hep-ph0701076/IC.tex
1: \documentclass[a4paper,showpacs,showkeys,nofootinbib,aps]{revtex4}
2: %\usepackage{a4}
3: %\usepackage{amsmath}
4: \usepackage{epsfig}
5: %%%%%%%%%%%%%%%%%% Page format %%%%%%%%%%%%%%%%%%%
6: %\textwidth 173mm
7: %\textheight 215mm
8: %\topmargin -50pt
9: \topmargin 1cm %
10: \voffset -50pt %
11: %\oddsidemargin -0.5cm %
12: %\evensidemargin -0.5cm
13: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
14: %\bibliographystyle{apsrev}
15: %\reversemarginpar
16: %\usepackage{axodraw}
17: %\usepackage[dvipdfm,pdfstartview=FitH]{hyperref}
18: %\usepackage[bookmarksnumbered,colorlinks,plainpages,backref]{hyperref}
19: %\hypersetup{pdfstartview=FitH, bookmarks=true}
20: \begin{document}
21: \title{Azimuthal Asymmetries in DIS as a Probe \\
22: of Intrinsic Charm Content of the Proton}
23: \author{L.N.~Ananikyan}
24:  \email{lev@web.am}
25:  \affiliation{Yerevan Physics Institute, Alikhanian Br.2, 375036 Yerevan, Armenia}
26: \author{N.Ya.~Ivanov}
27:  \email{nikiv@uniphi.yerphi.am}
28: \affiliation{Yerevan Physics Institute, Alikhanian Br.2, 375036 Yerevan, Armenia}
29: \date{\today}
30: \begin{abstract}
31: \noindent We calculate the azimuthal dependence of the heavy-quark-initiated ${\cal O}(\alpha_{s})$
32: contributions to the lepton-nucleon deep inelastic scattering (DIS). It is shown that, contrary to
33: the photon-gluon fusion (GF) component, the photon-quark scattering (QS) mechanism is practically
34: $\cos2\varphi$-independent. We investigate the possibility to discriminate experimentally between
35: the GF and QS contributions using their strongly different azimuthal distributions. Our analysis
36: shows that the GF and QS predictions for the azimuthal $\cos2\varphi$ asymmetry are quantitatively
37: well defined in the fixed flavor number scheme: they are stable, both parametrically and
38: perturbatively. We conclude that measurements of the azimuthal distributions at large Bjorken $x$
39: could directly probe the intrinsic charm content of the proton. As to the variable flavor number
40: schemes, the charm densities of the recent CTEQ and MRST sets of parton distributions have a
41: dramatic impact on the $\cos2\varphi$ asymmetry in the whole region of $x$ and, for this reason,
42: can easily be measured.
43: \end{abstract}
44: \pacs{12.38.-t, 13.60.-r, 13.88.+e}%
45: \keywords{Perturbative QCD, Heavy Flavor Leptoproduction, Intrinsic Charm, Azimuthal Asymmetries}
46: \maketitle
47: \section{Introduction}
48: The notion of the intrinsic charm (IC) content of the proton has been introduced over 25 years ago
49: in Refs~\cite{BHPS,BPS}. It was shown that, in the light-cone Fock space picture
50: \cite{brod1,brod2}, it is natural to expect a five-quark state contribution to the proton wave
51: function. The probability to find in a nucleon the five-quark component $\left\vert
52: uudc\bar{c}\right\rangle$ is of higher twist since it scales as $1/m^{2}$ where $m$ is the
53: $c$-quark mass \cite{polyakov}. This component can be generated by $gg\rightarrow c\bar{c}$
54: fluctuations inside the proton where the gluons are coupled to different valence quarks. Since all
55: of the quarks tend to travel coherently at same rapidity in the $\left\vert
56: uudc\bar{c}\right\rangle $ bound state, the heaviest constituents carry the largest momentum
57: fraction. For this reason, one would expect that the intrinsic charm component to be dominate the
58: $c$ -quark production cross sections at sufficiently large Bjorken $x$. So, the original concept of
59: the charm density in the proton \cite{BHPS,BPS} has nonperturbative nature and will be referred to
60: in the present paper as nonperturbative IC.
61: 
62: A decade ago another point of view on the charm content of the proton has been proposed in the
63: framework of the variable flavor number scheme (VFNS) \cite{ACOT,collins}. The VFNS is an approach
64: alternative to the traditional fixed flavor number scheme (FFNS) where only light degrees of
65: freedom ($u,d,s$ and $g$) are considered as active. It is well known that a heavy quark production
66: cross section contains potentially large logarithms of the type $\alpha_{s}\ln\left(
67: Q^{2}/m^{2}\right)$ whose contribution dominates at high energies, $Q^{2}\rightarrow\infty$. Within
68: the VFNS, these mass logarithms are resummed through the all orders into a heavy quark density
69: which evolves with $Q^{2}$ according to the standard DGLAP \cite{grib-lip,dokshitzer,alt-par}
70: evolution equation. Hence the VFN schemes introduce the parton distribution functions (PDFs) for
71: the heavy quarks and change the number of active flavors by one unit when a heavy quark threshold
72: is crossed. We can say that the charm density arises within the VFNS perturbatively via the
73: $g\rightarrow c\bar{c}$ evolution and will call it the perturbative IC.
74: 
75: Presently, both perturbative and nonperturbative IC are widely used for a phenomenological
76: description of available data. (A recent review of the theory and experimental constraints on the
77: charm quark distribution can be found in Refs.~\cite{pumplin,brod-higgs}. See also
78: Appendix~\ref{exp} in the present paper). In particular, practically all the recent versions of the
79: CTEQ \cite{CTEQ6} and MRST \cite{MRST2004} sets of PDFs are based on the VFN schemes and contain a
80: charm density. At the same time, the key question remains open: How to measure the intrinsic charm
81: content of the proton? The basic theoretical problem is that radiative corrections to the fixed
82: order predictions for the production cross sections are large. In particular, the next-to-leading
83: order (NLO) corrections increase the leading order (LO) results for both charm and bottom
84: production cross sections by approximately a factor of two at energies of the fixed target
85: experiments. Moreover, soft gluon resummation of the threshold Sudakov logarithms indicates that
86: higher-order contributions are also essential. (For a review see Refs.~\cite{kid2,kid1}). On the
87: other hand, perturbative instability leads to a high sensitivity of the theoretical calculations to
88: standard uncertainties in the input QCD parameters. For this reason, it is difficult to compare
89: pQCD results for spin-averaged cross sections with experimental data directly, without additional
90: assumptions. The total uncertainties associated with the unknown values of the heavy quark mass,
91: $m$, the factorization and renormalization scales, $\mu _{F}$ and $\mu _{R}$, $ \Lambda _{QCD}$ and
92: the PDFs are so large that one can only estimate the order of magnitude of the pQCD predictions for
93: production cross sections \cite{Mangano-N-R,Frixione-M-N-R}.
94: 
95: Since production cross sections are not perturbatively stable, it is of special interest to study
96: those observables that are well-defined in pQCD. A nontrivial example of such an observable was
97: proposed in Refs.~\cite{we1,we2,we4,we3} where the azimuthal $\cos2\varphi$ asymmetry in heavy
98: quark photo- and leptoproduction has been analyzed~\footnote{The well-known examples are the shapes
99: of differential cross sections of heavy flavor production which are sufficiently stable under
100: radiative corrections.}. In particular, the Born level results have been considered \cite{we1} and
101: the NLO soft-gluon corrections to the basic mechanism, photon-gluon fusion (GF), have been
102: calculated \cite{we2,we4}. It was shown that, contrary to the production cross sections, the
103: azimuthal asymmetry in heavy flavor photo- and leptoproduction is quantitatively well defined in
104: pQCD: the contribution of the dominant GF mechanism to the asymmetry is stable, both parametrically
105: and perturbatively. Therefore, measurements of this asymmetry would provide an ideal test of pQCD.
106: As was shown in Ref.~\cite{we3}, the azimuthal asymmetry in open charm photoproduction could have
107: been measured with an accuracy of about ten percent in the approved E160/E161 experiments at SLAC
108: \cite{E161} using the inclusive spectra of secondary (decay) leptons.
109: 
110: In the present paper we study the IC contribution to the azimuthal asymmetries in heavy quark
111: leptoproduction:
112: \begin{equation}
113: l(\ell )+N(p)\rightarrow l(\ell -q)+Q(p_{Q})+X[\overline{Q}](p_{X}). \label{1}
114: \end{equation}
115: Neglecting the contribution of $Z-$boson as well as the target mass effects, the cross section of
116: the reaction (\ref{1}) for unpolarized initial states  may be written as
117: \begin{equation}\label{2}
118: \frac{\text{d}^{3}\sigma _{lN}}{\text{d}x\text{d}Q^{2}\text{d}\varphi }=\frac{\alpha _{em}}{(2\pi
119: )^{2}}\frac{1}{xQ^{2}}\frac{y^2}{1-\varepsilon}\left[\sigma _{T}( x,Q^{2})+ \varepsilon\sigma_{L}(
120: x,Q^{2})+ \varepsilon\sigma_{A}( x,Q^{2})\cos 2\varphi+
121: 2\sqrt{\varepsilon(1+\varepsilon)}\sigma_{I}( x,Q^{2})\cos \varphi\right],
122: \end{equation}
123: The quantity $\varepsilon$ measures the degree of the longitudinal polarization of the virtual
124: photon in the Breit frame \cite{dombey},
125: \begin{equation}\label{3}
126: \varepsilon=\frac{2(1-y)}{1+(1-y)^2},
127: \end{equation}
128: and the kinematic variables are defined by
129: \begin{eqnarray}
130: \bar{S}=\left( \ell +p\right) ^{2},\qquad &Q^{2}=-q^{2},\qquad &x=\frac{Q^{2}}{%
131: 2p\cdot q},  \nonumber \\
132: y=\frac{p\cdot q}{p\cdot \ell },\qquad &Q^{2}=xy\bar{S},\qquad &\rho =\frac{4m^{2}%
133: }{\bar{S}}.  \label{4}
134: \end{eqnarray}
135: The cross sections $\sigma _{i}$ $(i=T,L,A,I)$ in Eq.~(\ref{2}) are related to the structure
136: functions $F_{i}(x,Q^{2})$ as follows:
137: \begin{eqnarray}
138: F_{i}(x,Q^{2}) &=&\frac{Q^{2}}{8\pi^{2}\alpha _{em}x}\,\sigma_{i}(x,Q^{2}), \qquad (i=T,L,A,I)\nonumber \\
139: F_{2}(x,Q^{2}) &=&\frac{Q^{2}}{4\pi^{2}\alpha _{em}}\,\sigma_{2}(x,Q^{2}),\label{6}
140: \end{eqnarray}
141: where $F_{2}=2x(F_{T}+F_{L})$ and $\sigma_{2}=\sigma_{T}+\sigma_{L}$.
142: \begin{figure}
143: \begin{center}
144: \mbox{\epsfig{file=graph.eps,width=200pt}}
145: \caption{\label{Fg.1}\small Definition of the azimuthal
146: angle $\varphi$ in the nucleon rest frame.}
147: \end{center}
148: \end{figure}
149: In Eq.~(\ref{2}), $\sigma _{T}\,(\sigma _{L})$ is the usual $\gamma ^{*}N$ cross section describing
150: heavy quark production by a transverse (longitudinal) virtual photon. The third cross section,
151: $\sigma _{A}$, comes about from interference between transverse states and is responsible for the
152: $\cos2\varphi$ asymmetry which occurs in real photoproduction using linearly polarized photons
153: \cite{we1,we2,we3}. The fourth cross section, $\sigma _{I}$, originates from interference between
154: longitudinal and transverse components \cite{dombey}. In the nucleon rest frame, the azimuth
155: $\varphi $ is the angle between the lepton scattering plane and the heavy quark production plane,
156: defined by the exchanged photon and the detected quark $Q$ (see Fig.~\ref{Fg.1}). The covariant
157: definition of $\varphi $ is
158: \begin{eqnarray}
159: \cos \varphi &=&\frac{r\cdot n}{\sqrt{-r^{2}}\sqrt{-n^{2}}},\qquad \qquad
160: \sin \varphi =\frac{Q^{2}\sqrt{1/x^{2}+4m_{N}^{2}/Q^{2}}}{2\sqrt{-r^{2}}%
161: \sqrt{-n^{2}}}~n\cdot \ell ,  \label{7} \\
162: r^{\mu } &=&\varepsilon ^{\mu \nu \alpha \beta }p_{\nu }q_{\alpha }\ell _{\beta },\qquad \qquad
163: \quad n^{\mu }=\varepsilon ^{\mu \nu \alpha \beta }q_{\nu }p_{\alpha }p_{Q\beta }.  \label{8}
164: \end{eqnarray}
165: In Eqs.~(\ref{4}) and (\ref{7}), $m$ and $m_{N}$ are the masses of the heavy quark and the target,
166: respectively. Usually, the azimuthal asymmetry associated with the $\cos 2\varphi $ distribution,
167: $A_{2\varphi}(\rho ,x,Q^{2})$, is defined by
168: \begin{eqnarray}
169: A_{2\varphi}(\rho ,x,Q^{2})&=&2\langle \cos 2\varphi \rangle(\rho ,x,Q^{2})=\frac{\text{d}^{3}
170: \sigma _{lN}(\varphi =0)+\text{d}^{3}\sigma _{lN}(\varphi =\pi )-2\text{d}^{3}\sigma _{lN}(\varphi
171: =\pi /2)}{ \text{d}^{3}\sigma _{lN}(\varphi =0)+\text{d}^{3}\sigma _{lN}(\varphi =\pi
172: )+2\text{d}^{3}\sigma _{lN}(\varphi =\pi /2)} \nonumber\\
173: &=&\frac{\varepsilon \,\sigma _{A}( x,Q^{2}) }{\sigma _{T}( x,Q^{2}) +\varepsilon \,\sigma _{L}(
174: x,Q^{2}) }=A(x,Q^{2})\frac{\varepsilon+\varepsilon R(x,Q^{2})}{1+\varepsilon R(x,Q^{2})},
175: \label{9}
176: \end{eqnarray}
177: where $\text{d}^{3}\sigma _{lN}(\varphi )\equiv {{\displaystyle{\text{d}^{3}\sigma _{lN} \over
178: \text{d}x\text{d}Q^{2}\text{d}\varphi }} }( \rho ,x,Q^{2},\varphi)$ and the mean value of $\cos
179: n\varphi$ is
180: \begin{equation}
181: \langle \cos n\varphi \rangle (\rho ,x,Q^{2})= \frac{\int\limits_{0}^{2\pi }\text{d}\varphi \cos
182: n\varphi {\displaystyle {\text{d}^{3}\sigma _{lN} \over \text{d}x\text{d}Q^{2}\text{d}\varphi }} (
183: \rho ,x,Q^{2},\varphi ) }{\int\limits_{0}^{2\pi }\text{d}\varphi {\displaystyle {\text{d}^{3}\sigma
184: _{lN} \over \text{d}x\text{d}Q^{2}\text{d}\varphi }} ( \rho ,x,Q^{2},\varphi ) }.  \label{10}
185: \end{equation}
186: In Eq.~(\ref{9}), the quantities $R(x,Q^{2})$ and $A(x,Q^{2})$ are defined as
187: \begin{eqnarray}
188: R(x,Q^{2})&=&\frac{\sigma_{L}}{\sigma_{T}}(x,Q^{2})=\frac{F_{L}}{F_{T}}(x,Q^{2}), \label{11}\\
189: A(x,Q^{2})&=&\frac{\sigma_{A}}{\sigma_{2}}(x,Q^{2})=2x\frac{F_{A}}{F_{2}}(x,Q^{2}). \label{12}
190: \end{eqnarray}
191: Likewise, we can define the azimuthal asymmetry associated with the $\cos \varphi $ distribution,
192: $A_{\varphi}(\rho ,x,Q^{2})$:
193: \begin{eqnarray}
194: A_{\varphi}(\rho ,x,Q^{2})&=&2\langle \cos \varphi \rangle(\rho ,x,Q^{2})=\frac{2\text{d}^{3}
195: \sigma _{lN}(\varphi =0)-2\text{d}^{3}\sigma _{lN}(\varphi =\pi )}{ \text{d}^{3}\sigma
196: _{lN}(\varphi =0)+\text{d}^{3}\sigma _{lN}(\varphi =\pi
197: )+2\text{d}^{3}\sigma _{lN}(\varphi =\pi /2)} \nonumber\\
198: &=&\frac{2\sqrt{\varepsilon(1+\varepsilon)}\,\sigma _{I}( x,Q^{2}) }{\sigma _{T}( x,Q^{2})
199: +\varepsilon\,\sigma_{L}(x,Q^{2})}=A_{I}(x,Q^{2})\sqrt{\varepsilon(1+\varepsilon)/2}
200: \frac{1+R(x,Q^{2})}{1+\varepsilon R(x,Q^{2})}, \label{13}
201: \end{eqnarray}
202: where
203: \begin{equation}\label{14}
204: A_{I}(x,Q^{2})=2\sqrt{2}\frac{\frac{}{}\sigma_{I}}{\sigma_{2}}(x,Q^{2})=
205: 4\sqrt{2}\,x\frac{F_{I}}{F_{2}}(x,Q^{2}).
206: \end{equation}
207: Remember that $y\ll 1$ in most of the experimentally reachable kinematic range. Taking also into
208: account that $\varepsilon=1+{\cal{O}}(y^{2})$, we find:
209: \begin{equation}\label{15}
210: A_{2\varphi}(\rho ,x,Q^{2})=A(x,Q^{2})+{\cal{O}}(y^{2}), \qquad \qquad \qquad \qquad
211: A_{\varphi}(\rho ,x,Q^{2})=A_{I}(x,Q^{2})+{\cal{O}}(y^{2}).
212: \end{equation}
213: So, like the $\sigma _{2}(x,Q^{2})$ cross section in the $\varphi$-independent case, it is the
214: parameters $A(x,Q^{2})$ and $A_{I}(x,Q^{2})$ that can effectively be measured in the
215: azimuth-dependent production.
216: 
217: In this paper we concentrate on the azimuthal asymmetry $A(x,Q^{2})$ associated with the
218: $\cos2\varphi$-distribution. We have calculated the IC contribution to the asymmetry which is
219: described at the parton level by the photon-quark scattering (QS) mechanism given in
220: Fig.~\ref{Fg.2}. Our main result can be formulated as follows:
221: \begin{itemize}
222: \item[$\star$] Contrary to the basic GF component, the IC mechanism is practically
223: $\cos2\varphi$-independent. This is due to the fact that the QS contribution to the
224: $\sigma_{A}(x,Q^{2})$ cross section is absent (for the kinematic reason) at LO and is negligibly
225: small (of the order of $1\%$) at NLO.
226: \end{itemize}
227: As to the $\varphi$-independent cross sections, our parton level calculations have been compared
228: with the previous results for the IC contribution to $\sigma_{2}(x,Q^{2})$ and
229: $\sigma_{L}(x,Q^{2})$ presented in Refs.~\cite{HM,KS}. Apart from two trivial misprints uncovered
230: in \cite{HM} for $\sigma_{L}(x,Q^{2})$, a complete agreement between all the considered results is
231: found.
232: 
233: Since the GF and QS mechanisms have strongly different $\cos2\varphi$-distributions, we investigate
234: the possibility to discriminate between their contributions using the azimuthal asymmetry
235: $A(x,Q^{2})$. We analyze separately the nonperturbative IC in the framework of the FFNS and the
236: perturbative IC within the VFNS.
237: 
238: The following properties of the nonperturbative IC contribution to the azimuthal asymmetry within
239: the FFNS are found:
240: \begin{itemize}
241: \item  The nonperturbative IC is practically invisible at low $x$, but affects essentially the GF
242: predictions at large $x$. The dominance of the $\cos2\varphi$-independent IC component at large $x$
243: leads to a more rapid (in comparison with the GF predictions) decreasing of $A(x,Q^{2})$ with
244: growth of $x$.%
245: \item  Contrary to the production cross sections, the $\cos 2\varphi$ asymmetry in charm  azimuthal
246: distributions is practically insensitive to radiative corrections at $Q^{2}\sim m^{2}$.
247: Perturbative stability of the combined GF+QS result for $A(x,Q^{2})$ is mainly due to the
248: cancellation of large NLO corrections in Eq.~(\ref{12}).%
249: \item  pQCD predictions for the $\cos 2\varphi$ asymmetry are parametrically stable; the GF+QS
250: contribution to $A(x,Q^{2})$ is practically insensitive to most of the standard uncertainties in
251: the QCD input parameters: $\mu _{R}$, $\mu _{F}$, $\Lambda _{QCD}$ and PDFs.%
252: \item  Nonperturbative corrections to the charm azimuthal asymmetry due to the gluon transverse
253: motion in the target are of the order of $20\%$ at $Q^{2}\leq m^{2}$ and rapidly vanish at $Q^{2}>
254: m^{2}$.
255: \end{itemize}
256: We conclude that the contributions of both GF and IC components to the $\cos 2\varphi$ asymmetry in
257: charm leptoproduction are quantitatively well defined in the FFNS: they are stable, both
258: parametrically and perturbatively, and insensitive (at $Q^{2}> m^{2}$) to the gluon transverse
259: motion in the proton. At large Bjorken $x$, the $A(x,Q^{2})$ asymmetry could be a sensitive probe
260: of the nonperturbative IC.
261: 
262: The perturbative IC has been considered within the VFNS proposed in Refs.~\cite{ACOT,collins}. The
263: following features of the azimuthal asymmetry should be emphasized:
264: \begin{itemize}
265: \item[$\ast$]  Contrary to the nonperturbative IC component, the perturbative one is significant
266: practically at all values of Bjorken $x$ and $Q^{2}>m^{2}$.%
267: \item[$\ast$]  The charm densities of the recent CTEQ and MRST sets of PDFs  lead to a sizeable
268: reduction (by about 1/3) of the GF predictions for the $\cos2\varphi$ asymmetry.
269: \end{itemize}
270: We conclude that impact of the perturbative IC on the $\cos2\varphi$ asymmetry is sizeable in the
271: whole region of $x$ and, for this reason, can easily be detected.
272: 
273: Concerning the experimental aspects, azimuthal asymmetries in charm leptoproduction can, in
274: principle, be measured in the COMPASS experiment at CERN, as well as in future studies at the
275: proposed eRHIC \cite{eRHIC,EIC} and LHeC \cite{LHeC} colliders at BNL and CERN, correspondingly.
276: 
277: The paper is organized as follows. In Section~\ref{part} we analyze the QS and GF parton level
278: predictions for the $\varphi$-dependent charm leptoproduction in the single-particle inclusive
279: kinematics. In particular, we discuss our results for the NLO QS cross sections and compare them
280: with available calculations. Hadron level predictions for $A(x,Q^{2})$ are given in
281: Section~\ref{hadr}. We consider the IC contributions to the asymmetry within the FFNS and VFNS in a
282: wide region of $x$ and $Q^{2}$. Some details of our calculations of the QS cross sections are
283: presented in Appendix~\ref{virt}. An overview of the soft-gluon resummation for the photon-gluon
284: fusion mechanism is given in Appendix~\ref{soft}. Some experimental facts in favor of the
285: nonperturbative IC are briefly listed in Appendix~\ref{exp}.
286: 
287: \section{\label{part} Partonic Cross Sections}
288: \subsection{Quark Scattering Mechanism}
289: \begin{figure}
290: \begin{center}
291: \mbox{\epsfig{file=QSgraph.eps,width=445pt}}
292: \end{center}
293: \caption{\label{Fg.2}\small The LO (a) and NLO (b and c) photon-quark scattering diagrams.}
294: \end{figure}
295: The momentum assignment of the deep inelastic lepton-quark scattering will be denoted as
296: \begin{equation}
297: l(\ell )+Q(k_{Q})\rightarrow l(\ell -q)+Q(p_{Q})+X(p_{X}). \label{17}
298: \end{equation}
299: Taking into account the target mass effects, the corresponding partonic cross section can be
300: written as follows \cite{we6}
301: \begin{equation}\label{18}
302: \frac{\text{d}^{3}\hat{\sigma}_{lQ}}{\text{d}z\text{d}Q^{2}\text{d}\varphi }=\frac{\alpha
303: _{em}}{(2\pi )^{2}}\frac{y^2}{z Q^{2}}\frac{\sqrt{1+4\lambda
304: z^{2}}}{1-\hat{\varepsilon}}\left[\hat{\sigma}_{2,Q}(z,\lambda)-
305: (1-\hat{\varepsilon})\hat{\sigma}_{L,Q}(z,\lambda)+
306: \hat{\varepsilon}\hat{\sigma}_{A,Q}(z,\lambda)\cos 2\varphi+
307: 2\sqrt{\hat{\varepsilon}(1+\hat{\varepsilon})}\hat{\sigma}_{I,Q}(z,\lambda)\cos \varphi\right].
308: \end{equation}
309: In Eq.~(\ref{18}), we use the following definition of partonic kinematic variables:
310: \begin{equation}\label{19}
311: y=\frac{q\cdot k_{Q}}{ \ell\cdot k_{Q} },\qquad \qquad z=\frac{Q^{2}}{2q\cdot k_{Q}},\qquad
312: \qquad\lambda =\frac{m^{2}}{Q^{2}}.
313: \end{equation}
314: In the massive case, the (virtual) photon polarization parameter, $\hat{\varepsilon}$, has the form
315: \cite{we6}
316: \begin{equation}\label{20}
317: \hat{\varepsilon}=\frac{2(1-y-\lambda z^{2}y^{2})}{1+(1-y)^2+2\lambda z^{2}y^{2}}.
318: \end{equation}
319: At leading order, ${\cal O}(\alpha _{em})$, the only quark scattering subprocess is
320: \begin{equation}
321: \gamma ^{*}(q)+Q(k_{Q})\rightarrow Q(p_{Q}).  \label{21}
322: \end{equation}
323: The $\gamma ^{*}Q$ cross sections, $\hat{\sigma}_{k,Q}^{(0)}$ ($k=2,L,A,I$), corresponding to the
324: Born diagram (see Fig.~\ref{Fg.2}a) are:
325: \begin{eqnarray}
326: \hat{\sigma}_{2,Q}^{(0)}(z,\lambda)&=&\hat{\sigma}_{B}(z)\sqrt{1+4\lambda z^{2}}\,\delta(1-z), \nonumber\\
327: \hat{\sigma}_{L,Q}^{(0)}(z,\lambda)&=&\hat{\sigma}_{B}(z)\frac{4\lambda z^{2}}{\sqrt{1+4\lambda
328: z^{2}}}\,\delta(1-z), \label{22}\\
329: \hat{\sigma}_{A,Q}^{(0)}(z,\lambda)&=&\hat{\sigma}_{I,Q}^{(0)}(z,\lambda)=0,\nonumber
330: \end{eqnarray}
331: with
332: \begin{equation}\label{23}
333: \hat{\sigma}_{B}(z)=\frac{(2\pi)^2e_{Q}^{2}\alpha _{em}}{Q^{2}}\,z,
334: \end{equation}
335: where $e_{Q}$ is the quark charge in units of electromagnetic coupling constant.
336: 
337: To take into account the NLO ${\cal O}(\alpha _{em}\alpha _{s})$ contributions, one needs to
338: calculate the virtual corrections to the Born process  (given in Fig.~\ref{Fg.2}c) as well as the
339: real gluon emission (see Fig.~\ref{Fg.2}b):
340: \begin{equation}
341: \gamma ^{*}(q)+Q(k_{Q})\rightarrow Q(p_{Q})+g(p_{g}).  \label{24}
342: \end{equation}
343: 
344: The NLO $\varphi$-dependent cross sections, $\hat{\sigma}_{A,Q}^{(1)}$ and
345: $\hat{\sigma}_{I,Q}^{(1)}$, are described by the real gluon emission only. Corresponding
346: contributions are free of any type of singularities and the quantities $\hat{\sigma}_{A,Q}^{(1)}$
347: and $\hat{\sigma}_{I,Q}^{(1)}$ can be calculated directly in four dimensions.
348: 
349: In the $\varphi$-independent case, $\hat{\sigma}_{2,Q}^{(1)}$ and $\hat{\sigma}_{L,Q}^{(1)}$, we
350: also work in four dimensions. The virtual contribution (Fig.~\ref{Fg.2}c) contains ultraviolet (UV)
351: singularity that is removed using the on-mass-shell regularization scheme. In particular, we
352: calculate the absorptive part of the Feynman diagram which has no UV divergences.  The real part is
353: then obtained by using the appropriate dispersion relations. As to the infrared (IR) singularity,
354: it is regularized with the help of an infinitesimal gluon mass. This IR divergence is cancelled
355: when we add the bremsstrahlung contribution (Fig.~\ref{Fg.2}b). Some details of our calculations
356: are given in Appendix \ref{virt}.
357: 
358: The final (real+virtual) results for $\gamma ^{*}Q$ cross sections can be cast into the following
359: form:
360: \begin{eqnarray}
361: \hat{\sigma}_{2,Q}^{(1)}(z,\lambda)=\frac{\alpha_{s}}{2\pi}C_{F}\hat{\sigma}_{B}(1)
362: \sqrt{1+4\lambda}\,\delta(1-z)\Bigl\{-2+4\ln\lambda-\sqrt{1+4\lambda }\,\ln r+
363: \frac{1+2\lambda}{\sqrt{1+4\lambda}}\Bigl[2\text{Li}_{2}(r^{2})+4\text{Li}_{2}(-r)&& \nonumber\\
364: +3\ln^{2}(r)-4\ln r+4\ln r \ln(1+4\lambda)-2\ln r\ln\lambda\Bigr]\Bigr\}&& \nonumber \\
365: +\frac{\alpha_{s}}{4\pi}C_{F}\hat{\sigma}_{B}(z)\frac{1}{(1+4\lambda z^{2})^{3/2}}\biggl\{
366: \frac{1}{\left[1-(1-\lambda)z\right]^{2}}\Bigl[1-3z-4z^{2}+6z^{3}+8z^{4}-8z^{5} \qquad \qquad \qquad \;&& \nonumber \\
367: +6\lambda z\left(3-18z+13z^{2}+10z^{3}-8z^{4}\right) \qquad \qquad&&  \nonumber \\
368: +4\lambda^{2}z^{2}\left(8-77z+65z^{2}-2z^{3}\right)\biggr. \qquad \qquad \qquad \; \,&&  \label{25} \\
369: +16\lambda^{3}z^{3}\left(1-21z+12z^{2}\right)-128\lambda^{4}z^{5}\Bigr] \qquad \quad \; \,&&   \nonumber \\
370: +\frac{2\ln D(z,\lambda)}{\sqrt{1+4\lambda z^{2}}}\Bigl[-\left(1+z+2z^{2}+2z^{3}\right)+2\lambda
371: z\left(2-11z-11z^{2}\right)+8\lambda^{2}z^{2}\left(1-9z\right)\Bigr]&& \nonumber \\
372: -\frac{8(1+4\lambda)^{2}z^{4}}{\left(1-z\right)_{+}}-
373: \frac{4(1+2\lambda)(1+4\lambda)^{2}z^{4}}{\sqrt{1+4\lambda z^{2}}}\frac{\ln
374: D(z,\lambda)}{\left(1-z\right)_{+}}\biggr\},&& \nonumber
375: \end{eqnarray}
376: \begin{eqnarray}
377: \hat{\sigma}_{L,Q}^{(1)}(z,\lambda)=\frac{\alpha_{s}}{\pi}C_{F}\hat{\sigma}_{B}(1)
378: \frac{2\lambda}{\sqrt{1+4\lambda}}\delta(1-z)\Bigl\{-2+4\ln\lambda-\frac{4\lambda}{\sqrt{1+4\lambda
379: }}\,\ln r+\frac{1+2\lambda}{\sqrt{1+4\lambda}}\Bigl[2\text{Li}_{2}(r^{2})+4\text{Li}_{2}(-r)&& \nonumber\\
380: +3\ln^{2}(r)-4\ln r+4\ln r \ln(1+4\lambda)-2\ln r\ln\lambda\Bigr]\Bigr\}&& \nonumber \\
381: +\frac{\alpha_{s}}{\pi}C_{F}\hat{\sigma}_{B}(z)\frac{1}{(1+4\lambda z^{2})^{3/2}}\biggl\{
382: \frac{z}{\left[1-(1-\lambda)z\right]^{2}}\Bigl[(1-z)^{2}-
383: \lambda z\left(13-19z-2z^{2}+8z^{3}\right)\Bigr. \qquad \qquad&&   \nonumber \\
384: -2\lambda^{2}z^{2}\left(31-39z+8z^{2}\right)\Bigr. \qquad \qquad \qquad \qquad  \quad \; \,&& \label{26} \\
385: -8\lambda^{3}z^{3}\left(10-7z\right)-32\lambda^{4}z^{4}\Bigr]\Bigr.  \qquad \qquad \qquad  \qquad&&   \nonumber \\
386: -\frac{2\lambda z^{2}\ln D(z,\lambda)}{\sqrt{1+4\lambda z^{2}}}\left[3+3z+16\lambda z\right]\Bigr.
387: \qquad \qquad \qquad \qquad \; \,&& \nonumber \\
388: -\frac{8\lambda(1+4\lambda)z^{4}}{\left(1-z\right)_{+}}-
389: \frac{4\lambda(1+2\lambda)(1+4\lambda)z^{4}}{\sqrt{1+4\lambda z^{2}}}\frac{\ln
390: D(z,\lambda)}{\left(1-z\right)_{+}}\biggr\},&& \nonumber
391: \end{eqnarray}
392: \begin{equation}\label{27}
393: \hat{\sigma}_{A,Q}^{(1)}(z,\lambda)=\frac{\alpha_{s}}{2\pi}C_{F}\hat{\sigma}_{B}(z)\frac{z(1-z)}{(1+4\lambda
394: z^{2})^{3/2}}\biggl\{ \frac{1}{\left[1-(1-\lambda)z\right]}\left[1+2\lambda(4-3z)+8\lambda^2
395: z\right]+\frac{2\lambda \ln D(z,\lambda)}{\sqrt{1+4\lambda z^{2}}}\left[2+z+4\lambda
396: z\right]\biggr\},
397: \end{equation}
398: \begin{eqnarray}
399: \hat{\sigma}_{I,Q}^{(1)}(z,\lambda)=\frac{\alpha_{s}}{8\sqrt{2}}C_{F}\hat{\sigma}_{B}(z)
400: \frac{1}{(1+4\lambda z^{2})^{2}}\frac{\sqrt{z}}{\left[1-(1-\lambda)z\right]^{3/2}}\biggl\{
401: -(1-z)(1+2z)-4\lambda z\left(10-10z-z^{2}+2z^{3}\right)  \qquad \quad \; \;&& \nonumber \\
402: -8\lambda^{2}z^{2}\left(25-29z+8z^{2}\right)-96\lambda^{3}z^{3}\left(3-2z\right)-
403: 128\lambda^{4}z^{4}\biggr.\;&& \label{28} \\
404: +8\sqrt{\lambda z\left[1-(1-\lambda)z\right]}\left[1-z^{2}+\lambda
405: z(13-11z)+4\lambda^{2}z^{2}(7-4z)+16\lambda^{3}z^{3}\right]\biggr\}.&& \nonumber
406: \end{eqnarray}
407: In Eqs.~(\ref{25}-\ref{28}), $C_{F}=(N_{c}^{2}-1)/(2N_{c})$, where $N_{c}$ is number of colors,
408: while
409: \begin{equation}\label{29}
410: D(z,\lambda)=\frac{1+2\lambda z -\sqrt{1+4\lambda z^{2}}}{1+2\lambda z +\sqrt{1+4\lambda
411: z^{2}}},\qquad \qquad \qquad \qquad \qquad
412: r=\sqrt{D(z=1,\lambda)}=\frac{\sqrt{1+4\lambda}-1}{\sqrt{1+4\lambda}+1}.
413: \end{equation}
414: The so-called "plus" distributions are defined by
415: \begin{equation}\label{30}
416: \left[g(z)\right]_{+}=g(z)-\delta(1-z)\int\limits_{0}^{1}\text{d}\zeta\,g(\zeta).
417: \end{equation}
418: For any sufficiently regular test function $h(z)$, Eq.~(\ref{30}) gives
419: \begin{equation}\label{31}
420: \int\limits_{a}^{1}\text{d}z\,h(z)\left[\frac{\ln^{k}(1-z)}{1-z}\right]_{+}=
421: \int\limits_{a}^{1}\text{d}z\frac{\ln^{k}(1-z)}{1-z}\left[h(z)-h(1)\right]+
422: h(1)\frac{\ln^{k+1}(1-a)}{k+1}.
423: \end{equation}
424: 
425: To perform a numerical investigation of the inclusive partonic cross sections, $\hat{\sigma}_{k,Q}$
426: ($k=T,L,A,I$),{\large \ } it is convenient to introduce the dimensionless coefficient functions
427: $c_{k,Q}^{(n,l)}$,
428: \begin{equation}\label{32}
429: \hat{\sigma}_{k,Q}(\eta ,\lambda ,\mu ^{2})=\frac{e_{Q}^{2}\alpha _{em}\alpha _{s}(\mu
430: ^{2})}{m^{2}}\sum_{n=0}^{\infty }\left( 4\pi \alpha _{s}(\mu ^{2})\right)
431: ^{n}\sum_{l=0}^{n}c_{k,Q}^{(n,l)}(\eta ,\lambda )\ln ^{l}\left( \frac{\mu ^{2}}{m^{2}}\right) ,
432: \end{equation}
433: where $\mu$ is a factorization scale (we use $\mu=\mu_{F}=\mu_{R}$) and the variable $\eta$
434: measures the distance to the partonic threshold:
435: \begin{equation}\label{33}
436: \eta =\frac{s}{m^{2}}-1=\frac{1-z}{\lambda z},\qquad \qquad \qquad \qquad \qquad s =(q+k_{Q})^{2}.
437: \end{equation}
438: 
439: Our analysis of the quantity $c_{A,Q}^{(0,0)}(\eta ,\lambda)$ is given in Fig.~\ref{Fg.3}. One can
440: see that $c_{A,Q}^{(0,0)}$ is negative at low $Q^{2}$ ($\lambda^{-1}\lesssim 1$) and positive at
441: high $Q^{2}$ ($\lambda^{-1}> 20$). For the intermediate values of $Q^{2}$, $c_{A,Q}^{(0,0)}(\eta
442: ,\lambda)$ is an alternating function of $\eta$.
443: 
444: Our results for the coefficient function $c_{I,Q}^{(0,0)}(\eta,\lambda )$ at several values of
445: $\lambda$ are presented in Fig.~\ref{Fg.3}. It is seen that $c_{I,Q}^{(0,0)}$ is negative at all
446: values of $\eta$ and $\lambda$. Note also the threshold behavior of the coefficient function:
447: \begin{equation}\label{34}
448: c_{I,Q}^{(0,0)}(\eta\rightarrow 0,\lambda )=-\sqrt{2}\pi^{2}C_{F}\frac{\sqrt{\lambda}}{1+4\lambda}+
449: {\cal{O}}(\eta).
450: \end{equation}
451: This quantity takes its minimum value at $\lambda_{m}=1/4$: $c_{I,Q}^{(0,0)}(\eta = 0,\lambda_{m})
452: =-\pi^{2}C_{F}/\left(2\sqrt{2}\right)$.
453: 
454: \begin{figure}
455: \begin{center}
456: \begin{tabular}{cc}
457: \mbox{\epsfig{file=cAQ.eps,width=250pt}}
458: & \mbox{\epsfig{file=cIQ.eps,width=253pt}}\\
459: \end{tabular}
460: \caption{\label{Fg.3}\small  $c_{A,Q}^{(0,0)}(\eta,\lambda )$ and $c_{I,Q}^{(0,0)}(\eta,\lambda )$
461: coefficient functions at several values of $\lambda$.}
462: \end{center}
463: \end{figure}
464: 
465: \subsection{Comparison with Available Results}
466: For the first time, the NLO ${\cal O}(\alpha_{em}\alpha_{s})$ corrections to the
467: $\varphi$-independent IC contribution have been calculated a long time ago by Hoffmann and Moore
468: (HM) \cite{HM}. However, authors of Ref.~\cite{HM} don't give explicitly their definition of the
469: partonic cross sections that leads to a confusion in interpretation of the original HM results. To
470: clarify the situation, we need first to derive the relation between the lepton-quark DIS cross
471: section, $\text{d}\hat{\sigma}_{lQ}$, and the partonic cross sections, $\sigma^{(2)}$ and
472: $\sigma^{(L)}$, used in \cite{HM}. Using Eqs.~(C.1) and (C.5) in Ref.~\cite{HM}, one can express
473: the HM tensor $\sigma_{R}^{\mu\nu}$ in terms of "our" cross sections $\hat{\sigma}_{2,Q}$ and
474: $\hat{\sigma}_{L,Q}$ defined by Eq.~(\ref{18}) in the present paper. Comparing the obtained results
475: with the corresponding definition of $\sigma_{R}^{\mu\nu}$ via the HM cross sections $\sigma^{(2)}$
476: and $\sigma^{(L)}$ (given by Eqs.~(C.16) and (C.17) in Ref.~\cite{HM}), we find that
477: \begin{eqnarray}
478: \hat{\sigma}_{2,Q}(z,\lambda)&\equiv &\hat{\sigma}_{B}(z)\sqrt{1+4\lambda
479: z^{2}}\,\sigma^{(2)}(z,\lambda),  \label{35} \\
480: \hat{\sigma}_{L,Q}(z,\lambda)&\equiv &\frac{2\hat{\sigma}_{B}(z)}{\sqrt{1+4\lambda
481: z^{2}}}\left[\sigma^{(L)}(z,\lambda)+2\lambda z^{2}\sigma^{(2)}(z,\lambda)\right]. \label{36}
482: \end{eqnarray}
483: Now we are able to compare our results with original HM ones. It is easy to see that the LO cross
484: sections (defined by Eqs.~(37) in \cite{HM} and Eqs.~(\ref{22}) in our paper) obey both above
485: identities.  Comparing with each other the quantities $\sigma^{(2)}_{1}$ and
486: $\hat{\sigma}_{2,Q}^{(1)}$ (given by Eq.~(51) in \cite{HM} and Eq.~(\ref{25}) in this paper,
487: respectively), we find that identity (\ref{35}) is satisfied at NLO too. The situation with
488: longitudinal cross sections is more complicated. We have uncovered two misprints in the NLO
489: expression for $\sigma^{(L)}$ given by Eq.~(52) in \cite{HM}. First, the r.h.s. of this Eq. must be
490: multiplied by $z$. Second, the sign in front of the last term (proportional to $\delta (1-z)$) in
491: Eq.~(52) in Ref.~\cite{HM} must be changed \footnote{Note that this term originates from virtual
492: corrections and the virtual part of the longitudinal cross section given by Eq.~(39) in
493: Ref.~\cite{HM} also has wrong sign. See Appendix \ref{virt} for more details.}. Taking into account
494: these typos, we find that relation (\ref{36}) holds at NLO as well. So, our calculations of
495: $\hat{\sigma}_{2,Q}$ and $\hat{\sigma}_{L,Q}$ agree with the HM results.
496: 
497: Recently, the heavy quark initiated contributions to the $\varphi$-independent DIS structure
498: functions, $F_{2}$ and $F_{L}$, have been calculated by Kretzer and Schienbein (KS) \cite{KS}. The
499: final KS results are expressed in terms of the parton level structure functions $\hat{H}^{q}_{1}$
500: and $\hat{H}^{q}_{2}$. Using the definition of $\hat{H}^{q}_{1}$ and $\hat{H}^{q}_{2}$ given by
501: Eqs.~(7, 8) in Ref.~\cite{KS}, we obtain that
502: \begin{equation}\label{37}
503: \hat{\sigma}_{T,Q}(z,\lambda)\equiv
504: \frac{\alpha_{s}}{2\pi}\frac{\hat{\sigma}_{B}(z)}{\sqrt{1+4\lambda}}\frac{\hat{H}^{q}_{1}(\xi^{\prime},
505: \lambda)}{\sqrt{1+4\lambda z^{2}}},\qquad \qquad \qquad \qquad \hat{\sigma}_{2,Q}(z,\lambda)\equiv
506: \frac{\alpha_{s}}{2\pi}\hat{\sigma}_{B}(z)\sqrt{\frac{1+4\lambda}{1+4\lambda
507: z^{2}}}\,\hat{H}^{q}_{2}(\xi^{\prime},\lambda),
508: \end{equation}
509: where $\hat{\sigma}_{T,Q}=\hat{\sigma}_{2,Q}-\hat{\sigma}_{L,Q}$ and $\hat{\sigma}_{L,Q}$ are
510: defined by Eq.~(\ref{18}) in our paper and
511: $\xi^{\prime}=z\left(1+\sqrt{1+4\lambda}\right)\left/\left(1+\sqrt{1+4\lambda
512: z^{2}}\right)\right.$. To test identities (\ref{37}), one needs only to rewrite the NLO expressions
513: for the functions $\hat{H}^{q}_{1}(\xi^{\prime},\lambda)$ and
514: $\hat{H}^{q}_{2}(\xi^{\prime},\lambda)$ (given in Appendix C in Ref.~\cite{KS}) in terms of
515: variables $z$ and $\lambda$. Our analysis shows that relations (\ref{37}) hold at both LO and NLO.
516: Hence we coincide with the KS predictions for the $\gamma^{*}Q$ cross sections.
517: 
518: However, we disagree with the conclusion of Ref.~\cite{KS} that there are errors in the NLO
519: expression for $\sigma^{(2)}$ given in Ref.~\cite{HM} \footnote{In detail, the KS point of view on
520: the HM results is presented in PhD thesis \cite{KS-thesis}, pp.~158-160.}. As explained above, a
521: correct interpretation of the quantities $\sigma^{(2)}$ and $\sigma^{(L)}$ used in \cite{HM} leads
522: to a complete agreement between the HM, KS and our results for $\varphi$-independent cross
523: sections.
524: 
525: As to the $\varphi$-dependent DIS, pQCD predictions for the $\gamma^{*}Q$ cross sections
526: $\hat{\sigma}_{A,Q}(z,\lambda)$ and $\hat{\sigma}_{I,Q}(z,\lambda)$ in the case of arbitrary values
527: of $m^{2}$ and $Q^{2}$ are not, to our knowledge, available in the literature. For this reason, we
528: have performed several cross checks of our results against well known calculations in two limits:
529: $m^{2}\rightarrow 0$ and $Q^{2}\rightarrow 0$. In particular, in the chiral limit, we reproduce the
530: original results of Georgi and Politzer \cite{GP} and M\'{e}ndez \cite{Mendez} for
531: $\hat{\sigma}_{I,Q}(z,\lambda\rightarrow 0)$ and $\hat{\sigma}_{A,Q}(z,\lambda\rightarrow 0)$. In
532: the case of $Q^{2}\rightarrow 0$, our predictions for $\hat{\sigma}_{2,Q}(s,Q^{2}\rightarrow 0)$
533: and $\hat{\sigma}_{A,Q}(s,Q^{2}\rightarrow 0)$ given by Eqs.~(\ref{25},\ref{27}) reduce to the QED
534: textbook results for the Compton scattering of polarized photons \cite{Fano}.
535: 
536: \subsection{Photon-Gluon Fusion}
537: The gluon fusion component of the semi-inclusive DIS is the following parton level interaction:
538: \begin{equation}
539: l(\ell )+g(k_{g})\rightarrow l(\ell -q)+Q(p_{Q})+X[\overline{Q}](p_{X}). \label{d1}
540: \end{equation}
541: Corresponding lepton-gluon cross section, $\text{d}\hat{\sigma}_{lg}$, has the following
542: decomposition in terms of the helicity $\gamma ^{*}g$ cross sections:
543: \begin{equation}\label{d2}
544: \frac{\text{d}^{3}\hat{\sigma}_{lg}}{\text{d}z\text{d}Q^{2}\text{d}\varphi }=\frac{\alpha
545: _{em}}{(2\pi )^{2}}\frac{1}{z Q^{2}}\frac{y^2}{1-\varepsilon}\left[\hat{\sigma}_{2,g}(z,\lambda)-
546: (1-\varepsilon)\hat{\sigma}_{L,g}(z,\lambda)+ \varepsilon\hat{\sigma}_{A,g}(z,\lambda)\cos
547: 2\varphi+ 2\sqrt{\varepsilon(1+\varepsilon)}\hat{\sigma}_{I,g}(z,\lambda)\cos \varphi\right],
548: \end{equation}
549: where the quantity $\varepsilon$ is defined by Eq.~(\ref{3}) with $y=\left.(q\cdot k_{g})\right
550: /(\ell\cdot k_{g})$.
551: \begin{figure}
552: \begin{center}
553: \mbox{\epsfig{file=GFgraph.eps,width=355pt}}
554: \end{center}
555: \caption{\label{Fg.4}\small The LO photon-gluon fusion diagrams.}
556: \end{figure}
557: 
558: At LO, ${\cal O}(\alpha_{em}\alpha_{s})$, the only gluon fusion subprocess responsible for heavy
559: flavor production is
560: \begin{equation}\label{38}
561: \gamma ^{*}(q)+g(k_{g})\rightarrow Q(p_{Q})+\overline{Q}(p_{\stackrel{\_}{Q}}).
562: \end{equation}
563: The $\gamma ^{*}g$ cross sections, $\hat{\sigma}_{k,g}^{(0)}$ ($k=2,L,A,I$), corresponding to the
564: Born diagrams given in Fig.~\ref{Fg.4} have the form \cite{LW1,Watson}:
565: \begin{eqnarray}
566: \hat{\sigma}_{2,g}^{(0)}(z,\lambda)&=&\frac{\alpha_{s}}{2\pi}\hat{\sigma}_{B}(z)
567: \Bigl\{\left[(1-z)^{2}+z^{2}+4\lambda z(1-3z)-8\lambda^{2}z^{2}\right]
568: \ln\frac{1+\beta_{z}}{1-\beta_{z}}-
569: \left[1+4z(1-z)(\lambda-2)\right]\beta_{z}\Bigr\},  \nonumber\\
570: \hat{\sigma}_{L,g}^{(0)}(z,\lambda)&=&\frac{2\alpha_{s}}{\pi}\hat{\sigma}_{B}(z)z
571: \Bigl\{-2\lambda z\ln\frac{1+\beta_{z}}{1-\beta_{z}}+\left(1-z\right)\beta_{z}\Bigr\},  \nonumber\\
572: \hat{\sigma}_{A,g}^{(0)}(z,\lambda)&=&\frac{\alpha_{s}}{\pi}\hat{\sigma}_{B}(z)z
573: \Bigl\{2\lambda\left[1-2z(1+\lambda)\right]\ln\frac{1+\beta_{z}}{1-\beta_{z}}+
574: (1-2\lambda)(1-z)\beta_{z}\Bigr\},  \label{39} \\
575: \hat{\sigma}_{I,g}^{(0)}(z,\lambda)&=&0,  \nonumber
576: \end{eqnarray}
577: where $\hat{\sigma}_{B}(z)$ is defined by Eq.~(\ref{23}) and the following notations are used:
578: \begin{equation}\label{40}
579: z=\frac{Q^{2}}{2q\cdot k_{g}},\qquad \qquad\lambda =\frac{m^{2}}{Q^{2}}, \qquad \qquad
580: \beta_{z}=\sqrt{1-\frac{4\lambda z}{1-z}}.
581: \end{equation}
582: Note that the $\cos \varphi $ dependence vanishes in the GF mechanism due to the $Q\leftrightarrow
583: \overline{Q}$ symmetry which, at leading order, requires invariance under $\varphi \rightarrow
584: \varphi +\pi$ \cite{LW2}.
585: 
586: As to the NLO results, presently, only $\varphi$-independent quantities $\hat{\sigma}_{2,g}^{(1)}$
587: and $\hat{\sigma}_{L,g}^{(1)}$ are known exactly \cite{LRSN}. For this reason, we will use in our
588: analysis the so-called soft-gluon approximation for the NLO $\gamma ^{*}g$ cross sections (see
589: Appendix \ref{soft}). As shown in Refs.~\cite{Laenen-Moch,we2,we4}, at energies not so far from the
590: production threshold, the soft-gluon radiation is the dominant perturbative mechanism in the
591: $\gamma ^{*}g$ interactions.
592: 
593: \section{\label{hadr}Hadron Level Results}
594: \subsection{\label{ha}Fixed Flavor Number Scheme and Nonperturbative Intrinsic Charm}
595: In the fixed flavor number scheme \footnote{This approach is sometimes referred to as the
596: fixed-order perturbation theory (FOPT).}, the wave function of the proton consists of light quarks
597: $u,d,s$ and gluons $g$. Heavy flavor production in DIS is dominated by the gluon fusion mechanism.
598: Corresponding hadron level cross sections, $\sigma_{k,GF}(x,\lambda)$, have the form
599: \begin{eqnarray}
600: \sigma_{k,GF}(x,\lambda)&=&\int\limits_{\chi}^{1}\text{d}z\,g(z,\mu_{F})\,\hat{\sigma}_{k,g}
601: \!\left(x/z,\lambda,\mu_{F}\right), \qquad \qquad \qquad (k=2,L,A,I), \label{41}\\
602: \chi&=&x\left(1+4\lambda\right), \label{42}
603: \end{eqnarray}
604: where $g(z,\mu _{F})$ describes gluon density in the proton evaluated at a factorization scale $\mu
605: _{F}$. The lowest order GF cross sections, $\hat{\sigma}_{k,g}^{(0)}$ ($k=2,L,A,I$), are given by
606: Eqs.~(\ref{39}). The NLO results, $\hat{\sigma}_{k,g}^{(1)}$, to the next-to-leading logarithmic
607: accuracy are presented in Appendix \ref{soft}.
608: 
609: We neglect the $\gamma ^{*}q(\bar{q})$ fusion subprocesses. This is justified as their
610: contributions to heavy quark leptoproduction vanish at LO and are small at NLO \cite{LRSN}.
611: 
612: In the FFNS, the intrinsic heavy flavor component of the proton wave function is generated by
613: $gg\rightarrow Q\bar{Q}$ fluctuations where the gluons are coupled to different valence quarks. In
614: the present paper, this component is referred to as the nonperturbative intrinsic charm (bottom).
615: The probability of the corresponding five-quark Fock state, $\left|uudQ\bar{Q}\right\rangle$, is of
616: higher twist since it scales as $\Lambda_{QCD}^{2}\left/m^{2}\right.$ \cite{polyakov}. However,
617: since all of the quarks tend to travel coherently at same rapidity in the
618: $\left|uudQ\bar{Q}\right\rangle$ bound state, the heaviest constituents carry the largest momentum
619: fraction. For this reason, the heavy flavor distribution function has a more "hard" $z$-behavior
620: than the light parton densities. Since all of the densities vanish at $z\to 1$, the hardest PDF
621: becomes dominant at sufficiently large $z$ independently of normalization.
622: 
623: Convolution of PDFs with partonic cross sections does not violate this observation. In particular,
624: assuming a gluon density $g(z)\sim (1-z)^n$ (where $n=3-5$), we obtain that the LO GF contribution
625: to $F_{2}$ scales as $(1-\chi)^{n+3/2}$ at $\chi\to 1$, where $\chi$ is defined by Eq.~(\ref{42}).
626: In the case of Hoffman and Moore charm density (see below), the LO IC contribution is proportional
627: to $(1-\chi)$ at $\chi\to 1$. It is easy to see that, independently of normalizations, the IC
628: contribution to be dominate over the more "soft" GF component at large enough $x$.
629: 
630: For the first time, the intrinsic charm momentum distribution in the five-quark state
631: $\left|uudc\bar{c}\right\rangle$ was derived by Brodsky, Hoyer, Peterson and Sakai (BHPS) in the
632: framework of a light-cone model \cite{BHPS,BPS}. Neglecting the transverse motion of constituents,
633: they have obtained in the heavy quark limit that
634: \begin{equation}\label{43}
635: c(z)=\frac{N_{5}}{6}z^{2}\left[6z(1+z)\ln z+(1-z)(1+10z+z^{2})\right],
636: \end{equation}
637: where $N_{5}=36$ corresponds to a $1\%$ probability for IC in the nucleon:
638: $\int^{1}_{0}c(z)\text{d}z=0.01$.
639: 
640: Hoffmann and Moore (HM) \cite{HM} incorporated mass effects in the BHPS approach. They first
641: introduced a mass scaling variable $\xi$,
642: \begin{equation}\label{44}
643: \xi=\frac{2ax}{1+\sqrt{1+4\lambda_{N}x^{2}}},\qquad \qquad \qquad \qquad
644: a=\frac{1+\sqrt{1+4\lambda}}{2},
645: \end{equation}
646: where $\lambda_{N}=m^{2}_{N}\left/Q^{2}\right.$. To provide correct threshold behavior of the charm
647: density, the constraint $\xi\leq\gamma<1$ was imposed where
648: \begin{equation}\label{45}
649: \gamma=\frac{2a\hat{x}}{1+\sqrt{1+4\lambda_{N}\hat{x}^{2}}},\qquad \qquad \qquad \qquad
650: \hat{x}=\frac{1}{1+4\lambda-\lambda_{N}}.
651: \end{equation}
652: Resulting charm distribution function, $c(\xi,\gamma)$, has the following form in the HM approach:
653: \begin{equation}\label{46}
654: c(\xi,\gamma)=\left\{%
655: \begin{array}{ll}
656: c(\xi)-\displaystyle{\frac{\xi}{\gamma}}c(\gamma),&\qquad \xi\leq\gamma\\
657: 0,&\qquad \xi>\gamma
658: \end{array}%
659: \right.
660: \end{equation}
661: with $c(\xi)$ defined by Eq.~(\ref{43}). Corresponding hadron level cross sections for the
662: $(c+\bar{c})$ production, $\sigma_{k,QS}(x,\lambda)$, due to the heavy quark scattering (QS)
663: mechanism, are
664: \begin{equation}\label{47}
665: \sigma_{k,QS}(x,\lambda)=\int\limits_{\xi}^{\gamma}\frac{\text{d}z}{\sqrt{1+4\lambda\xi^{2}/z^{2}}}
666: \,c_{+}(z,\gamma)\,\hat{\sigma}_{k,c}\!\left(\xi/z,\lambda\right),\qquad \qquad \qquad (k=2,L,A,I),
667: \end{equation}
668: where the charm density $c_{+}(z,\gamma)\equiv c(z,\gamma)+\bar{c}(z,\gamma)$. The LO and NLO
669: expressions for the partonic cross sections $\hat{\sigma}_{k,c}(z,\lambda)$ are given by
670: Eqs.~(\ref{22}) and (\ref{25}-\ref{28}), respectively.
671: 
672: Note also that, in the FFNS, the full cross section for the charm production,
673: $\sigma_{k}(x,\lambda)$, is simply a sum of the GF and IC components:
674: \begin{equation}\label{48}
675: \sigma_{k}(x,\lambda)=\sigma_{k,GF}(x,\lambda)+\sigma_{k,QS}(x,\lambda), \qquad \qquad \qquad
676: \qquad \qquad (k=2,L,A,I).
677: \end{equation}
678: \begin{figure}
679: \begin{center}
680: \begin{tabular}{cc}
681: \mbox{\epsfig{file=AcF.eps,width=250pt}}
682: & \mbox{\epsfig{file=AcV.eps,width=250pt}}\\
683: \end{tabular}
684: \caption{\label{Fg.5}\small The quantity
685: $\bigl(\sigma^{(1)}_{A,QS}\,\bigl/\sigma^{(0+1)}_{2,QS}\bigr)(x,\lambda)$ in the FFNS with the HM
686: \cite{HM} charm density (\emph{left panel}) and in the VFNS with the CTEQ5M charm distribution
687: function (\emph{right panel}).}
688: \end{center}
689: \end{figure}
690: 
691: Let us discuss the FFNS predictions for the hadron level asymmetry parameter $A(x,Q^{2})$ defined
692: by Eq.~(\ref{12}). First we consider the ratio
693: $\bigl(\sigma^{(1)}_{A,QS}\,\bigl/\sigma^{(0+1)}_{2,QS}\bigr)(x,\lambda)$, i.e., the mere IC
694: contribution to $A(x,Q^{2})$. In Fig.~\ref{Fg.5}, the $x$-behavior of this quantity at various
695: values of $\lambda$ is presented. One can see that the ratio
696: $\bigl(\sigma^{(1)}_{A,QS}\,\bigl/\sigma^{(0+1)}_{2,QS}\bigr)(x,\lambda)$ is negligibly small (of
697: the order of $1\%$) practically at all values of $Q^{2}>m^{2}$. Note that this fact is independent
698: of the charm density we use (see, for instance, the right panel in Fig.~\ref{Fg.5}), but is only
699: due to the smallness of the partonic cross section $\hat{\sigma}^{(1)}_{A,c}(z,\lambda)$
700: \cite{we6}. So, the quantity $\sigma_{A,QS}(x,\lambda)$ is exactly zero at LO \footnote{Remember
701: that the LO quantity $\hat{\sigma}^{(0)}_{A,c}(z,\lambda)$ vanishes for the kinematic reason, see
702: Eqs.~(\ref{22}).} and negligibly small at NLO. This implies that the IC contribution has no
703: practically $\cos2\varphi$-dependence and, for this reason, we will neglect both
704: $\hat{\sigma}_{A,c}(z,\lambda)$ and $\sigma_{A,QS}(x,\lambda)$ cross sections in our further
705: analysis.
706: \begin{figure}
707: \begin{center}
708: \begin{tabular}{cc}
709: \mbox{\epsfig{file=AgcF1B.eps,width=252pt}}
710: & \mbox{\epsfig{file=AgcF4B.eps,width=245pt}}\\
711: \mbox{\epsfig{file=AgcF10B.eps,width=243pt}}
712: & \mbox{\epsfig{file=AgcF20B.eps,width=245pt}}\\
713: \mbox{\epsfig{file=AgcF50B.eps,width=250pt}}
714: & \mbox{\epsfig{file=AgcF100B.eps,width=247pt}}\\
715: \end{tabular}
716: \caption{\label{Fg.6}\small Azimuthal asymmetry parameter $A(x,\lambda)$ in the FFNS at several
717: values of $\lambda$ in the case of $\int^{1}_{0}c(z)\text{d}z=1\%$. The following contributions are
718: plotted: $\text{GF}^{\text{(LO)}}$ (solid lines), $\text{GF}^{\text{(LO)}}$+$k_{T}$-kick (dotted
719: lines), $\text{GF}^{\text{(NLO)}}$ (dashed lines),
720: $\text{GF}^{\text{(LO)}}$+$\text{QS}^{\text{(LO)}}$ (dash-dotted lines) and
721: $\text{GF}^{\text{(NLO)}}$+$\text{QS}^{\text{(NLO)}}$ (long-dashed lines).}
722: \end{center}
723: \end{figure}
724: 
725: Fig.~\ref{Fg.6} shows $A(x,\lambda)$ as a function of $x$ for several values of variable $\lambda$:
726: $\lambda^{-1}=1,4,10,20,50$ and 100. We display both LO and NLO predictions  of the GF mechanism as
727: well as the analogous results of the combined GF+QS contribution. The azimuthal asymmetry due to
728: the mere LO GF component is given by solid line. The NLO GF predictions are plotted by dashed line.
729: The LO and NLO results of the total GF+QS contribution are given by dash-dotted and long-dashed
730: lines, respectively. In our calculations, the CTEQ5M \cite{CTEQ5} parametrization of the gluon
731: distribution function is used and a $1\%$ probability for IC in the nucleon is assumed. Throughout
732: this paper, the value $\mu_{F}=\mu_{R}=\sqrt{m^{2}+Q^{2}}$ for both factorization and
733: renormalization scales is chosen. In accordance with the CTEQ5M parametrization, we use $m_{c}=1.3
734: $ GeV and $\Lambda_{4}=326$ MeV \cite{CTEQ5}.
735: 
736: One can see from Fig.~\ref{Fg.6} the following basic features of the azimuthal asymmetry,
737: $A(x,\lambda)$, within the FFNS. First, as expected, the nonperturbative IC contribution is
738: practically invisible at low $x$, but affects essentially the GF predictions at large $x$. Since,
739: contrary to the GF mechanism, the QS component is practically $\cos2\varphi$-independent, the
740: dominance of the IC contribution at large $x$ leads to a more rapid (in comparison with the GF
741: predictions) decreasing of $A(x,\lambda)$ with growth of $x$.
742: 
743: The most remarkable property of the azimuthal asymmetry is its perturbative stability. In
744: Refs.~\cite{we2,we4}, the NLO soft-gluon corrections to the GF predictions for the $\cos2\varphi$
745: asymmetry in heavy quark photo- and leptoproduction was calculated. It was shown that, contrary to
746: the production cross sections, the quantity $A(x,\lambda)$ is practically insensitive to soft
747: radiation. One can see from Fig.~\ref{Fg.6} in the present paper that the NLO corrections to the LO
748: GF predictions for $A(x,\lambda)$ are about few percent at not large $x$.  This implies that large
749: soft-gluon corrections to $\sigma_{A,GF}^{(LO)}$ and $\sigma_{2,GF}^{(LO)}$ (increasing both cross
750: sections by a factor of two) cancel each other in the ratio
751: $\bigl(\sigma_{A,GF}^{(NLO)}\!\bigl/\sigma_{2,GF}^{(NLO)}\bigr)(x,\lambda)$ with a good accuracy.
752: In terms of so-called $K$-factors,
753: $K_{k}(x,\lambda)=\bigl(\sigma_{k}^{(NLO)}\!\bigl/\sigma_{k}^{(LO)}\bigr)(x,\lambda)$ for
754: $k=2,L,A,I$, perturbative stability of the GF predictions for $A(x,\lambda)$ is provided by the
755: fact that the corresponding $K$-factors are approximately the same at not large $x$:
756: $K_{A,GF}(x,\lambda)\approx K_{2,GF}(x,\lambda)$.
757: 
758: Comparing with each other the dash-dotted and long-dashed curves in Fig.~\ref{Fg.6}, we see that
759: the NLO corrections to the combined GF+QS result for $A(x,\lambda)$ are also small. In this case,
760: three reasons are responsible for the closeness of the LO and NLO predictions. At small $x$, where
761: the nonperturbative IC contribution is negligible, perturbative stability of the asymmetry is
762: provided by the GF component. In the large-$x$ region, where the IC mechanism dominates, the
763: azimuthal asymmetry rapidly vanishes with growth of $x$ at both LO and NLO because the QS component
764: is practically $\cos2\varphi$-independent, $\hat{\sigma}^{(1)}_{A,c}(x,\lambda)\approx\hat{\sigma
765: }^{(0)}_{A,c}(x,\lambda)=0$ \footnote{Although the ratio $(A^{(NLO)}/A^{(LO)})(x,\lambda)$ is
766: sizeable at sufficiently large $x$, the absolute values of the quantities $A^{(LO)}(x,\lambda)$ and
767: $A^{(NLO)}(x,\lambda)$ become so small that it seems reasonable to consider the asymmetry as
768: equally negligible at both LO and NLO and treat the predictions as perturbatively stable.}. At
769: intermediate values of $x$, where both mechanisms are essential, perturbative stability of
770: $A(x,\lambda)$ is due to the similarity of the corresponding $K$-factors: $K_{2,GF}(x,\lambda)\sim
771: K_{2,QS}(x,\lambda)$ \footnote{Note however that this similarity takes only place at intermediate
772: values of $x$ where both GF and QS components are essential. In the low- and large-$x$ regions, the
773: factors $K_{2,GF}(x,\lambda)$ and $K_{2,QS}(x,\lambda)$ are strongly different.}.
774: 
775: Another remarkable property of the azimuthal asymmetry closely related to fast perturbative
776: convergence is its parametric stability \footnote{Of course, parametric stability of the fixed
777: order results does not imply a fast convergence of the corresponding series. However, a fast
778: convergent series must be parametrically stable. In particular, it must be $\mu _{R}$- and $\mu
779: _{F}$-independent.}. The analysis of Refs.~\cite{we1,we4} shows that the GF predictions for the
780: $\cos 2\varphi $ asymmetry are less sensitive to standard uncertainties in the QCD input parameters
781: ($m,\mu _{R},\mu _{F},\Lambda_{QCD}$ and PDFs) than the corresponding ones for the production cross
782: sections. We have verified that the same situation takes also place for the combined GF+QS results.
783: 
784: Let us discuss how the GF predictions for the azimuthal asymmetry are affected by nonperturbative
785: contributions due to the intrinsic transverse motion of the gluon in the target. Because of the
786: relatively low $c$-quark mass, these contributions are especially important in the description of
787: the cross sections for charmed particle production.
788: 
789: To introduce $k_{T}$ degrees of freedom, $\vec{k}_{g}\simeq \zeta\vec{p}+\vec{k}_{T}$, one extends
790: the integral over the parton distribution function in Eq. (\ref{41}) to $k_{T}$-space,
791: \begin{equation}  \label{49}
792: \text{d}\zeta\,g(\zeta,\mu _{F})\rightarrow \text{d}\zeta\,\text{d}^{2}k_{T}f\big(\vec{k}_{T}\big)
793: g(\zeta,\mu _{F}).
794: \end{equation}
795: The transverse momentum distribution, $f\big( \vec{k}_{T}\big) $, is usually taken to be a
796: Gaussian:
797: \begin{equation}  \label{50}
798: f\big( \vec{k}_{T}\big) =\frac{{\rm {e}}^{-\vec{k}_{T}^{2}/\langle k_{T}^{2}\rangle }}{\pi \langle
799: k_{T}^{2}\rangle }.
800: \end{equation}
801: In practice, an analytic treatment of $k_{T}$ effects is usually used. According to Ref.~\cite{kT},
802: the $k_{T}$-smeared differential cross section of the process (\ref{1}) is a two-dimensional
803: convolution:
804: \begin{equation}  \label{51}
805: \frac{\text{d}^{4}\sigma _{lN}^{{\rm {kick}}}}{\text{d}x\text{d}Q^{2}\text{d} p_{QT}\text{d}\varphi
806: }\left( \vec{p}_{QT}\right) =\int \text{d}^{2}k_{T} \frac{{\rm {e}}^{-\vec{k}_{T}^{2}/\langle
807: k_{T}^{2}\rangle }}{\pi \langle k_{T}^{2}\rangle }\frac{\text{d}^{4}\sigma
808: _{lN}}{\text{d}x\text{d}Q^{2} \text{d}p_{QT}\text{d}\varphi }\Big( \vec{p}_{QT}-\frac{1}{2}\vec{k}
809: _{T}\Big) .
810: \end{equation}
811: The factor $\frac{1}{2}$ in front of $\vec{k}_{T}$ in the r.h.s. of Eq.~(\ref {51}) reflects the
812: fact that the heavy quark carries away about one half of the initial energy in the reaction
813: (\ref{1}).
814: 
815: Values of the $k_{T}$-kick corrections to the LO GF predictions for the $\cos 2\varphi $ asymmetry
816: in the charm production are shown in Fig.~\ref{Fg.6} by dotted curves. Calculating the $k_{T}$-kick
817: effect we use $\langle k_{T}^{2}\rangle =0.5$ GeV$^{\text{2}}$. One can see that $k_{T}$-smearing
818: for $A(x,Q^{2})$ is about $20$-$25\%$ in the region of low $Q^{2}\lesssim m^{2}$ and rapidly
819: decreases at high $Q^{2}$.
820: 
821: In Fig.~\ref{Fg.6a}, the dependence of the asymmetry $A(x,\lambda)$ on the nonperturbative
822: intrinsic charm content of the proton is presented. We plot the LO predictions for $A(x,\lambda)$
823: as a function of $x$ for several values of the variable $\lambda$ and quantity
824: $P_{c}=\int^{1}_{0}c(z)\text{d}z$ describing a probability for IC in the nucleon. Dash-dotted
825: curves describe the $\text{GF}^{\text{(LO)}}$+$\text{QS}^{\text{(LO)}}$ contributions with
826: $P_{c}=5\%,~1\%,~0.1\%$ and $0.01\%$. Solid lines correspond to the case when $P_{c}=0$. Comparing
827: with each other Figs.~\ref{Fg.6} and \ref{Fg.6a}, one can see that even a $0.1\%$ contribution of
828: the nonperturbative IC to the proton wave function could be extracted from the $\cos 2\varphi $
829: asymmetry at large enough Bjorken $x$.
830: \begin{figure}
831: \begin{center}
832: \begin{tabular}{cc}
833: \mbox{\epsfig{file=A2F1B.eps,width=252pt}}
834: & \mbox{\epsfig{file=A2F4B.eps,width=245pt}}\\
835: \mbox{\epsfig{file=A2F10B.eps,width=243pt}}
836: & \mbox{\epsfig{file=A2F20B.eps,width=245pt}}\\
837: \mbox{\epsfig{file=A2F50B.eps,width=250pt}}
838: & \mbox{\epsfig{file=A2F100B.eps,width=247pt}}\\
839: \end{tabular}
840: \caption{\label{Fg.6a}\small The LO predictions for $A(x,\lambda)$ in the FFNS at several values of
841: $\lambda$ and $P_{c}=\int^{1}_{0}c(z)\text{d}z$. Dash-dotted curves describe the
842: $\text{GF}^{\text{(LO)}}$+$\text{QS}^{\text{(LO)}}$ contributions with $P_{c}=5\%,~1\%,~0.1\%$ and
843: $0.01\%$. Solid lines correspond to the case when $P_{c}=0$.}
844: \end{center}
845: \end{figure}
846: 
847: 
848: \subsection{\label{hb} Variable Flavor Number Scheme and Perturbative Intrinsic Charm}
849: 
850: One can see from Eqs.~(\ref{39}) that the GF cross section $\hat{\sigma}_{2,g}^{(0)}(z,\lambda)$
851: contains potentially large logarithm, $\ln (Q^{2}/m^{2})$. The same situation takes also place for
852: the QS cross section $\hat{\sigma}_{2,Q}^{(1)}(z,\lambda)$ given by Eq.~(\ref{25}). At high
853: energies, $Q^{2}\rightarrow \infty$, the terms of the form $\alpha_{s}\ln (Q^{2}/m^{2})$ dominate
854: the production cross sections. To improve the convergence of the perturbative series at high
855: energies, the so-called variable flavor number schemes (VFNS) have been proposed. Originally, this
856: approach was formulated by Aivazis, Collins, Olness and Tung (ACOT) \cite{AOT,ACOT}.
857: 
858: In the VFNS, the mass logarithms of the type $\alpha_s^n\ln^n (Q^{2}/m^{2})$ are resummed via the
859: renormalization group equations. In practice, the resummation procedure consists of two steps.
860: First, the mass logarithms have to be subtracted from the fixed order predictions for the partonic
861: cross sections in such a way that in the asymptotic limit $Q^{2}\rightarrow \infty$ the well known
862: massless $\overline{\text{MS}}$ coefficient functions are recovered. Instead, a charm parton
863: density in the hadron, $c(x,Q^{2})$, has to be introduced. This density obeys the usual massless
864: NLO DGLAP evolution equation with the boundary condition $c(x,Q^{2}=Q_{0}^2)=0$ where $Q_{0}^2\sim
865: m^{2}$. So, we may say that, within the VFNS, the charm density arises perturbatively from the
866: $g\rightarrow c\bar{c}$ evolution.
867: 
868: In the VFNS, the treatment of the charm depends on the values chosen for $Q^{2}$. At low
869: $Q^{2}<Q_{0}^2$, the production cross sections are described by the light parton contributions
870: ($u,d,s$ and $g$). The charm production is dominated by the GF process and its higher order QCD
871: corrections. At high $Q^{2}\gg m^{2}$, the charm is treated in the same way as the other light
872: quarks and it is represented by a charm parton density in the hadron, which evolves in $Q^{2}$. In
873: the intermediate scale region, $Q^{2}\sim m^{2}$, one has to make a smooth connection between the
874: two different prescriptions.
875: 
876: Strictly speaking, the perturbative charm density is well defined at high $Q^2\gg m^2$ but does not
877: have a clean interpretation at low $Q^2$. Since the perturbative IC originates from resummation of
878: the mass logarithms of the type $\alpha_s^n\ln^n (Q^{2}/m^{2})$, it is usually assumed that the
879: corresponding PDF vanishes with these logarithms, i.e. for $Q^{2}<Q_{0}^2\approx m^{2}$. On the
880: other hand, the threshold constraint $W^2=(q+p)^2=Q^2(1/x-1)>4m^2$ implies that $Q_0$ is not a
881: constant but "live" function of $x$. To avoid this problem, several solutions have been proposed
882: (see e.g. Refs.~\cite{chi,SACOT}). In this paper, we use the so-called ACOT($\chi$) prescription
883: \cite{chi} which guarantees (at least at $Q^2>m^2$) the correct threshold behavior of the
884: heavy-quark-initiated contributions.
885: 
886: Within the VFNS, the $\varphi$-independent charm production cross sections have three pieces:
887: \begin{equation}\label{52}
888: \sigma_{2}(x,\lambda)=\sigma_{2,GF}(x,\lambda)-\sigma_{2,SUB}(x,\lambda)+\sigma_{2,QS}(x,\lambda),
889: \end{equation}
890: where the first and third terms on the right hand side describe the usual (unsubtracted) GF and QS
891: contributions while the second (subtraction) term renders the total result infra-red safe in the
892: limit $m^{2}\rightarrow 0$. The only constraint imposed on the subtraction term is to reproduce at
893: high energies the familiar $\overline{\text{MS}}$ partonic cross section:
894: \begin{equation}\label{53}
895: \lim_{\lambda\rightarrow 0}\left[\hat{\sigma}_{2,g}(z,\lambda)-
896: \hat{\sigma}_{2,SUB}(z,\lambda)\right]=\hat{\sigma}^{\overline{\text{MS}}}_{2,g}(z).
897: \end{equation}
898: Evidently, there is some freedom in the choice of finite mass terms of the form $\lambda^{n}$ (with
899: a positive $n$) in $\hat{\sigma}_{2,SUB}(z,\lambda)$. For this reason, several prescriptions have
900: been proposed to fix the subtraction term. As mentioned above, we use the so-called ACOT($\chi$)
901: scheme \cite{chi}.
902: 
903: According to the ACOT($\chi$) prescription, the lowest order $\varphi$-independent cross section is
904: \begin{equation}\label{54}
905: \sigma^{(LO)}_{2}(x,\lambda)=\int\limits_{\chi}^{1}\text{d}z\,g(z,\mu_{F})\left[\hat{\sigma}_{2,g}^{(0)}
906: \!\left(x/z,\lambda\right)-\frac{\alpha_{s}}{\pi}\ln\frac{\mu_{F}^{2}}{m^{2}}
907: \;\hat{\sigma}_{B}\left(x/z\right)P^{(0)}_{g\rightarrow
908: c}\left(\chi/z\right)\right]+\hat{\sigma}_{B}(x)c_{+}(\chi,\mu_{F}),
909: \end{equation}
910: where $P^{(0)}_{g\rightarrow c}$ is the LO gluon-quark splitting function, $P^{(0)}_{g\rightarrow
911: c}(\zeta)=\left.\left[(1-\zeta)^{2}+\zeta^{2}\right]\right/2$, and the LO GF cross section
912: $\hat{\sigma}_{2,g}^{(0)}$ is given by Eqs.~(\ref{39}). Remember also that $\chi=x(1+4\lambda)$ and
913: $c_{+}(\zeta,\mu_{F})=c(\zeta,\mu_{F})+\bar{c}(\zeta,\mu_{F})$.
914: 
915: The asymptotic behavior of the subtraction terms is fixed by the parton level factorization
916: theorem. This theorem implies that the partonic cross sections d$\hat{\sigma}$ can be factorized
917: into process-dependent infra-red safe hard scattering cross sections d$\tilde{\sigma}$, which are
918: finite in the limit $m\rightarrow 0$, and universal (process-independent) partonic PDFs
919: $f_{a\rightarrow i}$ and fragmentation functions $d_{n\rightarrow Q}$:
920: \begin{equation}\label{add1}
921: \text{d}\hat{\sigma}(\gamma^{*}+a\rightarrow Q+X)=\sum_{i,n}f_{a\rightarrow
922: i}(\zeta)\otimes\text{d}\tilde{\sigma}(\gamma^{*}+i\rightarrow n+X)\otimes d_{n\rightarrow Q}(z).
923: \end{equation}
924: In Eq.~(\ref{add1}), the symbol $\otimes$ denotes the usual convolution integral, the indices
925: $a,i,n$ and $Q$ denote partons, $p_{i}=\zeta p_{a}$ and $p_{Q}=z p_{n}$. All the logarithms of the
926: heavy-quark mass (i.e., the singularities in the limit $m\rightarrow 0$) are contained in the PDFs
927: $f_{a\rightarrow i}$ and fragmentation functions $d_{n\rightarrow Q}$ while d$\tilde{\sigma}$ are
928: IR-safe (i.e., are free of the $\ln m^{2}$ terms). The expansion of Eq.~(\ref{add1}) can be used to
929: determine order by order the subtraction terms. In particular, for the LO GF contribution to the
930: charm leptoproduction one finds \cite{ACOT}
931: \begin{equation}\label{add2}
932: \hat{\sigma}^{(0)}_{k,SUB}\left(z,\ln\,(\mu_{F}^{2}/m^{2})\right)=f^{(1)}_{g\rightarrow
933: c}\left(\zeta,\ln\,(\mu_{F}^{2}/m^{2})\right)\otimes\hat{\sigma}^{(0)}_{k,QS}(z/\zeta), \qquad
934: \qquad  (k=2,L,A,I),
935: \end{equation}
936: where $f^{(1)}_{g\rightarrow c}\left(\zeta,\ln\,(\mu_{F}^{2}/m^{2})\right)=\left(
937: \alpha_{s}/2\pi\right)\ln\,(\mu_{F}^{2}/m^{2})\,P^{(0)}_{g\rightarrow c}\left(\zeta\right)$
938: describes the charm distribution in the gluon within the $\overline{\text{MS}}$ factorization
939: scheme.
940: 
941: One can see from Eq.~(\ref{add2}) that the azimuth-dependent GF cross sections
942: $\hat{\sigma}_{A,GF}$ and $\hat{\sigma}_{I,GF}$ don't have subtraction terms at LO because the
943: lowest order QS contribution is $\varphi$-independent. For this reason, the
944: $\cos2\varphi$-dependence within the VFNS has the same form as in the FFNS:
945: \begin{equation}\label{55}
946: \sigma^{(LO)}_{A}(x,\lambda)=\int\limits_{\chi}^{1}\text{d}z\,g(z,\mu_{F})\,\hat{\sigma}_{A,g}^{(0)}
947: \!\left(x/z,\lambda\right).
948: \end{equation}
949: \begin{figure}
950: \begin{center}
951: \begin{tabular}{cc}
952: \mbox{\epsfig{file=AgcV1.eps,width=250pt}}
953: & \mbox{\epsfig{file=AgcV4.eps,width=244pt}}\\
954: \mbox{\epsfig{file=AgcV10.eps,width=242pt}}
955: & \mbox{\epsfig{file=AgcV20.eps,width=244pt}}\\
956: \mbox{\epsfig{file=AgcV50.eps,width=244pt}}
957: & \mbox{\epsfig{file=AgcV100.eps,width=253pt}}\\
958: \end{tabular}
959: \caption{\label{Fg.7}\small Azimuthal asymmetry parameter $A(x,\lambda)$ in the VFNS at several
960: values of $\lambda$. The following contributions are plotted: $\text{GF}^{\text{(LO)}}$ (solid
961: curves), $\text{GF}^{\text{(LO)}}$$-\text{SUB}^{\text{(LO)}}$+$\text{QS}^{\text{(LO)}}$ with the
962: CTEQ5M set of PDFs (dotted curves) and $\text{GF}^{\text{(LO)}}$$-\text{SUB}^{\text{(LO)}}$+
963: $\text{QS}^{\text{(LO)}}$ with the CTEQ4M set of PDFs (dashed curves).}
964: \end{center}
965: \end{figure}
966: 
967: Fig.~\ref{Fg.7} shows the ACOT($\chi$) predictions for the asymmetry parameter $A(x,\lambda)$ at
968: several values of variable $\lambda$: $\lambda^{-1}=1,4,10,20,50$ and 100. For comparison, we plot
969: also the LO GF predictions (solid curves). In the ACOT($\chi$) case, we consider the CTEQ5M (dotted
970: lines) and CTEQ4M (dashed curves) parametrizations of the gluon and charm densities in the proton.
971: Corresponding values of the charm quark mass are $m_{c}=1.3 $ GeV \cite{CTEQ5} (for the CTEQ5M
972: PDFs) and $m_{c}=1.6 $ GeV \cite{CTEQ4} (for the CTEQ4M PDFs). The default value of the
973: factorization scale is $\mu_{F}=\sqrt{m^{2}+Q^{2}}$.
974: 
975: One can see from Fig.~\ref{Fg.7} the following properties of the azimuthal asymmetry,
976: $A(x,\lambda)$, within the VFNS. Contrary to the nonperturbative IC component, the perturbative one
977: is significant practically at all values of Bjorken $x$ and $Q^{2}>m^{2}$. The perturbative charm
978: contribution leads to a sizeable decreasing of the GF predictions for the $\cos2\varphi$-asymmetry.
979: In the ACOT($\chi$) scheme, the IC contribution reduces the GF results for $A(x,\lambda)$ by about
980: $30\%$. The origin of this reduction is straightforward: the QS component is practically
981: $\cos2\varphi$-independent.
982: 
983: The ACOT($\chi$) predictions for the asymmetry depend weakly on the parton distribution functions
984: we use. It is seen from Fig.~\ref{Fg.7} that the CTEQ5M and CTEQ4M sets of PDFs lead to very
985: similar results  for $A(x,\lambda)$. Note that we give the CTEQ5M predictions at low $x$ only
986: because of irregularities in the CTEQ5M charm density at large $x$.
987: 
988: We have also analyzed how the VFNS predictions depend on the choice of subtraction prescription. In
989: particular, the schemes proposed in Refs.~\cite{KS,SACOT} have been considered. We find that,
990: sufficiently above the production threshold, these subtraction prescriptions reduce the GF results
991: for the asymmetry by approximately $30\div 50 \%$.
992: 
993: One can conclude that impact of the perturbative IC on the $\cos2\varphi$ asymmetry is essential in
994: the whole region of Bjorken $x$ and therefore can be tested experimentally.
995: 
996: \section{Conclusion}
997: In the present paper, we consider the azimuthal dependence in charm leptoproduction as a probe of
998: the IC content of the proton. Our analysis is based on the fact that the GF and QS components have
999: strongly different $\cos2\varphi$-distributions. This fact follows from the NLO calculations of
1000: both parton level contributions. In the framework of the FFNS, we justify the most remarkable
1001: property of the hadron level azimuthal $\cos2\varphi$ asymmetry: the combined GF+QS predictions for
1002: $A(x,Q^{2})$ are perturbatively and parametrically stable. The nonperturbative IC contribution
1003: (resulting from the five-quark $\left\vert uudc\bar{c}\right\rangle$ component  of the proton wave
1004: function) is practically invisible at low $x$, but affects essentially the GF predictions for the
1005: asymmetry at large Bjorken $x$. We conclude that measurements of the $\cos2\varphi$ asymmetry at
1006: large $x$ could directly probe the nonperturbative intrinsic charm.
1007: 
1008: Within the VFNS, charm density originates perturbatively from the $g\rightarrow c\bar{c}$ process
1009: and obeys the DGLAP evolution equation. Presently, charm densities are included practically in all
1010: the global sets of PDFs like CTEQ and MRST. Our analysis shows that these charm distribution
1011: functions reduce dramatically (by about 1/3) the GF predictions for $A(x,Q^{2})$ practically at all
1012: values of $x$. For this reason, the perturbative IC contribution can easily be measured using the
1013: azimuthal $\cos2\varphi$-distributions in charm leptoproduction.
1014: 
1015: The VFN schemes have been proposed to resum the mass logarithms of the form $\alpha_{s}^{n}\ln^{n}
1016: (Q^{2}/m^{2})$ which dominate the production cross sections at high energies. Evidently, were the
1017: calculation done to all orders of $\alpha_{s}$, the VFNS and FFNS (without nonperturbative IC)
1018: would be exactly equivalent. There is a point of view advocated in Refs.~\cite{ACOT,collins} that,
1019: at high energies, the perturbative series converge better within the VFNS than in the FFNS. There
1020: is also another opinion \cite{BMSN,neerven} that the above logarithms do not vitiate the
1021: convergence of the perturbation expansion so that a resummation is, in principle, not necessary.
1022: Our analysis of the azimuthal dependence in leptoproduction indicates an experimental way to
1023: resolve this problem. First, contrary to the production cross sections, the azimuthal
1024: $\cos2\varphi$-asymmetry is well defined numerically in pQCD. Second, sufficiently above the
1025: production threshold (i.e., at small enough Bjorken $x$), the LO VFNS predictions for the
1026: $\cos2\varphi$-asymmetry differ by more than $30\%$ from the corresponding FFNS ones. Third,
1027: nonperturbative contributions (like the intrinsic gluon motion in the target) can't compensate for
1028: this difference at non-small $Q^{2}$ where the VFNS is expected to be adequate. Therefore
1029: measurements of the azimuthal distributions in charm leptoproduction would make it possible to
1030: clarify the question whether the VFNS perturbative series for $A(x,Q^{2})$ converges better than
1031: the FFNS one.
1032: \begin{acknowledgments}
1033: We thank S.J. Brodsky for stimulating discussions and useful suggestions. We also would like to
1034: acknowledge interesting correspondence with I. Schienbein. This work was supported in part by
1035: the ANSEF grant 04-PS-hepth-813-98 and NFSAT grant GRSP-16/06.
1036: \end{acknowledgments}
1037: 
1038: \appendix
1039: \section{\label{virt}Virtual and Soft Contributions to the Quark Scattering}
1040: In this Appendix we reproduce some results of Hoffmann and Moore for the $\varphi$-independent QS
1041: cross sections, and correct two misprints uncovered in Ref.~\cite{HM}. We work in four dimensions,
1042: in the Feynman gauge and use the on-mass-shell renormalization scheme. We compute the absorptive
1043: part of the Feynman diagram (which is free of the UV divergences) and then restore the real part
1044: using the appropriate dispersion relations.
1045: 
1046: In the on-mass-shell scheme, the renormalized fermion self-energy vanishes like
1047: $(\hat{p}_{Q}-m)^{2}$ which means that the second and third diagrams in Fig.~\ref{Fg.2}c do not
1048: contribute to the cross section when the external quark legs are on-shell, $\hat{p}_{Q}\rightarrow
1049: m$. The first graph in Fig.~\ref{Fg.2}c describes the NLO corrections to the quark-photon vertex
1050: function:
1051: \begin{equation}  \label{71}
1052: \Lambda _{\mu }(q) =f\left( Q^{2}\right) \gamma _{\mu }- \frac{g\left( Q^{2}\right) }{2m}\sigma
1053: _{\mu \nu }q^{\nu },
1054: \end{equation}
1055: where $\sigma _{\mu \nu }=\frac{1}{2}\left( \gamma _{\mu }\gamma _{\nu }-\gamma _{\nu }\gamma _{\mu
1056: }\right) $ while $f\left( Q^{2}\right) $ and $g\left( Q^{2}\right)$ are the quark electromagnetic
1057: formfactors. At the lowest order $\Lambda _{\mu }^{(0)}=\gamma _{\mu }$.
1058: 
1059: The virtual lepton-quark cross section, $\hat{\sigma}_{lQ}^{V}$, is obtained from the interference
1060: term between the virtual and the Born amplitude. The result can be written in terms of the
1061: electromagnetic formfactors as:
1062: \begin{equation} \label{72}
1063: \frac{\text{d}^{\text{2}}\hat{\sigma}_{lQ}^{V}}{\text{d}z\text{d}Q^{2}}= \frac{\alpha_{em}}{\pi
1064: }\frac{\hat{\sigma}_{B}(z)}{zQ^{2}}\delta (1-z)\left\{ \left[ 1+(1-y)^{2}-2\lambda
1065: z^{2}y^{2}\right] f\left( Q^{2}\right) +y^{2}g\left( Q^{2}\right) \right\}.
1066: \end{equation}
1067: Taking into account the definition of the HM cross sections, $\sigma^{(2)}$ and $\sigma^{(L)}$,
1068: given by Eqs.~(\ref{35}), (\ref{36}) and (\ref{18}), we find that corresponding virtual parts are:
1069: \begin{equation}\label{73}
1070: \sigma_{1V}^{(2)}(z,Q^{2})=2\delta (1-z)f^{(1)}\left( Q^{2}\right), \qquad \qquad \qquad \sigma
1071: _{1V}^{(L)}(z,Q^{2})=-\delta (1-z)g^{(1)}\left( Q^{2}\right),
1072: \end{equation}
1073: where $f^{(1)}\left( Q^{2}\right)$ and $g^{(1)}\left( Q^{2}\right)$ are the NLO corrections to the
1074: electromagnetic formfactors. For the NLO HM  cross sections, $\sigma_{1V}^{(2)}$ and
1075: $\sigma_{1V}^{(L)}$, we use exactly the same notations as in Ref.~\cite{HM}.
1076: 
1077: In the on-mass-shell renormalization scheme, the renormalized vertex correction vanishes as the
1078: photon virtuality goes to zero, $f^{(1)}\left(0\right)=0$. This is a consequence of the Ward
1079: identity and the fact that the real photon field (like the massive fermion one) is unrenormalized
1080: in first order QCD. To satisfy the condition $f^{(1)}\left(0\right)=0$ automatically, we should use
1081: for $f^{(1)}\left(q^{2}\right)$ the dispersion relation with one subtraction. The second
1082: formfactor, $g^{(1)}\left(q^{2}\right)$, has no singularities. For this reason, we use for
1083: $g^{(1)}\left(q^{2}\right)$ the dispersion relation without subtractions:
1084: \begin{equation}\label{74}
1085: f^{(1)}\left( q^{2}\right) =\frac{q^{2}}{\pi }\int\limits_{4m^{2}}^{\infty }
1086: \frac{\text{d}t\;\,\text{Im }f^{(1)}(t)}{t(t-q^{2}-i0)},\qquad \qquad \qquad g^{(1)}\left(
1087: q^{2}\right) =\frac{1}{\pi }\int\limits_{4m^{2}}^{\infty } \frac{\text{d}t\;\,\text{Im
1088: }g^{(1)}(t)}{t-q^{2}-i0}.
1089: \end{equation}
1090: Calculating the imaginary parts of the formafactors and restoring their real parts with the help of
1091: Eqs.~(\ref{74}) yields
1092: \begin{eqnarray}
1093: f^{(1)}\left( Q^{2}\right)&=&\frac{\alpha _{s}}{\pi }C_{F}\biggl\{ \left[ 1+ \frac{1+2\lambda
1094: }{\sqrt{1+4\lambda }}\ln r\right]\left(\ln \frac{m}{m_{g}}-1\right) \nonumber\\%
1095: &&+\frac{1+2\lambda}{\sqrt{1+4\lambda }}\left[ \text{Li}_{2}(-r)+ \frac{\pi ^{2}}{12}+
1096: \frac{1}{4}\ln^{2}r+\frac{1}{2}\ln r\ln\frac{1+4\lambda }{\lambda }\right]+\frac{1}{4}
1097: \frac{\ln r}{\sqrt{1+4\lambda }}\biggr\},\label{75}\\
1098: g^{(1)}\left( Q^{2}\right)&=&-\frac{\alpha_{s}}{\pi}C_{F}\frac{\lambda\ln
1099: r}{\sqrt{1+4\lambda}}.\label{76}
1100: \end{eqnarray}
1101: In Eqs.~(\ref{75}) and (\ref{76}), $C_{F}=(N_{c}^{2}-1)/(2N_{c})$, where $N_{c}$ is number of
1102: colors, and $r$ is defined by Eq.~(\ref{29}). Taking into account that $\Lambda
1103: ^{(1)\mu}=\left(\alpha_{s}\left/\alpha_{em}\right.\right)C_{F}\,\Lambda_{QED}^{(1)\mu}$, we see
1104: that Eqs.~(\ref{75}, \ref{76}) reproduce the textbook QED results.
1105: 
1106: It is now straightforward to obtain the virtual contribution to the longitudinal cross section.
1107: Combining Eqs.~(\ref{73}) and (\ref{76}) yields:
1108: \begin{equation}\label{77}
1109: \sigma _{1V}^{(L)}(z,Q^{2})=\frac{\alpha _{s}}{\pi }C_{F}\delta (1-z)\frac{\lambda \ln
1110: r}{\sqrt{1+4\lambda }}.
1111: \end{equation}
1112: Comparing the above result with the corresponding one given by Eq.~(39) in Ref.~\cite{HM}, we see
1113: that the HM expression for $\sigma _{1V}^{(L)}$ has opposite sign. Note also that this typo
1114: propagates into the final result for $\sigma _{1}^{(L)}$ given by Eq.~(52) \cite{HM}.
1115: 
1116: Calculation of the bremsstrahlung contribution to the longitudinal cross section, $\sigma
1117: _{1B}^{(L)}(z,Q^{2})$, is also straightforward. We coincide with the HM result for $\sigma
1118: _{1B}^{(L)}(z,Q^{2})$ given by Eq.~(49) in Ref.~\cite{HM}. However there is one more misprint in
1119: the HM expression for $\sigma _{1}^{(L)}$: the r.h.s of  Eq.~(52) \cite{HM} should be multiplied by
1120: $z$.
1121: 
1122: In the case of $\sigma _{1}^{(2)}(z,Q^{2})$, the situation is slightly more complicated due to the
1123: need to take into account the IR singularities. One can see from Eq.~(\ref{75}) that $f^{(1)}\left(
1124: Q^{2}\right)$ has an IR divergence which is regularized with the help of an infinitesimal gluon
1125: mass $m_{g}$. This singularity is cancelled when one adds the so-called soft contribution
1126: originating from the real gluon emission. For this purpose we introduce another infinitesimal
1127: parameter $\delta z$, $\left(m_{g}\!\left/m\right.\right)\ll\delta z\ll 1$. The full bremsstrahlung
1128: contribution, $\sigma _{1B}^{(2)}$, can then be splitted into the soft and hard pieces as follows:
1129: \begin{equation}\label{78}
1130: \sigma _{1soft}^{(2)}(z,Q^{2})=\theta (z+\delta z-1)\sigma _{1B}^{(2)}(z,Q^{2}),\qquad \qquad
1131: \qquad \qquad \sigma _{1hard}^{(2)}(z,Q^{2})=\theta (1-z-\delta z)\sigma _{1B}^{(2)}(z,Q^{2}),
1132: \end{equation}
1133: where $\theta (1-z-\delta z)$ is the Heaviside step function. The soft cross section should be
1134: calculated in the eikonal approximation, $\vec{p_{g}}\rightarrow 0$, taking into account the
1135: infinitesimal gluon mass $m_{g}$. As a result, the sum of the virtual and soft contributions is IR
1136: finite:
1137: \begin{eqnarray}
1138: \sigma _{1V}^{(2)}+\sigma _{1soft}^{(2)} &=&\frac{\alpha _{s}}{\pi } C_{F}\delta (1-z)\biggl\{-2\ln
1139: (\delta z)\left[ 1+\frac{1+2\lambda }{\sqrt{ 1+4\lambda }}\ln r\right] +2\ln \lambda
1140: -1-\frac{\sqrt{1+4\lambda }}{2}\ln r \nonumber\\
1141: &&+\frac{1+2\lambda}{\sqrt{1+4\lambda}}\left[\text{Li}_{2}(r^{2})+2\text{Li}_{2}(-r)+\frac{3}{2}
1142: \ln^{2}r-2\ln r-\ln r\ln\lambda+2\ln r\ln (1+4\lambda)\right] \biggr\}. \label{79}
1143: \end{eqnarray}
1144: Adding to the above expression the hard cross section $\sigma _{1hard}^{(2)}$ defined by
1145: Eq.~(\ref{78}), we reproduce  in the limit $\delta z\rightarrow 0$ the full result for $\sigma
1146: _{1}^{(2)}$ given by Eq.~(51) in Ref.~\cite{HM}.
1147: 
1148: \section{\label{soft}NLO Soft-Gluon Corrections to the Photon-Gluon Fusion}
1149: This Appendix provides an overview of the NLO soft-gluon approximation for the photon-gluon fusion
1150: mechanism. We present the final results for the parton level cross sections to the next-to-leading
1151: logarithmic (NLL) accuracy. More details can be found in Refs.~\cite{Laenen-Moch,we2,we4}.
1152: 
1153: To take into account the NLO contributions to the GF mechanism, one needs to calculate the virtual
1154: ${\cal O}(\alpha _{em}\alpha _{s}^{2})$ corrections to the Born process (\ref{38}) and the real
1155: gluon emission:
1156: \begin{equation}  \label{60}
1157: \gamma ^{*}(q)+g(k_{g})\rightarrow Q(p_{Q})+\overline{Q}(p_{\stackrel{\_}{Q}})+g(p_{g}).
1158: \end{equation}
1159: The partonic invariants describing the single-particle inclusive (1PI) kinematics are
1160: \begin{eqnarray}
1161: s^{\prime }=2q\cdot k_{g}=s+Q^{2}=\zeta S^{\prime },\qquad \qquad &&t_{1}=\left(
1162: k_{g}-p_{Q}\right) ^{2}-m^{2}=\zeta T_{1},  \nonumber \\
1163: s_{4}=s^{\prime }+t_{1}+u_{1},\qquad \qquad &&u_{1}=\left( q-p_{Q}\right) ^{2}-m^{2}=U_{1},
1164: \label{61}
1165: \end{eqnarray}
1166: where $\zeta$ is defined by $\vec{k}_{g}= \zeta\vec{p}\,$ and $s_{4}$ measures the inelasticity of
1167: the reaction (\ref{60}). The corresponding 1PI hadron level variables describing the reaction
1168: (\ref{1}) are
1169: \begin{eqnarray}
1170: S^{\prime }=2q\cdot p=S+Q^{2},\qquad \qquad &&T_{1}=\left( p-p_{Q}\right)
1171: ^{2}-m^{2},  \nonumber \\
1172: S_{4}=S^{\prime }+T_{1}+U_{1},\qquad \qquad &&U_{1}=\left( q-p_{Q}\right) ^{2}-m^{2}.  \label{62}
1173: \end{eqnarray}
1174: 
1175: The exact NLO calculations of the unpolarized heavy quark production in $\gamma g$
1176: \cite{Ellis-Nason,Smith-Neerven}, $\gamma ^{*}g$ \cite{LRSN}, and $gg$
1177: \cite{Nason-D-E-1,Nason-D-E-2,Nason-D-E-3,BKNS} collisions show that, near the partonic threshold,
1178: a strong logarithmic enhancement of the cross sections takes place in the collinear, $\vec{p}_{g,T}
1179: $ $\rightarrow 0$, and soft, $\vec{p}_{g}\rightarrow 0$, limits. This threshold (or soft-gluon)
1180: enhancement has universal nature in the perturbation theory and originates from incomplete
1181: cancellation of the soft and collinear singularities between the loop and the bremsstrahlung
1182: contributions. Large leading and next-to-leading threshold logarithms can be resummed to all orders
1183: of perturbative expansion using the appropriate evolution equations
1184: \cite{Contopanagos-L-S,Laenen-O-S,Kidonakis-O-S}. The analytic results for the resummed cross
1185: sections are ill-defined due to the Landau pole in the coupling strength $\alpha _{s}$. However, if
1186: one considers the obtained expressions as generating functionals of the perturbative theory and
1187: re-expands them at fixed order in $\alpha _{s}$, no divergences associated with the Landau pole are
1188: encountered.
1189: 
1190: Soft-gluon resummation for the photon-gluon fusion has been performed in Ref.~\cite{Laenen-Moch}
1191: and checked in Refs.~\cite{we2,we4}. To NLL accuracy, the perturbative expansion for the partonic
1192: cross sections, d$^{2}\hat{\sigma}_{k,g}/$d$t_{1}$d$u_{1}$ ($k=T,L,A,I$), can be written in a
1193: factorized form as
1194: \begin{equation}  \label{63}
1195: s^{\prime 2}\frac{\text{d}^{2}\hat{\sigma}_{k,g}}{\text{d}t_{1}\text{d}u_{1}}( s^{\prime
1196: },t_{1},u_{1}) =B_{k,g}^{\text{{\rm Born}}}( s^{\prime },t_{1},u_{1})\left\{\delta (s^{\prime
1197: }+t_{1}+u_{1}) +\sum_{n=1}^{\infty }\left( \frac{\alpha _{s}C_{A}}{\pi}\right)^{n}K^{(n)}(
1198: s^{\prime },t_{1},u_{1})\right\} ,
1199: \end{equation}
1200: with the Born level distributions $B_{k,g}^{\text{{\rm Born}}}$ given by
1201: \begin{eqnarray}
1202: B_{T,g}^{\text{{\rm Born}}}( s^{\prime },t_{1},u_{1}) &=&\pi e_{Q}^{2}\alpha _{em}\alpha _{s}\left[
1203: \frac{t_{1}}{u_{1}}+\frac{u_{1}}{t_{1} }+4\left( \frac{s}{s^{\prime }}-\frac{m^{2}s^{\prime
1204: }}{t_{1}u_{1}}\right) \left( \frac{s^{\prime }(m^{2}-Q^{2}/2)}{t_{1}u_{1}}+\frac{Q^{2}}{s^{\prime}
1205: }\right) \right] ,  \nonumber \\
1206: B_{L,g}^{\text{{\rm Born}}}( s^{\prime },t_{1},u_{1}) &=&\pi e_{Q}^{2}\alpha _{em}\alpha _{s}\left[
1207: \frac{8Q^{2}}{s^{\prime }}\left( \frac{s}{s^{\prime }}-\frac{m^{2}s^{\prime }}{t_{1}u_{1}}\right)
1208: \right] ,
1209: \nonumber \\
1210: B_{A,g}^{\text{{\rm Born}}}( s^{\prime },t_{1},u_{1}) &=&\pi e_{Q}^{2}\alpha _{em}\alpha _{s}\left[
1211: 4\left( \frac{s}{s^{\prime }}-\frac{ m^{2}s^{\prime }}{t_{1}u_{1}}\right) \left(
1212: \frac{m^{2}s^{\prime }}{
1213: t_{1}u_{1}}+\frac{Q^{2}}{s^{\prime }}\right) \right] ,  \label{64} \\
1214: B_{I,g}^{\text{{\rm Born}}}( s^{\prime },t_{1},u_{1}) &=&\pi e_{Q}^{2}\alpha _{em}\alpha _{s}\left[
1215: 4\sqrt{Q^{2}}\left( \frac{t_{1}u_{1}s }{s^{\prime 2}}-m^{2}\right)
1216: ^{1/2}\frac{t_{1}-u_{1}}{t_{1}u_{1}}\left( 1-\frac{2Q^{2}}{s^{\prime }}-\frac{2m^{2}s^{\prime
1217: }}{t_{1}u_{1}}\right) \right] .  \nonumber
1218: \end{eqnarray}
1219: 
1220: Note that the functions $K^{(n)}( s^{\prime },t_{1},u_{1}) $ in Eq.~(\ref{63}) originate from the
1221: collinear and soft limits. Radiation of soft and collinear gluons does not affect the transverse
1222: momentum of detected particles and therefore the azimuthal angle $\varphi$. For this reason, the
1223: functions $ K^{(n)}(s^{\prime },t_{1},u_{1}) $ are the same for all helicity cross sections
1224: $\hat{\sigma}_{k,g}$ ($k=T,L,A,I$). At NLO, the soft-gluon corrections to NLL accuracy in the
1225: $\overline{\text{MS}}$ scheme are
1226: \begin{eqnarray}
1227: K^{(1)}( s^{\prime },t_{1},u_{1}) &=&2\left[ \frac{\ln \left( s_{4}/m^{2}\right) }{s_{4}}\right]
1228: _{+}-\left[ \frac{1}{s_{4}}\right] _{+}\left\{ 1+\ln \left( \frac{u_{1}}{t_{1}}\right) -\left(
1229: 1-\frac{2C_{F}}{ C_{A}}\right) \left( 1+\text{Re}L_{\beta }\right) +\ln \left( \frac{\mu ^{2}
1230: }{m^{2}}\right) \right\}   \nonumber \\
1231: &&+\delta ( s_{4}) \ln \left( \frac{-u_{1}}{m^{2}}\right) \ln \left( \frac{\mu ^{2}}{m^{2}}\right),
1232: \label{65}
1233: \end{eqnarray}
1234: where we use $\mu =\mu _{F}=\mu_{R}$. In Eq.~(\ref{65}), $C_{A}=N_{c}$ and $
1235: C_{F}=(N_{c}^{2}-1)/(2N_{c})$, where $N_{c}$ is number of colors, while $ L_{\beta
1236: }=(1-2m^{2}/s)\{\ln [(1-\beta )/(1+\beta )]+$i$\pi\}$ with $\beta=\sqrt{1-4m^{2}/s}$. The
1237: single-particle inclusive ''plus`` distributions are defined by
1238: \begin{equation}  \label{66}
1239: \left[\frac{\ln^{l}\left( s_{4}/m^{2}\right) }{s_{4}}\right]_{+}=\lim_{\epsilon \rightarrow
1240: 0}\left\{\frac{\ln^{l}\left(s_{4}/m^{2}\right) }{s_{4}}\theta ( s_{4}-\epsilon)+\frac{1}{l+1}\ln
1241: ^{l+1}\left(\frac{\epsilon }{m^{2}}\right) \delta ( s_{4})\right\}.
1242: \end{equation}
1243: For any sufficiently regular test function $h(s_{4})$, Eq.~(\ref{66}) gives
1244: \begin{equation}\label{67}
1245: \int\limits_{0}^{s_{4}^{\max }}\text{d}s_{4}\,h(s_{4})\left[ \frac{\ln ^{l}\left(
1246: s_{4}/m^{2}\right) }{s_{4}}\right] _{+}=\int\limits_{0}^{s_{4}^{\max }}\text{d}s_{4}\left[
1247: h(s_{4})-h(0)\right] \frac{\ln ^{l}\left( s_{4}/m^{2}\right) }{s_{4}}+\frac{1}{l+1}h(0)\ln
1248: ^{l+1}\left( s_{4}^{\max }/m^{2}\right) .
1249: \end{equation}
1250: 
1251: In Eq.~(\ref{65}), we have preserved the NLL terms for the scale-dependent logarithms too. Note
1252: also that the results (\ref{64}) and (\ref{65}) agree to NLL accuracy with the exact ${\cal
1253: O}(\alpha_{em}\alpha _{s}^{2})$ calculations of the photon-gluon cross sections
1254: $\hat{\sigma}_{T,g}$ and $\hat{\sigma}_{L,g}$ given in Ref.~\cite{LRSN}.
1255: 
1256: To investigate the scale dependence of the results (\ref{63}$-$\ref{65}), it is convenient to
1257: introduce for the fully inclusive (integrated over $t_{1}$ and $u_{1}$) cross sections,
1258: $\hat{\sigma}_{k,g}$ ($k=T,L,A,I$),{\large \ }the dimensionless coefficient functions
1259: $c_{k,g}^{(n,l)}$ defined by Eq.~(\ref{32}). Concerning the NLO scale-independent coefficient
1260: functions, only $ c_{T,g}^{(1,0)}$ and $c_{L,g}^{(1,0)}$ are known exactly
1261: \cite{LRSN,Harris-Smith}. As to the $\mu$-dependent coefficients, they can by calculated explicitly
1262: using the evolution equation:
1263: \begin{equation}  \label{68}
1264: \frac{\text{d}\hat{\sigma}_{k,g}(z ,Q^{2},\mu ^{2})}{\text{d}\ln \mu
1265: ^{2}}=-\int\limits_{\zeta_{\min }}^{1}\text{d}\zeta\,\hat{\sigma}_{k,g}(z/\zeta,Q^{2},\mu
1266: ^{2})P_{gg}(\zeta),
1267: \end{equation}
1268: where $z=Q^{2}/s^{\prime}$, $\zeta_{\min }=z(1+4\lambda)$, $\hat{\sigma}_{k,g}(z,Q^{2},\mu )$ are
1269: the cross sections resummed to all orders in $\alpha _{s}$ and $P_{gg}(\zeta)$ is the corresponding
1270: (resummed) Altarelli-Parisi gluon-gluon splitting function. Expanding Eq.~(\ref{68}) in $\alpha
1271: _{s}$, one can find \cite{Laenen-Moch,we2}
1272: \begin{equation}  \label{69}
1273: c_{k,g}^{(1,1)}(z, \lambda)=\frac{1}{4\pi ^{2}}\int\limits_{\zeta_{\min}}^{1}\text{d}\zeta\left[
1274: b_{2}\delta (1-\zeta)-\,P_{gg}^{(0)}(\zeta)\right] c_{k,g}^{(0,0)}(z/\zeta,\lambda),
1275: \end{equation}
1276: where $b_{2}=(11C_{A}-2n_{f})/12$ is the first coefficient of the $\beta(\alpha _{s})$-function
1277: expansion and $n_{f}$ is the number of active quark flavors. The one-loop gluon splitting function
1278: is:
1279: \begin{equation}  \label{70}
1280: P_{gg}^{(0)}(\zeta)=\lim_{\epsilon \rightarrow 0}\left\{ \left( \frac{\zeta}{1-\zeta}+
1281: \frac{1-\zeta}{\zeta}+\zeta(1-\zeta)\right) \theta(1-\zeta-\epsilon )+\delta(1-\zeta)\ln \epsilon
1282: \right\} C_{A}+b_{2}\delta(1-\zeta).
1283: \end{equation}
1284: 
1285: With Eq.~(\ref{69}) in hand, it is possible to check the quality of the NLL approximation against
1286: exact answers. As shown in Ref.~\cite{we4}, the soft-gluon corrections reproduce satisfactorily the
1287: threshold behavior of the available exact results for $\lambda\sim1$. Since the gluon distribution
1288: function supports just the threshold region, the soft-gluon contribution dominates the
1289: photon-hadron cross sections $\sigma_{k,GF}$ ($k=T,L,A,I$) at energies not so far from the
1290: production threshold and at relatively low virtuality $Q^{2}\lesssim m^{2}$.
1291: 
1292: \section{\label{exp}Nonperturbative IC and Relevant Experimental Facts}
1293: The most clean probe of the charm quark distribution function (both perturbative and
1294: nonperturbative) is the semi-inclusive deep inelastic lepton-proton scattering, $lp\rightarrow
1295: l'cX$. To measure the nonperturbative IC contribution, one needs data on the charm production at
1296: sufficiently large Bjorken $x$. The only experiment which has investigated the large $x$ domain is
1297: the European Muon Collaboration (EMC) \cite{EMC} where the decay lepton spectra have been used to
1298: detect the produced charmed particles. In Ref.~\cite{harris-emc}, a re-analysis of the EMC data on
1299: $F_{2}^{c}(x,Q^{2})$ have been performed using the NLO results for both GF and QS components. The
1300: analysis \cite{harris-emc} shows that a nonperturbative intrinsic charm contribution to the proton
1301: wave function of the order of $1\%$ is needed to fit the EMC data in the large $x$ region. This
1302: value of the nonperturbative IC is consistent with the estimates based on the operator product
1303: expansion \cite{polyakov}. Note however that the EMC data are of limited statistics and, for this
1304: reason, more accurate measurements of charm leptoproduction at large $x$ are necessary.
1305: 
1306: It is also possible to extract useful information on the IC from diffractive dissociation processes
1307: such as $p\rightarrow p J/\psi$ on a nuclear target. Comprehensive measurements of the
1308: $pA\rightarrow J/\psi X$ and $\pi A \rightarrow J/\psi X$ cross sections have been performed in the
1309: fixed target experiments NA3 at CERN \cite{NA3} and E886 at FNAL \cite{E886}. According to the
1310: arguments presented in Refs.~\cite{brod-psi-1,hoyer-psi-1,brod-psi-2}, the IC contribution is
1311: predicted to be strongly shadowed in the above reactions that is in a complete agreement with the
1312: observed nuclear dependence of the high Feynman $x_{F}$ component of the $J/\psi$ hadroproduction.
1313: 
1314: A non-vanishing five-quark Fock component $\left\vert uudc\bar{c}\right\rangle$ leads to the
1315: production of open charm states such as $\Lambda_{c}(cud)$ and $D^{-}(\bar{c}d)$ with large Feynman
1316: $x_{F}$. This may occur either through a coalescence of the valence and charm quarks which are
1317: moving with the same rapidity or via hadronization of the produced $c$ and $\bar{c}$. As shown in
1318: Refs.~\cite{barger-lead,brod-lead}, a model based on the nonperturbative intrinsic charm naturally
1319: explains the leading particle effect in the $pp\rightarrow DX$ and $pp\rightarrow \Lambda_{c}X$
1320: processes that has been observed at the ISR \cite{ISR-lead} and Fermilab
1321: \cite{E791-lead-1,E791-lead-2}.
1322: 
1323: As to the high-$x_{F}$ hadroproduction of open bottom states like $\Lambda_{b}(bud)$, corresponding
1324: cross sections are predicted to be suppressed as $m_{c}^{2}/m_{b}^{2}\sim 1/10$ in comparison with
1325: the case of charm production. Evidence for the forward $\Lambda_{b}$ production in the $pp$
1326: collisions at the ISR energy was reported in Refs.~\cite{ISR-B-1,ISR-B-2}.
1327: 
1328: Rare seven-quark fluctuations of the type $\left\vert uudc\bar{c}c\bar{c}\right\rangle$ in the
1329: proton wave function can lead to the production of two $J/\psi$ \cite{brod-7-quark} or a
1330: double-charm baryon state at large $x_{F}$ and low $p_{T}$. Double $J/\psi$ events with a high
1331: combined $x_{F}\geq 0.5$ have been detected in the NA3 experiment \cite{NA3-7-quark}. An
1332: observation of the double-charmed baryon $\Xi^{+}_{cc}(3520)$ with mean $\langle x_{F}\rangle\simeq
1333: 0.33$ has been reported by the SELEX collaboration at FNAL \cite{SELEX-7-quark}.
1334: \begin{thebibliography}{99}
1335: \bibitem{BHPS} S.~J.~Brodsky, P.~Hoyer, C.~Peterson, and N.~Sakai,
1336: Phys.\ Lett.\ B {\bf 93}, 451 (1980).
1337: \bibitem{BPS} S.~J.~Brodsky, C.~Peterson, and N.~Sakai,
1338: Phys.\ Rev.\ D {\bf 23}, 2745 (1981).
1339: \bibitem{brod1} S.~J.~Brodsky, \emph{"Light-front QCD"}, hep-ph/0412101.
1340: \bibitem{brod2} S.~J.~Brodsky, Few Body Syst. {\bf 36}, 35 (2005).
1341: \bibitem{polyakov} M.~Franz, V.~Polyakov, and K.~Goeke,
1342: Phys.\ Rev.\ D {\bf 62}, 074024 (2000).
1343: \bibitem{ACOT} M.~A.~G.~Aivazis, J.~C.~Collins, F.~I.~Olness, and W.~-K.~Tung,
1344: Phys.\ Rev.\ D {\bf 50}, 3102 (1994).
1345: \bibitem{collins} J.~C.~Collins, Phys.\ Rev.\ D {\bf 58}, 094002 (1998).
1346: \bibitem{grib-lip} V.~N.~Gribov and L.~N.~Lipatov, Sov.\ J.\ Nucl.\ Phys. {\bf 15}, 438 (1972).
1347: \bibitem{dokshitzer} Y.~L.~Dokshitzer, Sov.\ Phys.\ JETP {\bf 46}, 641 (1977).
1348: \bibitem{alt-par} G.~Altarelli and G.~Parisi, Nucl.\ Phys.\ B {\bf 126}, 298 (1977).
1349: \bibitem{pumplin} J.~Pumplin, Phys.\ Rev.\ D {\bf 73}, 114015 (2006).
1350: \bibitem{brod-higgs} S.~J.~Brodsky, B.~Kopeliovich, I.~Schmidt, and J.~Soffer,
1351: Phys.\ Rev.\ D {\bf 73}, 113005 (2006).
1352: \bibitem{CTEQ6} J.~Pumplin, D.~R.~Stump, J.~Huston, H.~L.~Lai, P.~Nadolsky,
1353: and W.~K.~Tung, JHEP {\bf 0207}, 012 (2002).
1354: \bibitem{MRST2004} A.~D.~Martin, R.~G.~Roberts, W.~J.~Stirling, and R.~S.~Thorne,
1355: Phys.\ Lett.\ B {\bf 604}, 61 (2004).
1356: \bibitem{kid2} N.~Kidonakis, Phys.\ Rev.\ D {\bf 73}, 034001 (2006).
1357: \bibitem{kid1} N.~Kidonakis, Phys.\ Rev.\ D {\bf 64}, 014009 (2001).
1358: \bibitem{Mangano-N-R} M.~L.~Mangano, P.~Nason, and G.~Ridolfi,
1359: Nucl.\ Phys.\ B {\bf 373}, 295 (1992).
1360: \bibitem{Frixione-M-N-R} S.~Frixione, M.~L.~Mangano, P.~Nason, and G.~Ridolfi,
1361: Nucl.\ Phys.\ B {\bf 412}, 225 (1994).
1362: \bibitem{we1} N.~Ya.~Ivanov, A.~Capella, and A.~B.~Kaidalov,
1363: Nucl.\ Phys.\ B {\bf 586}, 382 (2000).
1364: \bibitem{we2} N.~Ya.~Ivanov, Nucl.\ Phys.\ B {\bf 615}, 266 (2001).
1365: \bibitem{we4} N.~Ya.~Ivanov, Nucl.\ Phys.\ B {\bf 666}, 88 (2003).
1366: \bibitem{we3} N.~Ya.~Ivanov, P.~E.~Bosted, K.~Griffioen, and S.~E.~Rock,
1367: Nucl.\ Phys.\ B {\bf 650}, 271 (2003).
1368: \bibitem{E161} SLAC E161 (2000), http://www.slac.stanford.edu/exp/e160.
1369: \bibitem{dombey} N.~Dombey, Rev.\ Mod.\ Phys.\ {\bf 41}, 236 (1969).
1370: \bibitem{HM} E.~Hoffman and R.~Moore, Z.\ Phys.\ C {\bf 20}, 71 (1983).
1371: \bibitem{KS} S.~Kretzer and I.~Schienbein, Phys.\ Rev.\ D {\bf 58}, 094035 (1998).
1372: \bibitem{eRHIC} A.~Deshpande, R.~Milner, R.~Venugopalan, and W.~Vogelsang,
1373: Ann.\ Rev.\ Nucl.\ Part.\ Sci.\ {\bf 55}, 165 (2005).
1374: \bibitem{EIC} See also http://www.bnl.gov/eic for information concernig
1375: the eRHIC/EIC project.
1376: \bibitem{LHeC} J.~B.~Dainton, M.~Klein, P.~Newman, E.~Perez, and F.~Willeke,
1377: hep-ex/0603016.
1378: \bibitem{we6} L.~N.~Ananikyan and N.~Ya.~Ivanov, hep-ph/0609074.
1379: \bibitem{KS-thesis} I. Schienbein, hep-ph/0110292.
1380: \bibitem{GP} H.~Georgi and H.~D.~Politzer, Phys.\ Rev.\ Lett.\ {\bf 40}, 3 (1978).
1381: \bibitem{Mendez} A.~M\'{e}ndez, Nucl.\ Phys.\ B {\bf 145}, 199 (1978).
1382: \bibitem{Fano} U.~Fano, Phys.\ Rev.\ {\bf 93}, 121 (1954).
1383: \bibitem{LW1} J.~P.~Leveille and T.~Weiler, Phys.\ Rev.\ D {\bf 24}, 1789 (1981).
1384: \bibitem{Watson} A.~D.~Watson, Z.\ Phys.\ C {\bf 12}, 123 (1982).
1385: \bibitem{LW2} J.~P.~Leveille and T.~Weiler, Nucl.\ Phys.\ B {\bf 147}, 147 (1979).
1386: \bibitem{LRSN} E.~Laenen, S.~Riemersma, J.~Smith, and W.~L.~van Neerven,
1387: Nucl.\ Phys.\ B {\bf 392}, 162 (1993).
1388: \bibitem{Laenen-Moch} E.~Laenen and S.~-O. Moch, Phys.\ Rev.\ D {\bf 59}, 034027 (1999).
1389: \bibitem{CTEQ5} H.~L.~Lai et al., Eur.\ Phys.\ J.\ C {\bf 12}, 375 (2000).
1390: \bibitem{kT} L.~Apanasevich et al., Phys.\ Rev.\ D {\bf 59}, 074007 (1999).
1391: \bibitem{AOT} M.~A.~G.~Aivazis, F.~I.~Olness, and W.~-K.~Tung,
1392: Phys.\ Rev.\ D {\bf 50}, 3085 (1994).
1393: \bibitem{chi} W.~-K. Tung, S.~Kretzer, and C.~Schmidt, J.\ Phys.\ G {\bf 28}, 983 (2002).
1394: \bibitem{CTEQ4} H.~L.~Lai et al., Phys.\ Rev.\ D {\bf 55}, 1280 (1997).
1395: \bibitem{SACOT} M.~Kramer, F.~I.~Olness, and D.~E.~Soper,
1396: Phys.\ Rev.\ D {\bf 62}, 096007 (2000).
1397: \bibitem{neerven} W.~L.~van Neerven, hep-ph/0107193.
1398: \bibitem{BMSN} M.~Buza, Y.~Matiounine, J.~Smith, and W.~L.~van Neerven,
1399: Eur.\ Phys.\ J.\ C {\bf 1}, 301 (1998).
1400: \bibitem{Ellis-Nason} R.~K.~Ellis and P.~Nason, Nucl.\ Phys.\ B {\bf 312}, 551 (1989).
1401: \bibitem{Smith-Neerven} J.~Smith and W.~L.~van Neerven,
1402: Nucl.\ Phys.\ B {\bf 374}, 36 (1992).
1403: \bibitem{BKNS} W.~Beenakker, H.~Kuijf, W.~L.~van Neerven, and J.~Smith,
1404: Phys.\ Rev.\ D {\bf 40}, 54 (1989).
1405: \bibitem{Nason-D-E-1} P.~Nason, S.~Dawson, and R.~K.~Ellis,
1406: Nucl.\ Phys.\ B {\bf 303}, 607 (1988).
1407: \bibitem{Nason-D-E-2} P.~Nason, S.~Dawson, and R.~K.~Ellis,
1408: Nucl.\ Phys.\ B {\bf 327}, 49 (1989).
1409: \bibitem{Nason-D-E-3} P.~Nason, S.~Dawson, and R.~K.~Ellis,
1410: Nucl.\ Phys.\ B {\bf 335}, 260 (1990).
1411: \bibitem{Contopanagos-L-S} H.~Contopanagos, E.~Laenen, and G.~Sterman,
1412: Nucl.\ Phys.\ B {\bf 484}, 303 (1997).
1413: \bibitem{Laenen-O-S} E.~Laenen, G.~Oderda, and G.~Sterman,
1414: Phys.\ Lett.\ B {\bf 438}, 173 (1998).
1415: \bibitem{Kidonakis-O-S} N.~Kidonakis, G.~Oderda, and G.~Sterman,
1416: Nucl.\ Phys.\ B {\bf 531}, 365 (1998).
1417: \bibitem{Harris-Smith} B.~W.~Harris and J.~Smith, Nucl.\ Phys.\ B {\bf 452}, 109 (1995).
1418: \bibitem{EMC} J.~J.~Aubert et al., Nucl.\ Phys.\ B {\bf 213}, 31 (1983).
1419: \bibitem{harris-emc} B.~W.~Harris, J.~Smith, and R.~Vogt,
1420: Nucl.\ Phys.\ B {\bf 461}, 181 (1996).
1421: \bibitem{NA3} J.~Badier et al., Z.\ Phys.\ C {\bf 20}, 101 (1983).
1422: \bibitem{E886} M.~J.~Leitch et al., Phys.\ Rev.\ Lett.\ {\bf 84}, 3256 (2000).
1423: \bibitem{brod-psi-1} S.~J.~Brodsky and P.~Hoyer,
1424: Phys.\ Rev.\ Lett.\ {\bf 63}, 1566 (1989).
1425: \bibitem{hoyer-psi-1} P.~Hoyer, M.~Vanttinen, and U.~Sukhatme,
1426: Phys.\ Lett.\ B {\bf 246}, 217 (1990).
1427: \bibitem{brod-psi-2} S.~J.~Brodsky, P.~Hoyer, A.~H.~Mueller, and W.~K.~Tang,
1428: Nucl.\ Phys.\ B {\bf 369}, 519 (1992).
1429: \bibitem{barger-lead} V.~D.~Barger, F.~Halzen, and W.~Y.~Keung,
1430: Phys.\ Rev.\ D {\bf 25}, 112 (1982).
1431: \bibitem{brod-lead} R.~Vogt and S.~J.~Brodsky, Nucl.\ Phys.\ B {\bf 478}, 311 (1996).
1432: \bibitem{ISR-lead} P.~Chauvat et al., Phys.\ Lett.\ B {\bf 199}, 304 (1987).
1433: \bibitem{E791-lead-1} E.~M.~Aitala et al., Phys.\ Lett.\ B {\bf 495}, 42 (2000).
1434: \bibitem{E791-lead-2} E.~M.~Aitala et al., Phys.\ Lett.\ B {\bf 539}, 218 (2002).
1435: \bibitem{ISR-B-1} M.~Basile et al., Nuovo\ Cim.\ A {\bf 65}, 408 (1981).
1436: \bibitem{ISR-B-2} G.~Bari et al., Nuovo\ Cim.\ A {\bf 104}, 1787 (1991).
1437: \bibitem{brod-7-quark} R.~Vogt and S.~J.~Brodsky, Phys.\ Lett.\ B {\bf 349}, 569 (1995).
1438: \bibitem{NA3-7-quark} J.~Badier et al., Phys.\ Lett.\ B {\bf 114}, 457 (1982).
1439: \bibitem{SELEX-7-quark} A.~Ocherashvili et al., Phys.\ Lett.\ B {\bf 628}, 18 (2005).
1440: \end{thebibliography}
1441: \end{document}
1442: