1997.
3061: 
3062: \bibitem{HJL99}
3063: P.~Hoodbhoy, X.~Ji, and W.~Lu.
3064: \newblock {\em Phys. Rev.}, D59:014013, 1999.
3065: 
3066: \bibitem{JTH96}
3067: X.~Ji, J.~Tang, and P.~Hoodbhoy.
3068: \newblock {\em Phys. Rev. Lett.}, 76:740, 1996.
3069: 
3070: \bibitem{Burkardt00}
3071: M.~Burkardt.
3072: \newblock {\em Phys. Rev.}, D62:071503, 2000.
3073: 
3074: \bibitem{Burkardt03}
3075: M.~Burkardt.
3076: \newblock {\em Int. J. Mod. Phys.}, A18:173, 2003.
3077: 
3078: \bibitem{RP02}
3079: J.P.~Ralston and B.~Pire.
3080: \newblock {\em Phys. Rev.}, D66:111501, 2002.
3081: 
3082: \bibitem{DFJK05}
3083: M.~Diehl, Th.~Feldmann, R.~Jacob and P.~Kroll.
3084: \newblock {\em Eur. Phys. J.}, C39:1, 2005.
3085: 
3086: \bibitem{Diehl05}
3087: M.~Diehl.
3088: \newblock {\em hep-ph/0510221}.
3089: 
3090: \bibitem{LHPC02}
3091: D.~Dolgov et al.
3092: \newblock {\em Phys. Rev.}, D66:034506, 2002.
3093: 
3094: \bibitem{LHPC04}
3095: Ph.~H\"{a}gler, J.W~Negele, D.B.~Renner, W.~Schroers, Th.~Lippert
3096: and K.~Schilling.
3097: \newblock {\em Phys. Rev. Lett.}, 93:112001, 2004.
3098: 
3099: \bibitem{LHPC05}
3100: Ph.~H\"{a}gler, J.W~Negele, D.B.~Renner, W.~Schroers, Th.~Lippert
3101: and K.~Schilling.
3102: \newblock {\em Eur. Phys. J.}, A24:29, 2005.
3103: 
3104: \bibitem{QCDSF05}
3105: QCDSF Collaboration~: M.~G\"{o}ckeler, Ph.~H\"{a}gler, R.~Horsley,
3106: D. Pleiter, P.E.L.~Rakow, A.~Sch\"{a}fer, G.~Schierholz, and 
3107: J.M.~Zanotti.
3108: \newblock {\em Few-Body Systems}, 36:111, 2005.
3109: 
3110: \bibitem{QCDSF04a}
3111: M.~G\"{o}ckeler, R.~Horsley, D. Pleiter, P.E.L.~Rakow, A.~Sch\"{a}fer,
3112: G.~Schierholz, and W.~Schroers.
3113: \newblock {\em Phys. Rev. Lett.}, 92:042002, 2004.
3114: 
3115: \bibitem{QCDSF04b}
3116: M.~G\"{o}ckeler, T.R.~Hemmert, R.~Horsley, D. Pleiter, P.E.L.~Rakow,
3117: A.~Sch\"{a}fer, G.~Schierholz, and W.~Schroers.
3118: \newblock {\em Nucl. Phys. (Proc. Suppl.)}, 128:203, 2004.
3119: 
3120: \bibitem{DPP88}
3121: D.I.~Diakonov, V.Yu.~Petrov, and P.V.~Pobylitsa.
3122: \newblock {\em Nucl. Phys.}, B306:809, 1988.
3123: 
3124: \bibitem{WY91}
3125: M.~Wakamatsu and H.~Yoshiki.
3126: \newblock {\em Nucl. Phys.}, A524:561, 1991.
3127: 
3128: \bibitem{KR84}
3129: S.~Kahana and G.~Ripka.
3130: \newblock {\em Nucl. Phys.}, A429:462, 1984.
3131: 
3132: \bibitem{KRS84}
3133: S.~Kahana, G.~Ripka, and V.~Soni.
3134: \newblock {\em Nucl. Phys.}, A415:351, 1984.
3135: 
3136: \bibitem{KWW99}
3137: T.~Kubota, M.~Wakamatsu, and T.~Watabe.
3138: \newblock {\em Phys. Rev.}, D60:014016, 1999.
3139: 
3140: \bibitem{Ossmann05}
3141: J. Ossmann, M.V.~Polyakov, P.~Schweitzer, D.~Urbano, and K.~Goeke.
3142: \newblock {\em Phys. Rev.}, D71:034011, 2005.
3143: 
3144: \bibitem{WT05}
3145: M.~Wakamatsu and H.~Tsujimoto.
3146: \newblock {\em Phys. Rev.}, D71:074001, 2005.
3147: 
3148: \bibitem{Petrov98}
3149: V.Yu. Petrov, P.V. Pobylitsa, M.V. Polyakov, I.~B\"{o}ring, K.~Goeke
3150: and C.~Weiss.
3151: \newblock {\em Phys. Rev.}, D57:4325, 1998.
3152: 
3153: \bibitem{DPPPW96}
3154: D.I. Diakonov, V.Yu. Petrov, P.V. Pobylitsa, M.V. Polyakov,
3155: and C.~Weiss.
3156: \newblock {\em Nucl. Phys.}, B480:341, 1996.
3157: 
3158: \bibitem{BHMS01}
3159: S.J.~Brodsky, D.S.~Hwang, B.-Q.~Ma, and I. Schmidt.
3160: \newblock {\em Nucl. Phys.}, B593:311, 2001.
3161: 
3162: \bibitem{Wakamatsu92R}
3163: M.~Wakamatsu.
3164: \newblock {\em Prog. Theor. Phys. Suppl.}, 109:115, 1992.
3165: 
3166: \bibitem{Christov96}
3167: Chr.V.~Christov, A.~Blotz, H.-C.~Kim, P.~Pobylitsa, T.~Watabe
3168: Th.Meissner, E.Ruiz-Arriola, and K.~Goeke.
3169: \newblock {\em Prog. Part. Nucl. Phys.}, 37:91, 1996.
3170: 
3171: \bibitem{ARW96}
3172: R.~Alkofer, H.~Reinhardt, and H.~Weigel.
3173: \newblock {\em Phys. Rep.}, 265:139, 1996.
3174: 
3175: \bibitem{WGR96}
3176: H.~Weigel, L.~Gamberg, and H.~Reinhardt.
3177: \newblock {\em Mod. Phys. Lett.}, A11:3021, 1996.
3178: 
3179: \bibitem{GRW98}
3180: L.~Gamberg, H.~Reinhardt, and H.~Weigel.
3181: \newblock {\em Phys. Rev.}, D58:054014, 1998.
3182: 
3183: \bibitem{DPPPW97}
3184: D.I. Diakonov, V.Yu. Petrov, P.V. Pobylitsa, M.V. Polyakov, and C.~Weiss.
3185: \newblock {\em Phys. Rev.}, D56:4069, 1997.
3186: 
3187: \bibitem{WK98}
3188: M.~Wakamatsu and T.~Kubota.
3189: \newblock {\em Phys. Rev.}, D57:5755, 1998.
3190: 
3191: \bibitem{WK99}
3192: M.~Wakamatsu and T.~Kubota.
3193: \newblock {\em Phys. Rev.}, D60:034020, 1999.
3194: 
3195: \bibitem{Wakamatsu03a}
3196: M.~Wakamatsu.
3197: \newblock {\em Phys. Rev.}, D67:034005, 2003.
3198: 
3199: \bibitem{Wakamatsu03b}
3200: M.~Wakamatsu.
3201: \newblock {\em Phys. Rev.}, D67:034006, 2003.
3202: 
3203: \bibitem{PPGWW99}
3204: P.V. Pobylitsa, M.V. Polyakov, K.~Goeke, T.~Watabe, and C.~Weiss.
3205: \newblock {\em Phys. Rev.}, D59:034024, 1999.
3206: 
3207: \bibitem{Wakamatsu92}
3208: M.~Wakamatsu.
3209: \newblock {\em Phys. Rev.}, D46:3762, 1992.
3210: 
3211: \bibitem{Christov95}
3212: Chr.V.~Christov, A.Z.~G\'{o}rski, K.~Goeke, and P.V.~Pobylitsa.
3213: \newblock {\em Phys. Rev.}, A592:513, 1995.
3214: 
3215: \bibitem{Wakamatsu05}
3216: M.~Wakamatsu.
3217: \newblock {\em Phys. Rev.}, D72:074006, 2005.
3218: 
3219: \bibitem{Teryaev99}
3220: O.V.~Teryaev.
3221: \newblock {\em hep-ph/9904376}, 1999.
3222: 
3223: \bibitem{Teryaev98}
3224: O.V.~Teryaev.
3225: \newblock {\em hep-ph/9803403}, 1998.
3226: 
3227: \bibitem{Teryaev03}
3228: O.V.~Teryaev.
3229: \newblock {\em Czech. J. Phys.}, 53:47, 2003.
3230: 
3231: \bibitem{WW93}
3232: M.~Wakamatsu and T.~Watabe.
3233: \newblock {\em Phys. Lett.}, B312:184, 1993.
3234: 
3235: \bibitem{CBGPPWW94}
3236: Chr.V.~Christov, A.~Blotz, K.~Goeke, P.~Pobylitsa, V.Yu.~Petrov,
3237: M.~Wakamatsu, and T.~Watabe.
3238: \newblock {\em Phys. Lett.}, B325:467, 1994.
3239: 
3240: \bibitem{Wakamatsu96}
3241: M.~Wakamatsu.
3242: \newblock {\em Prog. Theor. Phys.}, 95:143, 1996.
3243: 
3244: \bibitem{NMC91}
3245: NMC Collaboration~: P.~Amaudruz et al.
3246: \newblock {\em Phys. Rev. Lett.}, 66:2712, 1991.
3247: 
3248: \bibitem{Saga96}
3249: M.~Miyama and S.~Kumano.
3250: \newblock {\em Compt. Phys. Commun.}, 94:185, 1996.
3251: 
3252: \bibitem{HM91}
3253: E.M.~Henley and G.A.~Miller.
3254: \newblock {\em Phys. Lett.}, B251:453, 1991.
3255: 
3256: \bibitem{KL91}
3257: S.~Kumano and J.T.~Londergan.
3258: \newblock {\em Phys. Rev.}, D43:59, 1991.. 
3259: AIP Conf. Proc. 223, 1991.
3260: 
3261: \bibitem{HERMES98}
3262: HEREMES Collaboration~: K.~Ackerstaff.
3263: \newblock {\em Phys. Rev. Lett.}, 81:5519, 1998.
3264: 
3265: \bibitem{E866}
3266: E866 Collaboration~: E.A.~Hawker et al.
3267: \newblock {\em Phys. Rev. Lett.}, 80:3715, 1998.
3268: 
3269: \bibitem{Arnold86}
3270: R.~Arnold et al.
3271: \newblock {\em Phys. Rev. Lett.}, 57:174, 1986.
3272: 
3273: \bibitem{FRS77}
3274: E.G.~Floratos, D.A..~Ross, and C.T.~Sachrajda.
3275: \newblock {\em Nucl. Phys.}, B129:66, 1977,
3276: and Erratum, B139:545, 1978.
3277: 
3278: \bibitem{LY81}
3279: C.~Lop\'{e}z and F.J.~Yndur\'{a}in.
3280: \newblock {\em Nucl. Phys.}, B183:157, 1981.
3281: 
3282: \bibitem{ABY97}
3283: K.~Adel, F.~Barreiro, and F.J.~Yndur\'{a}in.
3284: \newblock {\em Nucl. Phys.}, B495:221, 1997.
3285: 
3286: \bibitem{MRST98}
3287: A.D.~Martin, R.G.~Roberts, W.J.~Stirling, and R.S.~Thorne.
3288: \newblock {\em Eur. Phys. J.}, C4:463, 1998.
3289: 
3290: \bibitem{MRST2004}
3291: A.D.~Martin, W.J.~Stirling, and R.S.~Thorne.
3292: \newblock {\em hep-ph/0603143}, 2006.
3293: 
3294: \bibitem{CTEQ5}
3295: L.H.~Lai, J.~Huston, S.~Kuhlmann, J.~Morfin, F.~Olness,
3296: J.F.~Owen, J.~Pumplin, and W.K.~Tung.
3297: \newblock {\em Eur. Phys. J.}, C12:375, 2000.
3298: 
3299: \bibitem{GRSV2000}
3300: M.~Gl\"{u}ck. E.~Reya, and M.~Strattmann, and W.~Vogelsang.
3301: \newblock {\em Phys. Rev.}, D63:094005, 2001.
3302: 
3303: \bibitem{LSS2005}
3304: E.~Leader. A.V.~Sidorov, and D.B.~Stamenov.
3305: \newblock {\em Phys. Rev.}, D73:034023, 2006.
3306: 
3307: \bibitem{DNS2005}
3308: D.de~Florian, G.A.~Navarro, and R.~Sassot.
3309: \newblock {\em Phys. Rev.}, D71:094018, 2005.
3310: 
3311: \bibitem{GRSV96}
3312: M.~Gl\"{u}ck. E.~Reya, M.~Strattmann, and W.~Vogelsang.
3313: \newblock {\em Phys. Rev.}, D53:4775, 1996.
3314: 
3315: \bibitem{LSS2000}
3316: E.~Leader. A.V.~Sidorov, and D.B.~Stamenov.
3317: \newblock {\em Phys. Lett.}, B488:283, 2000.
3318: 
3319: \bibitem{Cheng98}
3320: H.-Y.~Cheng.
3321: \newblock {\em Phys. Lett.}, B427:371, 1998.
3322: 
3323: \bibitem{BFR96}
3324: R.D.~Ball. S.~Forte, and G.~Ridolfi.
3325: \newblock {\em Phys. Lett.}, B378:255, 1996.
3326: 
3327: \bibitem{ABFS97}
3328: G.~Altarelli, R.D.~Ball. S.~Forte, and G.~Ridolfi.
3329: \newblock {\em Nucl. Phys.}, B496:337, 1997.
3330: 
3331: \bibitem{MDLMM00}
3332: N.~Mathur, S.J.~Dong. K.F.~Liu, L.~Mankiewicz, and N.C.~Mukhopadhyay.
3333: \newblock {\em Phys. Rev.}, D62:114504, 2000.
3334: 
3335: \bibitem{FKOU95}
3336: M.~Fukugita. Y.~Kuramashi, M.~Okawa, and A.~Ukawa.
3337: \newblock {\em Phys. Phys. Lett.}, 75:2092, 1995.
3338: 
3339: \bibitem{DLL95}
3340: S.J.~Dong. J.-F.Laga\"{e}, and K.F.~Liu.
3341: \newblock {\em Phys. Phys. Lett.}, 75:2096, 1995.
3342: 
3343: \end{thebibliography}
3344: 
3345: \end{document}
3346: %
3347: % ****** End of file template.aps ******
3348: