hep-ph0610137/Sf2.tex
1: \documentclass[11pt]{article}
2: 
3: 
4: \usepackage{psfrag,amsfonts,varioref,epsfig,tabularx,color}
5: %\usepackage{enumerate,verbatim,latexsym,graphicx,graphics}
6: 
7: \definecolor{mygreen}{rgb}{0.08860759493670886, 0.8, 0.16877637130801687}
8: \definecolor{mycyan}{rgb}{0, 0.75, 1}
9: \definecolor{mygold}{rgb}{0.6, 0.6, 0}
10: \definecolor{myred}{rgb}{0.8, 0, 0}
11: \usepackage{hhline}
12: \usepackage{multirow}
13: 
14: \setlength{\paperwidth}{21cm}
15: \setlength{\paperheight}{29.7cm}\setlength{\textwidth}{15.5cm}
16: \setlength{\textheight}{22.3cm}\setlength{\oddsidemargin}{0.19truecm}
17: \setlength{\evensidemargin}{0.19truecm}\setlength{\topmargin}{-0.62truecm}\setlength{\parindent}{2\parindent}
18: 
19: 
20: 
21: \newcommand{\rmt}[1]{\textrm{\tiny{#1}}}
22: \newcommand{\MSbar}{\ensuremath{\overline{{\rm MS}}}}
23: \newcommand{\LMS}{\Lambda_{\rmt{\MSbar}}}
24: \newcommand{\pmpm}[2]{\scriptsize{$\begin{array}{l}\!\!\!#1\\\!\!\!#2\end{array}$}}
25: \newcommand{\pmpmc}[2]{\scriptsize{$\begin{array}{c}\!\!\!#1\\\!\!\!#2\end{array}$}}
26: \def\nn{\nonumber}
27: \def\qq {\qquad}
28: \def\MI{{\cal I}}
29: \def\MC{{\cal C}}
30: \def\MO{{\cal O}}
31: \def\MM{{\cal M}}
32: 
33: \def\MR{{\cal R}}
34: \def\MS{{\cal S}}
35: \newcommand{\Rt}{\tilde{R}}
36: \newcommand{\Rh}{\hat{R}}
37: 
38: \newcommand{\lb}{\left}
39: \newcommand{\rb}{\right}
40: \newcommand{\rt}{\tilde{r}}
41: \newcommand{\pt}{\tilde{p}}
42: \newcommand{\tF}{\tilde{F}}
43: \newcommand{\tL}{\tilde{\Lambda}}
44: \newcommand{\nf}{\ensuremath{N_{f}}}
45: \newcommand{\tx}{\tilde{x}}
46: \newcommand{\tX}{\tilde{X}}
47: \newcommand{\Xh}{\hat{X}}
48: \newcommand{\rmGeV}{{\rm GeV}}
49: \newcommand{\GeVV}{{\rm GeV}^{2}}
50: \newcommand{\rmss}[1]{\textrm{\scriptsize{#1}}}
51: \newcommand{\D}{\Delta}
52: \newcommand{\hA}{\hat{A}}
53: 
54: 
55: 
56: 
57: \begin{document}
58: 
59: %%%%%%%%%%%%TITLE PAGE%%%%%%%%%%%%%%%%%%
60: \thispagestyle{empty}
61: 
62: \title{Improved analysis of moments of $F_3$ in neutrino-nucleon scattering using the Bernstein polynomial method}
63: \author{\textbf{Paul M. Brooks\footnote{e-mail:\texttt{brooks.pm@googlemail.com}}\,\, and C.J.
64: Maxwell\footnote{e-mail:\texttt{\texttt{c.j.maxwell@durham.ac.uk}}}}}
65: \date{\vskip10mm\emph{Institute for Particle Physics Phenomenology, University of Durham,\\South Road, Durham, DH1 3LE,
66: UK.}}
67: \vspace{5cm}
68: \maketitle %\thispagestyle{empty}
69: \large \vspace{-8.45cm}\hfill\vbox{\hbox{IPPP/06/64}
70:             \hbox{DCPT/06/128}
71:              \hbox{October 2006}}
72: \normalsize \vspace{8.45cm}
73: \begin{abstract}
74: \setlength{\baselineskip}{13pt}
75: 
76: %%%%%%%%%%%%ABSTR${AC}$T%%%%%%%%%%%%%%%%%%
77: We use recently calculated next-to-next-to-leading order (NNLO)
78: anomalous dimension coefficients for the moments of the $x{F}_{3}$
79: structure function in ${\nu}N$ scattering, together with the
80: corresponding three-loop Wilson coefficients, to obtain improved QCD
81: predictions for both odd and even moments of $F_{3}$. To investigate
82: the issue of renormalization scheme dependence, the Complete
83: Renormalization Group Improvement (CORGI) approach is used, in which
84: all dependence on renormalization and factorization scales is
85: avoided by a complete resummation of RG-predictable scale
86: logarithms. We also consider predictions using the method of
87: effective charges, and compare with the standard `physical scale'
88: choice. The Bernstein polynomial method is used to construct
89: experimental moments (from the $x{F}_{3}$ data of the CCFR
90: collaboration) that are insensitive to the value of $xF_{3}$ in the
91: region of $x$ which is inaccessible experimentally. Direct fits for
92: $\LMS^{(5)}$ ($\alpha_s(M_Z)$) are then performed. The CORGI fits
93: including target mass corrections give a value
94: $\alpha_s(M_Z)={0.1189}^{+0.0019}_{-0.0019}$, consistent with the
95: world average. The effective charge and physical scale fits give
96: slightly smaller values, which are still consistent within the
97: errors.
98: 
99: \end{abstract}
100: \newpage
101: 
102: \renewcommand{\thepage}{\arabic{page}}
103: \pagestyle{headings} \setlength{\baselineskip}{16pt}
104: \setcounter{page}{2} \setlength{\parskip}{8pt}
105: 
106: \pagestyle{plain}
107: \section{Introduction}\vspace{-\parskip}
108: %%%%%%%%%%%Introduction%%%%%%%%%%%
109: The measurements of the CCFR collaboration provide a precise
110: determination of the non-singlet deep inelastic scattering (DIS)
111: structure functions of neutrinos and anti-neutrinos on nucleons,
112: $xF_{3}(x,Q^2)$ \cite{r1}. Recently the NuTev collaboration have
113: also published measurements of $F_2(x,Q^2)$ and $F_3(x,Q^2)$
114: \cite{r1a}. The $Q^2$-dependence of moments of structure functions
115: can be predicted in perturbative QCD, and fits to the data can be
116: used to infer $\LMS$ (or equivalently $\alpha_s(M_Z)$). In doing
117: this, the principal difficulty is that there are upper and lower
118: limits on the experimentally accessible range of $x$ at low and high
119: $Q^2$, respectively. The moments are potentially sensitive to this
120: missing information, and this propagates into an additional level of
121: uncertainty in the resultant prediction of $\LMS$.
122: 
123: An approach which has been applied in the past is to use Bernstein
124: polynomials, which are peaked in a rather limited $x$-range, to
125: construct linear combinations of moments which are insensitive to
126: the missing $x$ regions \cite{r2,r3}. The analysis of Ref.~\cite{r2}
127: chose, as is customary, to work in the $\MSbar$ scheme and set both
128: renormalization and factorization scales to $Q$ (physical scale (PS)
129: choice). This was extended in Ref.~\cite{r4} to consider predictions
130: for $F_3$ obtained in the `complete renormalization group improved'
131: (CORGI) approach \cite{r5} in which all dependence on the
132: renormalization scale $\mu$ and the factorization scale $M$ is
133: eliminated by an all-orders resummation of RG-predictable scale
134: logarithms.  
135: 
136: The analyses of Refs.~\cite{r2,r3} and \cite{r4} used the then
137: state-of-the-art three-loop (NNLO) results for the anomalous
138: dimension and coefficient function, which were restricted to a
139: subset of odd moments $n=1,3,5,\ldots,13$ \cite{r6}. Recent progress
140: has yielded NNLO results for these quantities for {\it any} value of
141: $n$ \cite{r7,r8}. Consequently the set of Bernstein moments used in
142: the fits can now be greatly extended. The Bernstein polynomials
143: defined in Refs.~\cite{r2,r3} were linear combinations of {\it odd}
144: moments, but the new results of Refs.~\cite{r7,r8} mean that, by a
145: slight redefinition of the polynomials, even moments can now also be
146: studied.
147: 
148: 
149: In this paper we intend to perform such an extended analysis. We
150: shall fit the CCFR data \cite{r1} to PS and CORGI NNLO QCD
151: predictions, and will in addition compare with the predictions in
152: the closely-related method of effective charges approach (EC)
153: \cite{r9}. Target mass corrections and higher twist effects will
154: also be considered. We shall also compare our results to those
155: obtained using a fitting technique based on Jacobi Polynomials
156: \cite{r9A,r9B,r9C}.
157: 
158: 
159: The plan of the paper is to give a brief review of the factorization
160: and renormalization scheme dependence of structure function moments
161: in Section 2. In Section 3 we discuss the  CORGI and effective
162: charge approaches for leptoproduction moments. We take this
163: opportunity to correct an error in the expression for the NNLO CORGI
164: result for the scheme invariant $X_2$ derived in Ref.~\cite{r4}.
165: Section 4 will contain a  description of the Bernstein polynomial
166: averages to be employed in the fits. We shall show how to modify the
167: definition of the polynomials to accommodate both odd and even
168: moments. We then constrain the set of acceptable Bernstein moments
169: to be used in the fits by comparing how four different methods of
170: extrapolation (to obtain $xF_3$ on the full $x$-range) differ; this
171: enables us to define a `modelling error' to be combined with the
172: other sources of error in our analysis. In section 5 we give details
173: of the fitting procedure and in section 6 we present the results of
174: the fits to the PS, CORGI and EC predictions for the moments, and
175: consider how the fits change if target mass corrections and higher
176: twist corrections are included. Section 7 contains a discussion and
177: conclusions.
178: 
179: \section{Factorization and renormalization scheme dependence of the moments}\vspace{-\parskip}
180: 
181: \label{s:FRS}
182: 
183: The moments we are concerned with in this paper are those derived from
184: $F_{3}$ in (anti)neutrino-nucleon scattering. They are defined as:
185: \begin{eqnarray}
186: \MM_{3}^{\nu N}(n;Q^{2})&=&\int_{0}^{1}dx\;x^{n-1}F_{3}^{\nu
187: N}(x,Q^{2}). \label{eq:6moms}
188: \end{eqnarray}
189: These moments can be factorized in the following form,
190: \begin{eqnarray}
191: \MM_{3}^{\nu N}(n;Q^{2})&=&\langle
192: N|\MO_{n,\;\rmt{NS}}(M)|N\rangle\,\MC^{(3)}_{n}(Q,M,\mu,a(\mu)),
193: \label{eq:momfac}
194: \end{eqnarray}
195: where $\langle N|\MO_{n,NS}(M)|N\rangle$ is the non-singlet (NS)
196: operator matrix element of nucleon states and
197: $\MC^{(3)}_{n}(Q,M,\mu,a(\mu))$ is the coefficient function. Here
198: $a\equiv{\alpha}_{s}/\pi$ is the RG-improved coupling. The operator
199: matrix element is factorized at the scale $M$ into a
200: non-perturbative component and a perturbative expression, written in
201: terms of the coupling evaluated at the factorization scale,
202: $a=a(M)$. The factorization scale dependence is governed by the
203: anomalous dimension equation,
204: \begin{eqnarray}
205: M\frac{\partial}{\partial M}\langle
206: N|\MO_{n,\;\rmt{NS}}(M)|N\rangle&=&\langle
207: N|\MO_{n,\;\rmt{NS}}(M)|N\rangle\gamma_{n,\;\rmt{NS}}(a)
208: \label{eq:ad6a}.
209: \end{eqnarray}
210: Here ${\gamma}_{n,\;\rmt{NS}}(a)$ is the anomalous dimension of the
211: moment. It has the following perturbative expansion,
212: \begin{eqnarray}
213: {\gamma}_{n,\;\rmt{NS}}(a)&=&-d(n)a-d_{1}(n)a^{2}-d_{2}(n)a^{3}-d_{3}(n)a^{4}-\ldots,
214: \label{eq:ad6b}
215: \end{eqnarray}
216: where $d(n)$ is factorization scheme invariant, and the higher
217: coefficients serve to label the factorization scheme dependence. The
218: $M$-dependence of the coupling is governed by the beta-function
219: equation,
220: \begin{eqnarray}
221: M\frac{\partial a}{\partial M}&=&{\beta}(a)\;\equiv\;
222: -ba^{2}(1+ca+c_{2}a^{2}+c_{3}a^{3}+\ldots). \label{eq:6bfe}
223: \end{eqnarray}
224: Here $b=(33-2{N}_{f})/6$ and $c=(153-19N_f)/12b$ are renormalization
225: scheme (RS) invariant. The higher coefficients serve to label the RS
226: dependence. Together, these two equations determine the perturbative
227: behaviour of the operator matrix element. For the remainder of this
228: paper we simplify our notation by dropping the sub- and superscripts
229: `$\nu N$', `$n$', `$(3)$' and `NS', from the quantities in
230: Eqs.~(\ref{eq:momfac}) and (\ref{eq:ad6a}). Also, although the
231: coefficients $d_{i}(n)$ in Eq.~(\ref{eq:ad6b}) are $n$-dependent, we
232: also suppress this.
233: 
234: A solution to Eq.~(\ref{eq:ad6a}) can be obtained in the form,
235: \begin{eqnarray}
236: \langle\mathcal{O}(M)\rangle&=&A_{n}\exp\lb\{\int_{0}^{a(M)}\frac{\gamma(x)}{\beta(x)}dx-\int_{0}^{\infty}\frac{\gamma^{(1)}(x)}{\beta^{(2)}(x)}dx\rb\},
237: \label{eq:6operatorme}
238:  \end{eqnarray}
239: where $\gamma^{(i)}$ and $\beta^{(i)}$ denote the anomalous
240: dimension and beta-function equations truncated after $i$ terms.
241: There is a distinct parallel between the above equation and the
242: solution to the beta function equation. The second integral in
243: Eq.~(\ref{eq:6operatorme}) is an infinite constant. We are free to
244: choose any form we wish for this term, subject to the constraint
245: that it must have the same singularity structure as the first
246: integral. However, a particular choice for this constant corresponds
247: to a particular definition of $A_{n}$. Consequently, $A_{n}$ can be
248: likened to the dimensional transmutation parameter, $\Lambda$, in
249: that it defines the missing boundary condition in
250: Eq.~(\ref{eq:ad6a}). $A_{n}$  is actually a (set of)
251: non-perturbative constant(s), generated by the factorization
252: process, and they are factorization and renormalization scheme (FRS)
253: invariant. Their precise values cannot be calculated within
254: perturbation theory, and hence must be obtained by comparison with
255: experimental data.
256: 
257: The coefficient function $\MC(Q,M,\mu,a(\mu))$, depends on both the
258: renormalization {\it and} factorization scheme adopted, and it takes
259: the form of an expansion in powers of the coupling evaluated at the
260: renormalization scale,
261: \begin{eqnarray}
262:   \MC(Q,M,\mu,{\tilde{a}}(\mu))&=&1+r_{1}\tilde{a}+r_{2}\tilde{a}^{2}
263: +r_{3}\tilde{a}^{3}+\ldots,
264: \end{eqnarray}
265: where $\tilde{a}=a(M=\mu)$. Using the above equation together with
266: Eqs.~(\ref{eq:momfac}) and (\ref{eq:6operatorme}), the moments can
267: be written as \cite{r4,r10,r11},
268: \begin{eqnarray}
269: \MM(n;Q^{2})&=&A_{n}\lb(\frac{ca}{1+ca}\rb)^{d/b}\exp(\mathcal{I}(a))(1+r_{1}\tilde{a}+r_{2}\tilde{a}^{2}+r_{3}\tilde{a}^{3}+\cdots),
270: \label{eq:6protomoments}
271: \end{eqnarray}
272: where,
273: \begin{eqnarray}
274: \MI(a)&=&\int_{0}^{a}dx\frac{d_{1}+(d_{1}c+d_{2}-dc_{2})x+(d_{3}+cd_{2}-c_{3}d)x^{2}+\cdots)}{b(1+cx)(1+cx+c_{2}x^{2}+c_{3}x^{3}+\cdots)}.
275: \label{MI}
276: \end{eqnarray}
277: The explicit $M$ dependence of the coupling can be obtained by solving the
278: following transcendental equation \cite{r9a},
279: \begin{eqnarray}
280: \frac{1}{a}+c\ln\frac{ca}{1+ca}&=&b\ln\frac{M}{\tilde{\Lambda}}-b\int_{0}^{a}\left[\frac{1}{\beta(x)}-\frac{1}{\beta^{(2)}(x)}\right].
281: \label{eq:6trans}
282: \end{eqnarray}
283: Equation (\ref{eq:6protomoments}) serves as a prototypical
284: expression for the moments, from which CORGI, EC and PS predictions
285: can be derived.
286: 
287: 
288: The self-consistency of  perturbation theory means that the
289: perturbative coefficient $r_1$ has a dependence on $M$ and $d_1$.
290: The higher coefficients also have a dependence on the parameters
291: specifying the FRS, ${r_k} ({\mu},M,{c_2},c_3,{\ldots},{c_k}
292: ;{d_1},{d_2},{\ldots},{d_k})$. The explicit form of this
293: FRS-dependence can be determined by demanding that on calculating
294: the moments up to O($a^k$) the partial derivative with respect to
295: each FRS parameter is O($a^{k+1}$) \cite{r10,r11} . The complete set
296: of partial derivatives required to derive the FRS-dependence of
297: $r_1$, $r_2$ and $r_3$ is, for the $\mu$-dependence
298: \begin{eqnarray}
299: \mu\frac{\partial r_{1}}{\partial\mu}\;=\;0, \qq \mu\frac{\partial
300: r_{2}}{\partial\mu}\;=\;r_{1}b, \qq\mu\frac{\partial
301: r_{3}}{\partial\mu}\;=\;2r_{2}b+br_{1}c.\label{eq:r3mu}
302: \end{eqnarray}
303: For the $M$-dependence we have
304: \begin{eqnarray}
305: &&M\frac{\partial r_{1}}{\partial M}\;=\;d,
306: %\label{eq:r1M}
307: \qq M\frac{\partial r_{2}}{\partial M}\;=\;dr_{1}-dL+d_{1}, \nn
308: \\[10pt]
309: &&M\frac{\partial r_{3}}{\partial
310: M}\;=\;d_{2}+d_{1}r_{1}+dr_{2}-dr_{1}L-2d_{1}L+dL^{2}-dcL,\label{eq:r3M}
311: \end{eqnarray}
312: where we have defined $L\equiv b{\ln}(M/\mu)$. For the
313: $c_2$-dependence we have
314: \begin{eqnarray}
315: &&\frac{\partial r_{1}}{\partial c_{2}}\;=\;0, \label{eq:r1c2}
316: \qq\frac{\partial r_{2}}{\partial c_{2}}\;=\;-\frac{d}{2b},
317: \label{eq:r2c2}
318: \nn \\[10pt]
319: &&\frac{\partial r_{3}}{\partial
320: c_{2}}\;=\;-\frac{r_{1}d}{2b}+\frac{Ld}{b}+\frac{cd}{3b}-\frac{2d_{1}}{3b}-r_{1}.\label{eq:r3c2}
321: \end{eqnarray}
322: For the $c_3$-dependence
323: \begin{eqnarray}
324: \frac{\partial r_{1}}{\partial c_{3}}\;=\;0, \qq \frac{\partial
325: r_{2}}{\partial c_{3}}\;=\;0, \qq\frac{\partial r_{3}}{\partial
326: c_{3}}\;=\;-\frac{d}{6b}.\label{eq:r3c3}
327: \end{eqnarray}
328: For the $d_1$-dependence
329: \begin{eqnarray}
330: &&\frac{\partial r_{1}}{\partial d_{1}}\;=\;-\frac{1}{b}, \qq
331: \frac{\partial r_{2}}{\partial
332: d_{1}}\;=\;\frac{c}{2b}-\frac{r_{1}}{b}+\frac{L}{b},\nn\\[10pt]
333: &&\frac{\partial r_{3}}{\partial
334: d_{1}}\;=\;\frac{cr_{1}}{2b}-\frac{c^{2}}{3b}-\frac{r_{2}}{b}+\frac{c_{2}}{3b}-\frac{L^{2}}{b}+\frac{Lr_{1}}{b}.\label{eq:r3d1}
335: \end{eqnarray}
336: For the $d_2$-dependence we have
337: \begin{eqnarray}
338: \frac{\partial r_{1}}{\partial d_{2}}\;=\;0,\qq \frac{\partial
339: r_{2}}{\partial d_{2}}\;=\;-\frac{1}{2b},\qq \frac{\partial
340: r_{3}}{\partial
341: d_{2}}\;=\;\frac{c}{3b}+\frac{1}{2b}(2L-r_{1}).\label{eq:r3d2}
342: \end{eqnarray}
343: Finally, for the $d_3$-dependence we have
344: \begin{eqnarray}
345: \frac{\partial{r}_{1}}{\partial{d_3}}\;=\;0, \qq
346: \frac{\partial{r_2}}{\partial{d_3}}\;=\;0, \qq \frac{\partial
347: r_{3}}{\partial d_{3}}\;=\;-\frac{1}{3b}.\label{17}
348: \end{eqnarray}
349: These results may now be integrated to obtain $r_1$, $r_2$ and $r_3$. For $r_1$ we obtain
350: \begin{eqnarray}
351: r_{1}&=&\frac{d}{b}\tau_{M}-\frac{d_{1}}{b}-X_{0}(Q),
352: \label{eq:corgi1}
353: \end{eqnarray}
354: where $\tau_{M}=b\ln\lb(M/\tL\rb)$. $X_{0}(Q)$ is an FRS invariant
355: quantity, generated as a constant of integration. One can define an
356: FRS invariant, non-universal scale parameter, $\Lambda_{\MM}$, via
357: the FRS invariant $X_{0}(Q)$. Thus,
358: \begin{eqnarray}
359: \frac{d}{b}\tau_{M}-\frac{d_{1}}{b}-r_{1}&=&X_{0}(Q)\;\equiv\;
360: d\;\ln\lb(\frac{Q}{\Lambda_{\MM}}\rb). \label{eq:X0Lm}
361: \end{eqnarray}
362: For $r_2$ we obtain
363: \begin{eqnarray}
364: r_{2}&=&\Bigg{(}\frac{1}{2}-\frac{b}{2d}\Bigg{)}r_{1}^{2}+\frac{b}{d}r_{1}\tilde{r}_{1}+\frac{d_{1}}{d}r_{1}-\frac{dc_{2}}{2b}+\frac{d_{1}^{2}}{2bd}
365: +\frac{cd_{1}}{2b}-\frac{d_{2}}{2b}+X_{2}, \label{eq:corgi2}
366: \end{eqnarray}
367: where we have defined,
368: \begin{eqnarray}
369: \tilde{r}_{1}&\equiv&r_{1}(M=\mu) \nn\\
370: &=&\frac{d}{b}\tau_{\mu}-\frac{d_{1}}{b}-X_{0}(Q).
371: \label{eq:rtau}
372: \end{eqnarray}
373: Here $X_2$ is another FRS-invariant constant of integration.
374: Crucially $X_2$ and higher invariants are independent of $Q$. Hence,
375: the complete $Q$-dependence of the observable is generated by
376: $X_0(Q)$. Similarly, for $r_3$ we obtain
377: \begin{eqnarray}
378: r_{3}&=&\frac{c_{2}dc}{3b}
379: -\frac{c_{3}d}{6b}
380: +\frac{cd_{2}}{3b}
381: -\frac{d_3}{3b}
382: -\frac{2d_{1}^{3}}{3bd^{2}}
383: -\frac{cd_{1}^{2}}{2bd}
384: -\frac{c^{2}d_{1}}{3b}
385: +\frac{d_{1}d_{2}}{bd}
386: +\frac{c_{2}d_{1}}{3b}\nn\\
387: &-&\frac{2d_{1}^{2}r_{1}}{d^{2}}
388: -\frac{d_{1}r_{1}^{2}}{d}
389: +\frac{b^{2}r_{1}^{3}}{3d^{2}}-
390: \frac{r_{1}^{3}}{3}
391: -\frac{bcr_{1}^{2}}{2d}
392: +\frac{d_{2}r_{1}}{d}\nn\\
393: &-&\frac{2bd_{1}r_{1}\rt_{1}}{d^{2}}
394: -\frac{b^{2}r_{1}\rt_{1}^{2}}{d^{2}}
395:  - \frac{br_{1}^{2}\rt_{1}}{d}
396:  +\frac{bcr_{1}\rt_{1}}{d}\nn\\
397: &+&\frac{2b\rt_{1}r_{2}}{d}
398: +\frac{2d_{1}r_{2}}{d}
399: + r_{1}r_{2}
400: +X_{3}.
401: \label{eq:r3FRS}
402: \end{eqnarray}
403: Again $X_3$ is a $Q$-independent FRS-invariant constant of
404: integration. Using Eq.~(\ref{eq:corgi2}) $X_3$ can be written in
405: terms of $r_1$, ${\tilde{r}}_{1}$ and the other FRS parameters. This
406: also holds for the higher invariants. The results of
407: Eqs.~(\ref{eq:r3mu}) - (\ref{17}) and of Eqs.~(\ref{eq:corgi2}) and
408: (\ref{eq:r3FRS}) replace, respectively, Eqs.~(15) and (18) of
409: Ref.~\cite{r4} which contain several errors. The invariant $X_2$ can
410: be obtained from NNLO results for the anomalous dimension
411: coefficients and coefficient function in {\it any} FRS. For instance
412: if we make the customary choice of $\MSbar$ with $M=\mu=Q$ then
413: $r_1={\tilde{r_1}}$ and we obtain
414: \begin{eqnarray}
415: X_{2}&=&\lb.r_{2}-\Bigg{(}\frac{1}{2}+\frac{b}{2d}\Bigg{)}r_{1}^{2}-\frac{d_{1}}{d}r_{1}+\frac{dc_{2}}{2b}-\frac{d_{1}^{2}}{2bd}
416: -\frac{cd_{1}}{2b}+\frac{d_{2}}{2b}\rb|_{\MSbar}. \label{eq:6:X2MS}
417: \end{eqnarray}
418: 
419: In summary, through Eqs.~(\ref{eq:corgi1}), (\ref{eq:corgi2}) and
420: (\ref{eq:r3FRS}), we have determined the explicit FRS dependence of
421: the coefficients $r_{i}$. In doing so, we have generated a set of
422: FRS invariant quantities $X_{i}$, the importance of which will
423: become clear when we come to consider the CORGI form of the moments
424: in the following section.
425: 
426: 
427: \section{PS, CORGI and EC predictions}\vspace{-\parskip}
428: 
429: The standard physical scale approach is to set $M=\mu=Q$ and adopt
430: $\MSbar$ subtraction. Setting $M=\mu$ implies that $a=\tilde{a}$,
431: and hence the moments have the form,
432: \begin{eqnarray}
433: \MM(n;Q^{2})&=&A_{n}\lb(\frac{ca}{1+ca}\rb)^{d/b}\lb(1+R_{1}a+R_{2}a^{2}+\ldots\rb).
434: \label{eq:PSM}
435: \end{eqnarray}
436: The coefficients $R_{i}$ can be determined by expanding
437: Eq.~(\ref{eq:6protomoments}) in powers of $a$,
438: \begin{eqnarray}
439: R_{1}&=&r_{1}+\frac{d_{1}}{b}\label{eq:6:R1}\\
440: R_{2}&=&r_{2}+\frac{d_{1}^{2}}{2b^{2}}-\frac{cd_{1}}{2b}+\frac{r_{1}d_{1}}{b}-\frac{dc_{2}}{2b}+\frac{d_{2}}{2b},
441: \label{eq:6:R2}
442: \end{eqnarray}
443: and the coupling in this expression is the three-loop $\MSbar$
444: coupling with $\mu=Q$.
445: 
446: The CORGI idea (see Ref.~\cite{r5} for a detailed discussion) is
447: that all RG-predictable information about higher perturbative
448: coefficients, available at a given fixed-order of calculation should
449: be resummed to all-orders. Given an NLO calculation for instance one
450: knows $X_0(Q)$ but not $X_2,\;X_3$ or higher FRS-invariants. One
451: should therefore resum to all-orders all the terms {\it not}
452: involving these unknown invariants. As discussed in Section 2 these
453: terms are multinomials in
454: $r_1,{\tilde{r}_{1}},c_2,\ldots,c_i,d_1,d_2,\ldots,d_i,\ldots$.
455: 
456: 
457: Crucially this all-orders sum must be FRS-invariant, as separately
458: must be the subset of terms involving $X_2$, $X_3\ldots$. One may
459: exploit this invariance and choose to use the FRS where all the FRS
460: parameters are zero, ${r_1}={\tilde{r}_{1}}=
461: c_2=\ldots={c}_{i}=\ldots=d_1=d_2=\ldots={d}_{i}=\ldots=0$.
462: 
463: Setting $r_1={\tilde{r}_{1}}=0$ means that $\mu=M$, setting  $d_1=0$
464: then implies that (from Eq.~(\ref{eq:X0Lm})) ${\tau}_{M}=
465: b{\ln}(Q/{\Lambda}_{\MM})$. Also, with $c_{i}={d}_{i}=0$ the
466: integral $\MI(a)$ of Eq.~(\ref{MI}) vanishes, and one finally
467: obtains the CORGI form of the moments,
468: \begin{eqnarray}
469: \MM(n;Q^{2})&=&A_{n}\lb(\frac{ca_0}{1+ca_0}\rb)^{d/b}\lb(1+X_{2}a_{0}^{2}+X_{3}a_{0}^{3}+\ldots\rb).
470: \label{eq:CORGImoms}
471: \end{eqnarray}
472: Here the CORGI coupling $a_0$ is the coupling in a 't Hooft scheme
473: \cite{r11a}, in which $c_i=0$ $(i>1)$. This can be written in terms
474: of the Lambert $W$ function defined implicitly by $W(z){e}^{W(z)}=z$
475: \cite{r12,r13},
476: \begin{eqnarray}
477: a_{0}(Q)&=&\frac{-1}{c\lb[1+W_{-1}(z(Q))\rb]}, \label{eq:6tHooft}
478: \end{eqnarray}
479: with,
480: \begin{eqnarray}
481: z(Q)&=&-\frac{1}{{\rm e}}\lb(\frac{Q}{{\Lambda}_{\MM}}\rb)^{-b/c}.
482: \end{eqnarray}
483: $W_{-1}$ refers to the branch of the Lambert $W$ function required
484: for asymptotic freedom, the nomenclature being that of
485: Ref.~\cite{r14}. $\Lambda_{\MM}$ is the invariant scale connected
486: with the $X_0(Q)$ FRS-invariant, defined in Eq.~(\ref{eq:X0Lm}).
487: Since it is an FRS-invariant it can be evaluated in any FRS.
488: Choosing the $\MSbar$ scheme with $M=\mu=Q$ one finds
489: \begin{eqnarray}
490: \Lambda_{\MM}&=&\LMS{\left(\frac{2c}{b}\right)}^{-c/b}\exp\lb\{\frac{d_{1}}{db}+\frac{r_{1}}{d}\rb\},
491: \label{eq:6:LMM}
492: \end{eqnarray}
493: with $r_{1}$ and $d_{1}$ calculated in \MSbar~with $M=\mu=Q$. The
494: factor of ${\left(\frac{2c}{b}\right)}^{-c/b}$ converts to the
495: standard convention for integrating the beta-function equation and
496: defining $\LMS$ (see Ref.~\cite{r14a} for further details). The
497: second factor on the r.h.s.~of Eq.~(\ref{eq:CORGImoms}) resums to
498: all-orders the RG-predictable terms not involving $X_2,X_3,\ldots$.
499: The $a_0^2$ term sums to all-orders the RG-predictable terms
500: involving $X_2$, but not $X_3,X_4,\ldots$, etc.
501: 
502: The CORGI result corresponds to  an $\MSbar$ scale choice
503: $M=\mu=xQ$, with
504: \begin{eqnarray}
505: x&=&x_{\rmt{CORGI}}\;\equiv\;\exp\lb\{-\frac{d_{1}}{db}-\frac{r_{1}}{d}\rb\}.
506: \end{eqnarray}
507: To illustrate how the CORGI scale differs from the PS choice ($x=1$)
508: we plot in table 1 the $x_{\rmt{CORGI}}$, for the first $20$ moments
509: $n=1,2,\ldots,20$. We also tabulate the corresponding $X_2(n)$ NNLO
510: CORGI invariants obtained from Eq.~(\ref{eq:6:X2MS}). The anomalous
511: dimension coefficients up to NNLO are taken from Refs.~\cite{r7,r8},
512: and the coefficient function from Ref.~\cite{r6}. We assume $N_f=5$
513: active quark flavours.
514: \begin{table}[t]
515: \begin{center}
516: \begin{tabular}{|r|rr|r|}
517: \hline $n$&$x_{\rmt{CORGI}}$&&$X_2(n)$
518: \\\hline
519: 1&0.4688  &&-1\\
520: 2&0.7156&&-3.01\\
521: 3&0.5074&&-3.28\\
522: 4&0.42966&&-3.485\\
523: 5&0.3838&&-3.627\\
524: 6&0.3530&&-3.713\\
525: 7&0.3300&&-3.766\\
526: 8&0.312 &&-3.792\\
527: 9&0.2974&&-3.8\\
528: 10&0.2853&&-3.793\\
529: \hline
530: \end{tabular}
531: \qq\qq\qq\begin{tabular}{|r|rr|r|} \hline
532: $n$&$x_{\rmt{CORGI}}$&&$X_2(n)$
533: \\\hline
534: 11&0.2749&&-3.776\\
535: 12&0.2659&&-3.749\\
536: 13&0.2580&&-3.716\\
537: 14&0.2510&&-3.677\\
538: 15&0.2447&&-3.633\\
539: 16&0.2390&&-3.586\\
540: 17&0.2338&&-3.536\\
541: 18&0.2291&&-3.483\\
542: 19&0.2248&&-3.428\\
543: 20&0.2207&&-3.372\\
544: \hline
545: \end{tabular}
546: %\begin{tabular}{|r|rr|r|}
547: %\hline $n$&$x_{\rmt{CORGI}}$&&$X_2(n)$
548: %\\\hline
549: %1&0.4688  &&-1\\
550: %2&0.7156&&-3.01\\
551: %3&0.5074&&-3.28\\
552: %4&0.42966&&-3.485\\
553: %5&0.3838&&-3.627\\
554: %6&0.3530&&-3.713\\
555: %7&0.3300&&-3.766\\
556: %8&0.312 &&-3.792\\
557: %9&0.2974&&-3.8\\
558: %10&0.2853&&-3.793\\
559: %11&0.2749&&-3.776\\
560: %12&0.2659&&-3.749\\
561: %13&0.2580&&-3.716\\
562: %14&0.2510&&-3.677\\
563: %15&0.2447&&-3.633\\
564: %16&0.2390&&-3.586\\
565: %17&0.2338&&-3.536\\
566: %18&0.2291&&-3.483\\
567: %19&0.2248&&-3.428\\
568: %20&0.2207&&-3.372\\
569: %\hline
570: %\end{tabular}
571: \end{center}
572: \caption{The numerical values of $x_{\rmt{CORGI}}$ and the NNLO
573: CORGI invariants ${X}_{2}(n)$ for the $n=1-20$ moments of $F_3$.}
574: \label{xcorgi}
575: \end{table}
576: We see from table \ref{xcorgi} that as $n$ increases the CORGI scale
577: decreases, becoming significantly less than $x=1$ (PS). The $X_2(n)$
578: invariants are seen to be moderate in size.
579: 
580: 
581: 
582: 
583: Finally, we discuss the third variant of perturbative QCD which we
584: shall consider. By setting $M=\mu$ and rearranging, we can recast
585: the perturbation series for $\MM(n;Q^{2})$ of
586: Eq.~(\ref{eq:6protomoments}) in the form
587: \begin{eqnarray}
588: \MM(n;Q^{2})&=&A_{n}\lb(c\Rt(a)\rb)^{d/b}, \label{eq:ECmoms}
589: \end{eqnarray}
590: where
591: \begin{eqnarray}
592: \Rt(a)&=&a+\Rt_{1}a^{2}+\Rt_{2}a^{3},
593: \end{eqnarray}
594: is an effective charge \cite{r9},
595: and the coefficients $\Rt_{i}$ have the form,
596: \begin{eqnarray}
597: \Rt_{1}&=&\frac{bcR_{1}}{d}-c^{2}\nn\\
598: &=&\frac{bcr_{1}}{d}+\frac{d_{1}c}{d}-c^{2},\\
599: \Rt_{2}&=&\frac{bcR_{2}}{d}+\frac{b^{2}R^{2}_{1}c}{2d^{2}}-\frac{bR_{1}^{2}c}{2d}-\frac{bR_{1}c^{2}}{d}+c^{3}\nn\\
600: &=&\frac{bcr_{2}}{d}+\frac{bd_{1}r_{1}c}{d^{2}}+\frac{d_{2}c}{2d}-\frac{c_{2}c}{2}\nn\\
601: &+&\frac{b^{2}r^{2}_{1}c}{2d^{2}}
602: -\frac{br_{1}^{2}c}{2d}
603: +\frac{d_{1}^{2}c}{2d^{2}}
604: -\frac{br_{1}c^{2}}{d}-\frac{3d_{1}c^{2}}{2d}+c^{3}.
605: \end{eqnarray}
606: Here $R_1$ and $R_2$ are the coefficients defined in
607: Eqs.~(\ref{eq:6:R1}) and (\ref{eq:6:R2}). Rather than integrating
608: the effective charge beta-function we shall instead apply CORGI to
609: the effective charge, avoiding the need to numerically solve a
610: transcendental equation which would make the fitting to data
611: considerably more complicated. At NLO the CORGI and EC results agree
612: exactly. We have the CORGI result
613: \begin{eqnarray}
614: \MM(n;Q^{2})&=&A_{n}c^{d/b}\lb(a_{0}+\tX_{2}a_{0}^{3}+\tX_{3}a_{0}^{4}+\ldots\rb)^{d/b}.
615: \label{eq:6ECH}
616: \end{eqnarray}
617: In this case, the $\tX_{i}$ coefficients are the CORGI invariants
618: corresponding to single scale RS-dependence \cite{r5}. They have the
619: form,
620: \begin{eqnarray}
621: \tX_{2}&=&\Rt_{2}-\Rt_{1}^{2}-c\Rt_{1}+c_{2},
622: \label{eq:x2Rt}\\
623: \tX_{3}&=&\Rt_{3}-3\Rt_{1}\Rt_{2}+2\Rt_{1}^{3}+\frac{c\Rt_{1}^{2}}{2}-\Rt_{1}c_{2}+\frac{1}{2}c_{3}.\label{eq:x3Rt}
624: \end{eqnarray}
625: The CORGI coupling $a_0$ is that of Eq.~(\ref{eq:6tHooft}) but with
626: the scale $\Lambda_{\MM}$ now defined by,
627: \begin{eqnarray}
628: \Lambda_{\MM}^{\rmt{EC}}&=&\lb(\frac{2c}{b}\rb)^{-c/b}\exp\lb(\frac{\Rt_{1}}{b}\rb)\LMS.
629: \label{eq:6echL}
630: \end{eqnarray}
631: We shall refer to this variant of perturbation theory as `EC' for
632: simplicity, even though as noted above it is really CORGI applied to
633: a single-scale effective charge.
634: 
635: 
636: We note that we can streamline the calculation of the FRS-invariants
637: $X_i$ by using the single-scale effective charge. If we set $M=\mu$
638: in Eq.~(\ref{eq:6protomoments}), then the moments reduce to a
639: single-scale problem \cite{r5}. We can then rearrange the resultant
640: expression in terms of an effective charge, $\Rh(a)$,
641: \begin{eqnarray}
642: \MM(n;Q^{2})&=&A_{n}\lb(\frac{c\Rh(a)}{1+c\Rh(a)}\rb)^{d/b}.\label{Malt}
643: \end{eqnarray}
644: $\Rh(a)$ has the form,
645: \begin{eqnarray}
646: \Rh(a)&=&a+\Rh_{1}a^{2}+\Rh_{2}a^{3}+\Rh_{3}a^{4}+\ldots.
647: \end{eqnarray}
648: The coefficients $\Rh_{i}$ can be determined by expanding
649: Eqs.~(\ref{eq:PSM}) and (\ref{Malt}) in powers of $a$ and then
650: equating coefficients. They are found to be,
651: \begin{eqnarray}
652: \Rh_{1}&=&\frac{b}{d}R_{1},\label{eq:6:Rh1}\\
653: \Rh_{2}&=&\frac{b}{d}\lb(R_{2}+cR_{1}-\frac{R_{1}^{2}}{2}+\frac{bR_{1}^{2}}{2d}\rb),
654: \label{eq:6:Rh2}
655: \end{eqnarray}
656: where $R_{1}$ and $R_{2}$ are given by Eqs.~(\ref{eq:6:R1}) and
657: (\ref{eq:6:R2}). If we then CORGI-ize this effective charge, we have
658: a new set of FRS
659: invariants \cite{r5},
660: \begin{eqnarray}
661: \Xh_{0}&=&b\ln\frac{M}{\tL}-\Rh_{1},
662: \label{eq:6:xh0}\\
663: \Xh_{2}&=&\Rh_{2}-\Rh_{1}^{2}-c\Rh_{1}+c_{2},
664: \label{eq:6:xh2}
665: \end{eqnarray}
666: and the moments become,
667: \begin{eqnarray}
668: \MM(n;Q^{2})&=&A_{n}\lb(\frac{c\lb(a_{0}+\Xh_{2}
669: a_{0}^{3}\rb)}{1+c\lb(a_{0}+\Xh_{2} a_{0}^{3}\rb)}\rb)^{d/b}.
670: \end{eqnarray}
671: Expanding this into a form which we can compare with Eq.~(\ref{eq:CORGImoms}), gives,
672: \begin{eqnarray}
673: \MM(n;Q^{2})&=&A_{n}\lb(\frac{ca_{0}}{1+ca_{0}}\rb)^{d/b}\lb(1+\frac{d}{b}\Xh_{2}\,a_{0}^{2}+\ldots\rb).
674: \label{eq:6:momsX2alt}
675: \end{eqnarray}
676: Isolating the $\MO(a_{0}^{2})$ in the RHS bracket of the above
677: equation, and then using Eqs.~(\ref{eq:6:xh2}), (\ref{eq:6:Rh1}),
678: (\ref{eq:6:Rh2}), (\ref{eq:6:R1}) and (\ref{eq:6:R2}), gives,
679: \begin{eqnarray}
680: \frac{d}{b}\Xh_{2}&=& \frac{d}{b}\lb(\Rh_{2}-\Rh_{1}^{2}-c\Rh_{1}+c_{2}\rb)
681: \label{eq:6:X2altRh}\\
682: &=&r_{2}-\Bigg{(}\frac{1}{2}+\frac{b}{2d}\Bigg{)}r_{1}^{2}-\frac{d_{1}}{d}r_{1}+\frac{dc_{2}}{2b}-\frac{d_{1}^{2}}{2bd}
683: -\frac{cd_{1}}{2b}+\frac{d_{2}}{2b}.
684: \end{eqnarray}
685: So we see that the coefficient of the $\MO(a^{2})$ term in
686: Eq.~(\ref{eq:6:momsX2alt}) is the FRS invariant $X_{2}$, of
687: Eq.~(\ref{eq:6:X2MS}) with $\mu=M$ ($r_{1}=\rt_{1}$). Isolating the
688: $a^{3}$ term will yield $X_{3}$, and so on for higher $X_{i}$.
689: 
690: The coupling in Eq.~(\ref{eq:6:momsX2alt}) is the 't Hooft coupling
691: of Eq.~(\ref{eq:6tHooft}), but with the scale parameter determined
692: by Eq.~(\ref{eq:6:xh0}). Evaluating Eq.~(\ref{eq:6:xh0}) in
693: \MSbar~with $M=\mu=Q$ and using the standard definition of the
694: single-scale RS invariant $\Xh_{0}$ \cite{r9,r9a} gives,
695: \begin{eqnarray}
696: \Xh_{0}(Q)&\equiv&b\ln\frac{Q}{{\Lambda}_{\MM}}\\
697: &=&b\ln\frac{Q}{\tL}-\Rh_{1}.
698: \end{eqnarray}
699: Comparison with Eq.~(\ref{eq:X0Lm}) then reveals that
700: $\Xh_{0}(Q)=(b/d)X_0(Q)$. Using the same procedure we can obtain
701: expressions for $X_3$ and higher CORGI invariants.
702: 
703: 
704: \subsubsection*{Non-perturbative effects}\vspace{-\parskip}
705: 
706: The three variants of NNLO perturbative QCD, PS, CORGI, and EC, can
707: all be computed given $\MSbar$ anomalous dimension coefficients up
708: to NNLO \cite{r7,r8}, and the coefficient function \cite{r6}.
709: However, these perturbative predictions will be subject to
710: non-perturbative corrections in the form of $\MO\lb(1/Q^{2}\rb)$
711: terms. The two principal sources of these terms are: higher twist
712: terms and effects due to the mass of the target hadron.
713: 
714: The perturbative form of the moments is derived under the assumption
715: that the mass of the target hadron is zero (in the limit
716: $Q^{2}\rightarrow\infty$). At intermediate and low $Q^{2}$ this
717: assumption will begin to break down and the moments will be subject
718: to potentially significant power corrections, of order
719: $\MO\lb(m_{N}^{2}/Q^{2}\rb)$, where $m_N$ is the mass of the
720: nucleon. These are known as target mass corrections (TMCs) and when
721: included, the $F_{3}$ moments have the form \cite{r15,r15a},
722: \begin{eqnarray}
723: \MM^{\rmt{TMC}}(n;Q^{2})&=&\MM(n;Q^{2})+\frac{n(n+1)}{n+2}\frac{m_{N}^{2}}{Q^{2}}\MM(n+2;Q^{2})+\MO\lb(\frac{m_{N}^{4}}{Q^{4}}\rb).\label{eq:6:TMC}
724: \end{eqnarray}
725: 
726: 
727: The moments will also be subject to corrections from sub-leading
728: twist contributions to the OPE. These effects are poorly understood
729: and hence we only estimate them; this is done by means of an unknown
730: parameter, $A_{\rmt{HT}}$. The estimate has the form \cite{r2},
731: \begin{eqnarray}
732: \MM^{\rmt{HT}}(n;Q^{2})&=&n\lb(A^{\rmt{HT}}\frac{\LMS^{2}}{Q^{2}}\rb)\MM(n;Q^{2}),
733: \end{eqnarray}
734: and the value of $A^{\rmt{HT}}$ is obtained by fitting to data. Due
735: to the poorly understood nature of these effects, we do not include
736: the above term in the full analysis. Rather, we perform the analysis
737: with and without this term included, and take the difference in the
738: results as an estimate of the error associated with our ignorance of
739: the true nature of these effects.
740: 
741: The bottom quark mass threshold is within the range of $Q^{2}$
742: spanned by the available data for $F_{3}$. It is therefore necessary
743: to evolve the expressions for the moments over this threshold, and
744: in order to do this we use the formalism of Ref.~\cite{r16a}. We use
745: massless QCD with 4 quarks for $Q^{2}\leq m_{b}^{2}$ and massless
746: QCD with 5 quarks for $Q^{2}>m_{b}^{2}$. Here $m_b$ is the pole mass
747: of the $b$-quark with ${m}_{b}=4.85\pm{0.15}$ MeV \cite{r16}. From
748: the decoupling theorem, one finds the following relation between the
749: coupling above and below a quark threshold (denoted by
750: $a_{f+1}(Q^{2})$ and $a_{f}(Q^{2})$ respectively) \cite{r16a}
751: \begin{eqnarray}
752: a_{f}(m_{b}^{2})&=&a_{f+1}(m_{b}^{2})+\frac{11}{72}\lb(a_{f+1}(m_{b}^{2})\rb)^{3}.
753: \label{eq:5:decoup}
754: \end{eqnarray}
755: In practice, this matching is implemented by adopting different
756: values of the scale parameter in different $N_{f}$ regions. This is
757: governed by the following equations \cite{r16a},
758: \begin{eqnarray}
759: {\Lambda}_{N_{f}+1}^{2}&=&{\Lambda}_{N_{f}}^{2}\left(\frac{m_{N_{f}+1}^{2}}{{\Lambda}_{N_{f}}^{2}}\right)^{1-\frac{b^{N_{f}}}{b^{N_{f}+1}}}\times\exp\left(\frac{\delta_{\rmt{NLO}}+\delta_{\rmt{NNLO}}}{2b^{N_{f}+1}}\right),
760: \label{eq:matching}
761: \end{eqnarray}
762: where $\delta_{\rmt{NLO}}$ and $\delta_{\rmt{NNLO}}$ are given by
763: \begin{eqnarray}
764:  \delta_{\rmt{NLO}}&=&4(c^{N_{f}+1}-c^{N_{f}})\ln
765: L_{m}-4c^{N_{f}+1}\ln\frac{b^{N_{f}+1}}{b^{N_{f}}},\\
766: \delta_{\rmt{NNLO}}&=&\frac{8}{b^{N_{f}}L_{m}}\left(\left(c^{N_{f}+1}-c^{n_f}\right)c^{N_{f}}\ln
767:   L_{m}+\left(c^{N_{f}+1}\right)^{2}-\left(c^{N_{f}}\right)^{2}\nn\right.\\
768: &+&\left.c_{2}^{N_{f}}-c_{2}^{N_{f}+1}+\frac{7}{384}\right).\label{eq:5:mat3}
769: \end{eqnarray}
770: Here, ${\Lambda}_{N_{f}}$ is the scale parameter in the region where
771: $N_{f}$ quarks are active, $m_{N_{f}}$ is the pole mass of the $f$
772: quark, $b^{N_{f}}$, $c^{N_{f}}$ and $c_{2}^{N_{f}}$ are simply $b$,
773: $c$ and $c_{2}$ evaluated for $N_{f}$ quark flavours and we have
774: defined $L_{m}\equiv\ln\lb(m_{\nf+1}^{2}/{\Lambda}_{\nf}^{2}\rb)$.
775: Furthermore, we also demand continuity of the moments at the
776: threshold i.e.
777: \begin{eqnarray}
778: \lb.\MM(n;m_{b}^{2})\rb|_{\nf=4}&=&\lb.\MM(n;m_{b}^{2})\rb|_{\nf=5}.
779: \end{eqnarray}
780: As a consequence of this, the parameters $A_{n}$ also have different values in the
781:  $\nf=4$ and $\nf=5$ regions and their values are related by,
782: \begin{eqnarray}
783: A_{n}^{(5)}&=&\lb(\lb.\frac{A_{n}}{\MM(n;m_{b}^{2})}\rb|_{\nf=5}\,\rb)\lb(\lb.\frac{\MM(n;m_{b}^{2})}{A_{n}}\rb|_{\nf=4}\,\rb)A_{n}^{\nf=4}.
784: \end{eqnarray}
785: 
786: \section{The method of Bernstein averages}\vspace{-\parskip}
787: \begin{figure}
788: \begin{center}
789: \begin{tabular}{c c c}
790: \hspace{-.4cm}\includegraphics[angle=270,width=0.36\textwidth]{fig1.ps}&\hspace{-1cm}
791: \includegraphics[angle=270,width=0.36\textwidth]{fig2.ps}&\hspace{-1cm}
792: \includegraphics[angle=270,width=0.36\textwidth]{fig3.ps}\\
793: \hspace{-.4cm}\includegraphics[angle=270,width=0.36\textwidth]{fig4.ps}&\hspace{-1cm}
794: \includegraphics[angle=270,width=0.36\textwidth]{fig5.ps}&\hspace{-1cm}
795: \includegraphics[angle=270,width=0.36\textwidth]{fig6.ps}\\
796: \hspace{-.4cm}\includegraphics[angle=270,width=0.36\textwidth]{fig7.ps}&\hspace{-1cm}
797: \includegraphics[angle=270,width=0.36\textwidth]{fig9.ps}&\hspace{-1cm}
798: \includegraphics[angle=270,width=0.36\textwidth]{fig8.ps}\\
799: \hspace{-.4cm}\includegraphics[angle=270,width=0.36\textwidth]{fig10.ps}&\hspace{-1cm}
800: \includegraphics[angle=270,width=0.36\textwidth]{fig11.ps}&\hspace{-1cm}
801: \includegraphics[angle=270,width=0.36\textwidth]{fig12.ps}
802: \end{tabular}
803: \end{center}
804: \caption{Data for $xF_{3}$ plotted against $x$ for the 12 different $Q^{2}$
805:   bins of the CCFR data.}
806: \label{f:CCFR}
807: \end{figure}
808: When comparing theoretical predictions for moments of structure
809: functions with experimental data, we are faced with the
810: long-standing issue of missing data regions at high and low $x$ for
811: low and high $Q^{2}$ respectively. This is demonstrated in
812: Fig.~\ref{f:CCFR} in which we plot the CCFR data \cite{r1} for 12
813: different values of $Q^{2}$. We can see that at the lower range of
814: $Q^{2}$ we are limited to low-$x$ data, and that at high $Q^{2}$ we
815: are limited to the high $x$ range.
816: 
817: 
818: In order to reliably evaluate a moment at a particular $Q^{2}$, we
819: require data for the whole range of $x$. This being unavailable, we
820: are forced to make some guess about how the structure function
821: behaves in the missing data region. That is to say, we have to
822: choose some method of modelling (extrapolating and interpolating)
823: the data to cover the full range of $x$. However we wish to make the
824: evaluation of the experimental moments as free from QCD input as
825: possible, thus making the comparison between theory and experiment
826: as direct as possible. To this end, we shall adopt the approach
827: involving Bernstein averages \cite{r2,r3}; objects which, though
828: related to the moments, have negligible dependence on the modelling
829: method adopted (and hence on the behaviour of the structure function
830: in the missing data regions).
831: 
832: 
833: We define the Bernstein
834: polynomials as follows,
835: \begin{eqnarray}
836: \qq
837: p_{nk}(x^{2})&=&2\frac{\Gamma\lb(n+\frac{3}{2}\rb)}{\Gamma\lb(k+\frac{1}{2}\rb)\Gamma\lb(n-k+1\rb)}x^{2k}(1-x^{2})^{n-k},\qq
838: n,k\in\mathbb{I}. \label{eq:BP}
839: \end{eqnarray}
840: These functions are constructed such that they are zero at the endpoints
841: $x=0$ and $x=1$, and they are also normalized such that
842: $\int_{0}^{1}p_{nk}(x)dx=1$. Furthermore, if we constrain $n$ and $k$ such that $n\geq
843: k\geq0$, then $p_{nk}(x)$ are peaked sharply in some region between the
844: two endpoints.
845: 
846: 
847: 
848: The Bernstein polynomials can be treated as a distribution, with a
849: mean,
850: \begin{eqnarray}
851: \overline{x}_{nk}&=&\int_{0}^{1}x\;p_{nk}(x)\;dx\\
852: &=&\frac{\Gamma(k+1)\Gamma(n+\frac{3}{2})}{\Gamma(k+\frac{1}{2})\Gamma(n+2)},
853: \label{eq:BPmean}
854: \end{eqnarray}
855: and variance,
856: \begin{eqnarray}
857: \Delta x_{nk}&=&\int_{0}^{1}\lb(x-\overline{x}_{nk}\rb)^{2}\;p_{nk}(x)\;dx\\
858: &=&\frac{k+\frac{1}{2}}{n+\frac{3}{2}}-\lb(\frac{\Gamma(k+1)\Gamma(n+\frac{3}{2})}{\Gamma(k+\frac{1}{2})\Gamma(n+2)}\rb)^{2}.
859: \label{eq:BPvar}
860: \end{eqnarray}
861: The Bernstein averages of $F_{3}$ are then defined by,
862: \begin{eqnarray}
863: F_{nk}(Q^{2})&=&\int_{0}^{1}p_{nk}(x^{2})F_{3}(x,Q^{2})dx.
864: \label{eq:BAs}
865: \end{eqnarray}
866: Thus $F_{nk}(Q^{2})$ is the average of the structure function
867: weighted such that the region around $\overline{x}_{nk}$ is
868: emphasized. By picking the values of $n$ and $k$ wisely, we can
869: construct a set of averages which enhance the region for which we
870: have data for $F_{3}$ and de-emphasize the regions where there are
871: gaps. Therefore, in the resultant averages, the dependence on the
872: missing data regions will be heavily suppressed.
873: 
874: 
875: Defining this more carefully,
876: for a given value of $Q^{2}$, we only consider averages for which the range,
877: \begin{eqnarray}
878: \overline{x}_{nk}-\sqrt{\Delta x_{nk}}\;\le\;
879: x\;\le\;\overline{x}_{nk}+\sqrt{\Delta x_{nk}}, \label{eq:interval}
880: \end{eqnarray}
881: lies entirely within the region for which we have data. The only exception to this is
882: that if the
883: highest-$x$ data point lies within this range, then we {\it do} accept this
884: average, but only if the data suggests that $xF_{3}$ vanishes rapidly beyond this point.
885: 
886: The construction of an acceptable average, and the resultant suppression of
887: the missing data region is demonstrated in Fig.~\ref{f:berndemo}. We see that
888: the shaded (dark grey) missing data regions almost disappear in the right
889: hand plot.
890: \begin{figure}
891: \begin{center}
892: \begin{tabular}{c c c}
893: \hspace{-.018\textwidth}\includegraphics[angle=270,width=0.33\textwidth]{f62.ps}&\hspace{-.036\textwidth}
894: \includegraphics[angle=270,width=0.33\textwidth]{dat9.ps}&\hspace{-.036\textwidth}
895: \includegraphics[angle=270,width=0.33\textwidth]{pnkdat9.ps}
896: \end{tabular}
897: \end{center}
898: \vspace*{-0.15\textheight}
899: \begin{eqnarray*}
900: \qquad\!\!\!\!\times\frac{1}{x}\qquad
901: \qquad\qquad\qquad\quad\qquad\;=\qquad
902: \end{eqnarray*}
903: \vspace*{0.8truecm}
904: \caption{Constructing the Bernstein average, $F_{62}(Q^{2}=50.1\;{\rm{GeV}}^{2})$. The
905:   light grey region represents the interval in Eq.~(\ref{eq:interval}) and the dark grey
906:   areas represent the missing data regions. The small size of the
907:   dark grey
908:   region in the right hand plot demonstrates that this average will have
909:   negligible dependence on the missing data regions. Note that the right
910:   hand plot actually shows the {\it integrand} of the Bernstein average. The average
911:   itself will be this function integrated over $[0,1]$.}
912: \label{f:berndemo}
913: \end{figure}
914: 
915: 
916: By expanding the integrand of Eq.~(\ref{eq:BAs}) in powers of $x$, and using
917: Eq.~(\ref{eq:6moms}), we can relate the averages directly
918: to the moments,
919: \begin{eqnarray}
920: F_{nk}(Q^{2})&=&\frac{2\Gamma(n+\frac{3}{2})}{\Gamma(k+\frac{1}{2})}\sum_{l=0}^{n-k}\frac{(-1)^{l}}{l!(n-k-l)!}\MM(2(k+l)+1;Q^{2}),
921: \label{eq:BAsMoms}
922: \end{eqnarray}
923: and so theoretical predictions for the averages can be obtained by
924: substitution of Eqs.~(\ref{eq:CORGImoms}), (\ref{eq:PSM}) and
925: (\ref{eq:6ECH}) into the above expression. The Bernstein average is
926: seen to be a linear combination of {\it odd} moments. Due to the
927: unavailability of results for $d_{2}$ for even $n$, previous NNLO
928: analyses of this kind have been limited to the inclusion of only odd
929: $F_{3}$ moments. However, now that the NNLO calculation of the NS
930: anomalous dimension is complete \cite{r7,r8}, we are no longer
931: constrained in such a way. In light of this, we define a new set of
932: {\it modified} Bernstein polynomials,
933: \begin{eqnarray}
934: \qq
935: \pt_{nk}(x^{2})&=&2\frac{\Gamma\lb(n+2\rb)}{\Gamma\lb(k+1\rb)\Gamma\lb(n-k+1\rb)}x^{2k+1}(1-x^{2})^{n-k},\qq
936: n,k\in\mathbb{I}, \label{eq:mBP}
937: \end{eqnarray}
938: which include only odd powers of $x$ and hence whose averages are
939: related to even moments. These modified Bernstein polynomials are
940: simply the original polynomials of Eq.~(\ref{eq:BP}), multiplied be
941: $x$, and then `re-normalized' such that they still satisfy
942: $\int_{0}^{1}\pt_{nk}(x)dx=1$. We can calculate the mean and
943: variance of $\pt_{nk}(x)$,
944: \begin{eqnarray}
945: \overline{\tx}&=&
946: \frac{\Gamma(k +\frac{3}{2})\Gamma(n+2)}{\Gamma(k+1)\Gamma(n+\frac{5}{2})},\label{eq:mBPmean}\\
947: \Delta \tx&=&
948: \frac{k+1}{n+2}
949: -\lb[\frac{\Gamma(k+\frac{3}{2})\Gamma(n+2)}{\Gamma(k+1)\Gamma(n+\frac{5}{2})}\rb]^{2},
950: \label{eq:mBPvar}
951: \end{eqnarray}
952: and in analogy with Eq.~(\ref{eq:BAs}), we define the modified
953: Bernstein averages,
954: \begin{eqnarray}
955: \tF_{nk}(Q^{2})&=&\int_{0}^{1}\pt_{nk}(x^{2})F_{3}(x,Q^{2})dx.
956: \label{eq:mBAs}
957: \end{eqnarray}
958: Again, we only accept experimental modified averages for which the range,
959: \begin{eqnarray}
960: \overline{\tx}_{nk}-\sqrt{\Delta \tx_{nk}}\;\le\; x\;\le\;\overline{\tx}_{nk}+\sqrt{\Delta \tx_{nk}},
961: \label{eq:minterval}
962: \end{eqnarray}
963: lies within the region for which we have data. We obtain theoretical
964: predictions for the modified averages using the equation,
965: \begin{eqnarray}
966: \tF_{nk}(Q^{2})&=&\frac{2\Gamma(n+2)}{\Gamma(k+1)}\sum_{l=0}^{n-k}\frac{(-1)^{l}}{l!(n-k-l)!}\MM(2(k+l)+2;Q^{2}),
967: \label{eq:mBAsMoms}
968: \end{eqnarray}
969: which is seen to be a linear combination of even moments.
970: 
971: In order to calculate averages from data for $F_{3}$, we need an
972: expression for $xF_{3}$ covering the entire range of $x$, for each
973: value of $Q^{2}$. As mentioned above, the values of the moments
974: calculated in this way will depend on how we model the structure
975: functions in the missing data regions but for the averages, this
976: dependence is suppressed. However, we would like to test this
977: assertion, and so we use four different methods of modelling $F_{3}$
978: and perform our analysis separately for each method. Significant
979: differences between the results would signify a failure of the
980: Bernstein average method, and in instances where this is the case,
981: we reject that particular average at that particular $Q^{2}$.
982: Moderate deviation however, is acceptable, provided that we use the
983: magnitude of the deviation as an estimate of the error associated
984: with the missing data region. This error is then included as a
985: `modelling error' in the final result. In this way, we can almost
986: completely remove any dependence on missing data regions, and also
987: quantify the error associated with any residual dependence.
988: 
989: The four extrapolation methods we use are described below:
990: \renewcommand{\labelenumi}{{\bf \Roman{enumi}}}
991: \begin{enumerate}
992: \item In the first method, we fit the function,
993: \begin{eqnarray}
994: xF_{3}(x)&=&{\cal{A}}x^{\cal{B}}(1-x)^{\cal{C}}, \label{eq:ABC}
995: \end{eqnarray}
996: to the data for each fixed value of $Q^{2}$. The parameters ${\cal{A}}$, ${\cal{B}}$ and ${\cal{C}}$
997: are obtained by performing $\chi^{2}$ fitting of Eq.~(\ref{eq:ABC}) to data
998: for $F_{3}$. They are $Q^{2}$-dependent
999: quantities, and errors on their values are obtained by performing
1000: the fitting with the data for $F_{3}$ shifted to the two extremes of the
1001: error bars.
1002: 
1003: A justification for the particular form of fitting function in
1004: Eq.~(\ref{eq:ABC}) can be found in Ref.~\cite{r17}. However, the
1005: simple fact that this function fits the data well is justification
1006: enough, since the Bernstein averages are independent of the
1007: extrapolation method.
1008: 
1009: \item The second method we use is linear interpolation between successive data
1010: points. We also extrapolate beyond the data range, to the endpoints $xF_{3}(x)|_{x=0}=xF_{3}
1011: (x)|_{x=1}=0$, in order to be consistent with method {\bf I}.
1012: 
1013: \item The third method consists of using the fitting function of Eq.~(\ref{eq:ABC}), but
1014: setting $xF_{3}(x)=0$ everywhere outside the region for which we have data.
1015: 
1016: \item In analogy with {\bf III}, in this method we use the linear interpolation of
1017: method {\bf II} but setting $xF_{3}(x)=0$ everywhere outside the data region.\label{en:mod4}
1018: 
1019: \end{enumerate}
1020: \renewcommand{\labelenumi}{\arabic{enumi}}
1021: 
1022: 
1023: The deviation between the results obtained from the above methods
1024: (in particular, the difference between the first two and the last
1025: two) will be a good measure of the effectiveness of the Bernstein
1026: average method.
1027: 
1028: 
1029: 
1030: 
1031: 
1032: Data for $xF_{3}$ in neutrino-nucleon scattering is available from
1033: the CCFR collaboration \cite{r1}. The data was obtained from the
1034: scattering of neutrinos off iron nuclei and the measurements span
1035: the  ranges $1.26\leq Q^{2}\leq199.5\;{\rm{GeV}}^2$ and $0.015\leq
1036: x\leq0.75$. The $x$-ranges covered at each $Q^{2}$ are depicted in
1037: Fig.~\ref{f:x-ranges}.
1038: \begin{figure}
1039: \begin{center}
1040:   \includegraphics[angle=270,width=0.5\textwidth]{xrange.ps}
1041: \caption{Diagram depicting the $x$-ranges covered by the CCFR data,
1042: at
1043:   different $Q^{2}$.}\label{f:x-ranges}
1044: \end{center}
1045: \end{figure}
1046: 
1047: 
1048: In Fig.~\ref{f:4fits} we show each of the four modelling methods
1049: applied to $F_{3}$ measured at $Q^{2}=79.4\;{\rm{GeV}}^2$. Also
1050: shown on these figures (in grey) are the fits which are used to
1051: determine the errors on the modelling, which propagate through to
1052: errors on the averages.
1053: \begin{figure}[h]
1054: \begin{center}
1055: \begin{tabular}{c c}
1056: \hspace{-.45cm}\includegraphics[angle=270,width=0.5\textwidth]{fitABC.ps}&
1057: \includegraphics[angle=270,width=0.5\textwidth]{fitINT.ps}\\[-20pt]
1058: \vspace{-0.6cm}\hspace{-.45cm}\includegraphics[angle=270,width=0.5\textwidth]{fitABC2.ps}&
1059: \includegraphics[angle=270,width=0.5\textwidth]{fitINT2.ps}
1060: \end{tabular}
1061: \caption{The four methods used for fitting the structure functions.
1062: Here we show the
1063:   measured values of $xF_{3}$ at $Q^{2}=79.4\;\GeVV$. The errors are
1064:   determined by re-performing the fitting for the data shifted to the extremes
1065:   of the error bars, and this is denoted by grey lines.}\label{f:4fits}
1066: \end{center}
1067: \end{figure}
1068: 
1069: From the CCFR (and using the methods {\bf I} - {\bf IV} outlined
1070: above) we can obtain expressions describing the behaviour of the
1071: structure function over the full range of $x$, for each value of
1072: $Q^{2}$. It is then possible to extract experimental values of the
1073: averages, using the methods outlined below:
1074: 
1075: In the case of {\bf I}, obtaining the averages is particularly simple. Substituting Eq.~(\ref{eq:ABC}) into
1076: Eqs.~(\ref{eq:BAs}) and (\ref{eq:mBAs}) gives,
1077: \begin{eqnarray}
1078: F^{(\rmss{exp})}_{nk}&=&\mathcal{A}\frac{2\Gamma(n+\frac{3}{2})}{\Gamma(k+\frac{1}{2})}\sum_{l=0}^{n-k}\frac{(-1)^{l}}{l!(n-k-l)!}{\rm
1079:   B}(2(k+l)+\mathcal{B},\mathcal{C}+1),
1080: \label{eq:BAexp}
1081: \end{eqnarray}
1082: for the Bernstein averages and,
1083: \begin{eqnarray}
1084: \tF^{(\rmss{exp})}_{nk}&=&\mathcal{A}\frac{2\Gamma(n+2)}{\Gamma(k+1)}\sum_{l=0}^{n-k}\frac{(-1)^{l}}{l!(n-k-l)!}{\rm
1085:   B}(2(k+l)+\mathcal{B}+1,\mathcal{C}+1),
1086: \label{eq:mBAexp}
1087: \end{eqnarray}
1088: for the modified Bernstein averages. Here,
1089: B$(x,y)\equiv\Gamma(x)\Gamma(y)/\Gamma(x+y)$ is the Beta function.
1090: Once values for ${\cal{A}}$, ${\cal{B}}$, and ${\cal{C}}$, have been
1091: obtained, substitution into the above expressions leads directly to
1092: the averages.
1093: 
1094: In the case of {\bf II}, each of the averages is split into $j+1$ sections (where $j$
1095: is the number of data points) and each section is an integral over a polynomial of
1096: order $2n+1$. It is then reasonably simple to evaluate the averages by
1097: computing this set of integrals. This approach also applies to method {\bf IV}, but
1098: in this case there are only $j-1$ integrals.
1099: 
1100: For method {\bf III} we simply integrate the fitting function,
1101: multiplied by the Bernstein polynomials, with the integration limits
1102: being the values of $x$ at the first and last data point.
1103: 
1104: Having outlined the method for obtaining the experimental averages
1105: we now turn our attention to which averages are acceptable at which
1106: energies. The highest moment we use is the 18th moment and the
1107: lowest the 1st. Inclusion of higher moments than this leads to {\it
1108: no} significant increase in the number of acceptable Bernstein
1109: averages. The upper limit of $n=18$ implies that the highest
1110: Bernstein averages included are $F_{8k}$ and $\tF_{8k}$ and that the
1111: lowest used are $F_{10}$ and $\tilde{F}_{10}$. We exclude the
1112: averages for which $n=k$ as they simply correspond to individual
1113: moments themselves. This leaves us with a total of 72 (36 +
1114: $\tilde{36}$) potential averages at our disposal for each value of
1115: $Q^{2}$. This number will be reduced when we come to exclude
1116: averages on the basis of the acceptance criteria. After applying the
1117: acceptance criteria, we are left with 132 data points for the
1118: standard Bernstein averages and 141 for the modified Bernstein
1119: averages. Exactly which averages we use at a particular $Q^{2}$, can
1120: be determined by inspecting the plots in the results section.
1121: 
1122: In Fig.~\ref{f:ranges} we plot the dominant regions of the Bernstein
1123: polynomials (given by Eq.~(\ref{eq:interval})) for each of the used
1124: averages. This is superimposed onto the data-range diagram of
1125: Fig.~\ref{f:x-ranges}. Figure \ref{f:ranges2} shows equivalent plots
1126: for the modified Bernstein averages. These plots can be used to
1127: identify which averages are acceptable for a particular value of
1128: $Q^{2}$.
1129: \begin{figure}[h!]
1130: \begin{center}
1131: \hspace*{-6pt}\begin{tabular}{c c} \psfrag{x}{$x$}
1132: \psfrag{Qt}{{\scriptsize$\!\!\!\!Q^{2}/\GeVV$}}
1133: \psfrag{F10}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{10}$}}}
1134: \psfrag{F20}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{20}$}}}
1135: \psfrag{F30}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{30}$}}}
1136: \psfrag{F40}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{40}$}}}
1137: \psfrag{F50}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{50}$}}}
1138: \psfrag{F60}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{60}$}}}
1139: \psfrag{F70}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{70}$}}}
1140: \psfrag{F80}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{80}$}}}
1141: \psfrag{F87}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{87}$}}}
1142: \includegraphics[width=0.49\textwidth]{accept1.eps}&
1143: \psfrag{x}{$x$} \psfrag{Qt}{{\scriptsize$\!\!\!\!Q^{2}/\GeVV$}}
1144: \psfrag{F21}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{21}$}}}
1145: \psfrag{F31}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{31}$}}}
1146: \psfrag{F41}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{41}$}}}
1147: \psfrag{F51}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{51}$}}}
1148: \psfrag{F61}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{61}$}}}
1149: \psfrag{F71}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{71}$}}}
1150: \psfrag{F81}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{81}$}}}
1151: \psfrag{F76}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{76}$}}}
1152: \psfrag{F86}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{86}$}}}
1153: \includegraphics[width=0.49\textwidth]{accept2.eps}\\
1154: \psfrag{x}{$x$}
1155: \psfrag{Qt}{{\scriptsize$\!\!\!\!Q^{2}/\GeVV$}}
1156: \psfrag{F32}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{32}$}}}
1157: \psfrag{F42}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{42}$}}}
1158: \psfrag{F52}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{52}$}}}
1159: \psfrag{F62}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{62}$}}}
1160: \psfrag{F72}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{72}$}}}
1161: \psfrag{F82}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{82}$}}}
1162: \psfrag{F65}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{65}$}}}
1163: \psfrag{F75}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{75}$}}}
1164: \psfrag{F85}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{85}$}}}
1165: \includegraphics[width=0.49\textwidth]{accept3.eps}&
1166: \psfrag{x}{$x$}
1167: \psfrag{Qt}{{\scriptsize$\!\!\!\!Q^{2}/\GeVV$}}
1168: \psfrag{F43}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{43}$}}}
1169: \psfrag{F53}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{53}$}}}
1170: \psfrag{F63}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{63}$}}}
1171: \psfrag{F73}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{73}$}}}
1172: \psfrag{F83}{\textcolor{white}{\scriptsize{$\!\!\!\!F_{83}$}}}
1173: \psfrag{F54}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{54}$}}}
1174: \psfrag{F64}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{64}$}}}
1175: \psfrag{F74}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{74}$}}}
1176: \psfrag{F84}{\textcolor{black}{\scriptsize{$\!\!\!\!F_{84}$}}}
1177: \includegraphics[width=0.49\textwidth]{accept4.eps}
1178: \end{tabular}
1179: \end{center}
1180: %\vspace{-5.5cm}\hspace{7.2cm}$x$\hspace{7.35cm}$x$\hspace{-7.35cm}\vspace{5.5cm}\hspace{-7.2cm}
1181: 
1182: \caption{The black and light grey bars ($\vdash\!\dashv$) show  $x$ ranges covered by the CCFR data at
1183:   different energies. Superimposed onto these, in various colours,
1184:   are the peaked regions of the individual Bernstein polynomials, defined by the interval in
1185:  Eq.~(\ref{eq:interval}).}
1186: \label{f:ranges}
1187: \end{figure}
1188: \begin{figure}[h!]
1189: \begin{center}
1190: \hspace*{-6pt}\begin{tabular}{c c}
1191: &\\
1192: \psfrag{x}{$x$}
1193: \psfrag{Qt}{{\scriptsize$\!\!\!\!Q^{2}/\GeVV$}}
1194: \psfrag{F10}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{10}$}}}
1195: \psfrag{F20}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{20}$}}}
1196: \psfrag{F30}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{30}$}}}
1197: \psfrag{F40}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{40}$}}}
1198: \psfrag{F50}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{50}$}}}
1199: \psfrag{F60}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{60}$}}}
1200: \psfrag{F70}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{70}$}}}
1201: \psfrag{F80}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{80}$}}}
1202: \psfrag{F87}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{87}$}}}
1203: \includegraphics[width=0.49\textwidth]{accept1x.eps}&
1204: \psfrag{x}{$x$}
1205: \psfrag{Qt}{{\scriptsize$\!\!\!\!Q^{2}/\GeVV$}}
1206: \psfrag{F21}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{21}$}}}
1207: \psfrag{F31}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{31}$}}}
1208: \psfrag{F41}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{41}$}}}
1209: \psfrag{F51}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{51}$}}}
1210: \psfrag{F61}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{61}$}}}
1211: \psfrag{F71}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{71}$}}}
1212: \psfrag{F81}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{81}$}}}
1213: \psfrag{F76}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{76}$}}}
1214: \psfrag{F86}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{86}$}}}
1215: \includegraphics[width=0.49\textwidth]{accept2x.eps}\\
1216: \psfrag{x}{$x$}
1217: \psfrag{Qt}{{\scriptsize$\!\!\!\!Q^{2}/\GeVV$}}
1218: \psfrag{F32}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{32}$}}}
1219: \psfrag{F42}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{42}$}}}
1220: \psfrag{F52}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{52}$}}}
1221: \psfrag{F62}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{62}$}}}
1222: \psfrag{F72}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{72}$}}}
1223: \psfrag{F82}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{82}$}}}
1224: \psfrag{F65}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{65}$}}}
1225: \psfrag{F75}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{75}$}}}
1226: \psfrag{F85}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{85}$}}}
1227: \includegraphics[width=0.49\textwidth]{accept3x.eps}&
1228: \psfrag{x}{$x$}
1229: \psfrag{Qt}{{\scriptsize$\!\!\!\!Q^{2}/\GeVV$}}
1230: \psfrag{F43}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{43}$}}}
1231: \psfrag{F53}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{53}$}}}
1232: \psfrag{F63}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{63}$}}}
1233: \psfrag{F73}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{73}$}}}
1234: \psfrag{F83}{\textcolor{white}{\tiny{$\!\!\!\!\tF_{83}$}}}
1235: \psfrag{F54}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{54}$}}}
1236: \psfrag{F64}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{64}$}}}
1237: \psfrag{F74}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{74}$}}}
1238: \psfrag{F84}{\textcolor{black}{\tiny{$\!\!\!\!\tF_{84}$}}}
1239: \includegraphics[width=0.49\textwidth]{accept4x.eps}
1240: \end{tabular}
1241: \end{center}
1242: \caption{The black and light grey bars ($\vdash\!\dashv$) show  $x$ ranges covered by the CCFR data at
1243:   different energies. Superimposed onto these, in various colours,
1244:   are the peaked regions of the individual modified Bernstein polynomials, defined by the interval in
1245: Eq.~(\ref{eq:minterval}).}
1246: \label{f:ranges2}
1247: \end{figure}
1248: 
1249: 
1250: \section{Fitting procedure}\vspace{-\parskip} We use $\chi^{2}$
1251: minimization to optimize the fits of the theoretical predictions to
1252: the data. The highest moment included in the experimental averages
1253: is the 18th, and so when TMCs are included, we will require
1254: predictions for the first 20 moments. Therefore, the set of fitting
1255: parameters comprises of $\{A_{1}\ldots A_{20}\}$ plus the QCD scale
1256: parameter $\LMS$\@. When we include higher twist corrections this
1257: set is expanded to include $A_{\rmt{HT}}$\@. To check consistency
1258: between the odd and even moments we perform the analysis for each of
1259: these sets of moments separately and then finally together, and
1260: compare the results.
1261: 
1262: Although we stated previously that 132 standard Bernstein averages
1263: are available to us, in the CORGI case this is reduced to 130 for
1264: the following reason: When fitting predictions to the data, we scan
1265: values of $\LMS^{(4)}$ between 0 and 590 MeV for a minimum in
1266: $\chi^{2}$. Unfortunately, for $n=17$ and 19, the values of
1267: $\Lambda_{\MM}/\LMS$ (see table \ref{xcorgi}) are such that
1268: $Q^{2}=7.9\;\GeVV$ is below the Landau pole in
1269: Eq.~(\ref{eq:6tHooft}). Consequently we cannot obtain CORGI
1270: predictions for Bernstein averages which include these moments. From
1271: Eqs.~(\ref{eq:6:TMC}) and (\ref{eq:BAsMoms}) we can determine that
1272: this excludes $F_{80}(Q^{2}=7.9 \;{\rm{GeV}}^{2})$ and $F_{70}(Q^{2}=7.9\;{\rm{GeV}}^{2})$ from the
1273: fit. In the case of the modified Bernstein averages,
1274: $\tF_{80}(Q^{2}=7.9\;{\rm{GeV}}^{2})$ and $\tF_{70}(Q^{2}=7.9\;{\rm{GeV}}^{2})$ are already excluded
1275: due to their failure to meet the acceptance criteria.
1276: 
1277: 
1278: The CCFR data includes statistical errors and 18 different sources
1279: of systematic error. These errors cannot be added in quadrature, and
1280: so we perform the analysis for each of these 19 sources of error
1281: separately and then add the variation in the results in quadrature
1282: to obtain the final total error. We also include, as additional
1283: sources of error the deviation in results associated with using the
1284: four different modelling methods (this forms the `modelling error'
1285: in our final result), and the deviation in the results obtained from
1286: performing the analysis with and without HT corrections included
1287: (forming the `HT error').
1288: 
1289: 
1290: \subsubsection*{Correlation of errors}\vspace{-\parskip}
1291: When fitting theoretical predictions to experimental data using
1292: $\chi^{2}$ minimization, care must be taken in order to take into
1293: account fully the correlation between data points.
1294: 
1295: 
1296: To construct $\chi^{2}$ from a set of $N$ uncorrelated data points
1297: $\{f_{i}^{\rmt{exp.}}\}$ ($i=1,\ldots N$), with errors
1298: $\{\sigma_{f_{i}}\}$ and corresponding theoretical predictions
1299: $\{f_{i}^{\rmt{theo.}}\}$, we have,
1300: \begin{eqnarray}
1301: \chi^{2}&=&\sum_{i=1}^{N}\lb(\frac{f_{i}^{\rmt{exp.}}-f_{i}^{\rmt{theo.}}}{\sigma_{f_{i}}}\rb)^{2}.\label{eq:6:chinaive}
1302: \label{eq:6:chi2}
1303: \end{eqnarray}
1304: The raw data for $xF_{3}$ are uncorrelated. However, we are not
1305: comparing predictions for the structure functions themselves with
1306: data directly; rather we are doing so indirectly via the Bernstein
1307: averages. For a given value of $Q^{2}$, the full set of Bernstein
1308: averages (and modified Bernstein averages) we obtain {\it will} be
1309: correlated, due to their being derived from the same set of
1310: ($xF_{3}$) data points (see Fig.~\ref{f:berndemo}).
1311: 
1312: In the case where $\{f_{i}^{\rmt{exp.}}\}$ are correlated the
1313: $\chi^{2}$ function becomes \cite{r16b},
1314: \begin{eqnarray}
1315: \chi^{2}&=&\sum_{i=1}^{N}\sum_{j=1}^{N}\lb(f_{i}^{\rmt{exp.}}-f_{i}^{\rmt{theo.}}\rb)V^{-1}_{ij}\lb(f_{j}^{\rmt{exp.}}-f_{j}^{\rmt{theo.}}\rb)\nn\\[10pt]
1316: &=&\lb(\textbf{f}^{\;\rmt{exp.}}-\textbf{f}^{\;\rmt{theo.}}\rb)^{
1317: T}\textbf{V}^{-1}\lb(\textbf{f}^{\;\rmt{exp.}}-\textbf{f}^{\;\rmt{theo.}}\rb).\label{eq:6:chicovar}
1318: \end{eqnarray}
1319: In the second line of the above equation we have constructed vectors
1320: from the data points and their predictions. \textbf{V} is known as
1321: the covariance matrix. It encodes the correlation between each of
1322: the data points; its elements are obtained as follows,
1323: \begin{eqnarray}
1324: V_{ij}&=&{\rm cov}\lb(f_{i}, f_{j}\rb)\nn\\
1325: &=&\langle f_{i} f_{j}\rangle-\langle f_{i}\rangle \langle
1326: f_{j}\rangle.
1327: \end{eqnarray}
1328: If the $f_{i}$ are themselves functions of $M$ variables $x_{k}$
1329: (representing the `raw' data), then we have,
1330: \begin{eqnarray}
1331: {\rm cov}\lb(f_{k}, f_{l}\rb)&=&
1332: \sum_{i=1}^{M}\sum_{j=1}^{M}\lb(\frac{\partial f_{k}}{\partial
1333: x_{i}}\rb) \lb(\frac{\partial f_{l}}{\partial x_{j}}\rb) {\rm
1334: cov}\lb(x_{i},x_{j}\rb).
1335: \end{eqnarray}
1336: If the data for $x_{i}$ are uncorrelated, this reduces to,
1337: \begin{eqnarray}
1338: {\rm cov}\lb(f_{k}, f_{l}\rb)&=& \sum_{i=1}^{M}\lb(\frac{\partial
1339: f_{k}}{\partial x_{i}}\rb) \lb(\frac{\partial f_{l}}{\partial
1340: x_{i}}\rb)\sigma_{x_{i}}^{2}.\label{eq:5:cov}
1341: \end{eqnarray}
1342: To obtain the covariance of the Bernstein averages we simply
1343: substitute $F_{nk}(Q^{2})$ for $f_{i}$ and $F_{3}(x)$ for $x_{i}$ in
1344: Eq.~(\ref{eq:5:cov}). We assume no correlation between different
1345: values of $Q^{2}$ and so Eq.~(\ref{eq:6:chicovar}) decouples into 7
1346: different matrix equations (one for each of the values of $Q^{2}$
1347: between 7.9 and 125.9 $\GeVV$). By using the trapezium rule to
1348: approximate the Bernstein average integrals, we obtain the following
1349: expression for the covariance matrices:
1350: \begin{eqnarray}
1351: V_{lm}(Q_{i}^{2})&=&\sum_{j=1}^{N_{i}}\frac{1}{4}\lb(p_{nk}(x_{j})\rb)_{l}\lb(p_{nk}(x_{j})\rb)_{m}\lb(x_{j+1}-x_{j-1}\rb)^{2}\sigma_{F_{3}(x_{j})}^{2}.
1352: \end{eqnarray}
1353: Here, the index $j$ runs over the number of data points we have for
1354: $xF_{3}$ at a given $Q_{i}^{2}$. The $j=0$ and $j=N_{i}+1$ terms are
1355: simply the $x=0$ and $x=1$ endpoints (see parts {\bf I} and {\bf II}
1356: of Fig.~\ref{f:4fits}). From this equation we can calculate elements
1357: of the covariance matrices for each value of $Q^{2}$. All that
1358: remains is for us to invert them. However, upon attempting to do so,
1359: we find that these matrices are ill-conditioned, with some of their
1360: eigenvalues being close to zero. Hence their inverses are
1361: intractable. As a result of this, it is impossible to perform a
1362: reliable $\chi^{2}$ analysis of the averages with their correlation
1363: taken into account.
1364: 
1365: We believe that this is principally due to the fact that the
1366: correlation between averages is significant in some cases, and this
1367: in turn is an artefact of the fact that the selection criteria
1368: systematically select Bernstein polynomials which are peaked in the
1369: same region and hence are of fairly similar shape. This situation
1370: arises because the intent behind the inclusion of more averages in
1371: the analysis is not to increase the amount of `data', rather it is
1372: to further ensure that the missing data regions are suppressed.
1373: 
1374: In light of this, we settle for the method adopted in
1375: Refs.~\cite{r2,r3}, in which the na\"{i}ve $\chi^{2}$ function of
1376: Eq.~(\ref{eq:6:chi2}) is used, but the error bars on the averages
1377: are modified in order to account for the `over-counting of degrees
1378: of freedom'. For example, for the standard averages, at each value
1379: of $Q^{2}$ we have theoretical information on 9 moments, but the
1380: number of experimental Bernstein averages we use is often more than
1381: this; e.g.~for $Q^{2}=20\,\GeVV$ we use 27 standard Bernstein
1382: averages. To remedy this, we adopt the following approach. For each
1383: value of $Q^{2}$ we count the number of averages above 9 as
1384: duplicate information. The number of duplicates we have altogether
1385: is 73 and so for the values of $Q^{2}$ for which we have more
1386: Bernstein averages than moments, we rescale the error on these
1387: averages by $\sqrt{130/(130-73)}=1.510$. Correspondingly, for the
1388: modified Bernstein averages we rescale the errors by a factor of
1389: $\sqrt{141/(141-87)}=1.616$. This rescaling has the effect of
1390: suppressing the contribution of the duplicate data points to
1391: $\chi^{2}$, relative to those values of $Q^{2}$ for which we have
1392: fewer Bernstein averages than moments.
1393: 
1394: \subsubsection*{Positivity constraints}\vspace{-\parskip}
1395: 
1396: 
1397: The fact that $xF_{3}$ is a positive definite function, and that the
1398: moments are simply integrals over these functions multiplied by a
1399: single power of $x$, means that we can impose certain positivity
1400: constraints on the parameters $A_{n}$, as follows.
1401: 
1402: We construct the following matrices from the moments,
1403: \begin{eqnarray}
1404: \hat{\MM}\;=\;\lb(\begin{array}{cccc}
1405: \MM_{1}&\MM_{2}&\cdots\qq&\MM_{9}\\[5pt]
1406: \MM_{2}&\MM_{3}& &\\
1407: \MM_{3}& &\ddots\qq&\\
1408: \vdots& & &\\
1409: \MM_{9}& & &\MM_{17}\\
1410: \end{array}\rb),\label{eq:6:det1}
1411: \end{eqnarray}
1412: and
1413: \begin{eqnarray}
1414: \D\hat{\MM}\;=\;\lb(\begin{array}{cccc}
1415: \D\MM_{1}&\D\MM_{2}&\cdots\qq&\D\MM_{9}\\[5pt]
1416: \D\MM_{2}&\D\MM_{3}& &\\
1417: \D\MM_{3}& &\ddots\qq&\\
1418: \vdots& & &\\
1419: \D\MM_{9}& & &\D\MM_{17}\\
1420: \end{array}\rb),\label{eq:6:det2}
1421: \end{eqnarray}
1422: where $\MM_{n}=\MM(n;Q^{2})$ and $\D\MM_{n}=\MM_{n}-\MM_{n+1}$.
1423: 
1424: In order for $\MM(n;Q^{2})$ to be moments of positive definite
1425: functions (as the structure functions must be), the determinants of
1426: the above matrices, and of all their minors, must be positive, for
1427: all values of $Q^{2}$ \cite{r3}. Evaluating these determinants at
1428: fixed $Q^{2}$ will translate to conditions on the parameters
1429: $A_{n}$. We do not implement these constraints as part of the
1430: fitting procedure. Rather, we perform checks on the values of the
1431: fitting parameters resulting from the $\chi^{2}$ minimization to
1432: ensure that they obey the above constraints.
1433: 
1434: 
1435: However, we do impose positivity constraints on the moments
1436: themselves. As a result of the determinantal constraints described
1437: above, and from the general form of the moments given in
1438: Eq.~(\ref{eq:6moms}), we can infer that the following inequalities
1439: must be satisfied,
1440: \begin{eqnarray}
1441: \MM(n;Q^{2})&>&0,\label{eq:6:con1}\\
1442: \MM(n;Q^{2})&>&\MM(n+1;Q^{2}),\label{eq:6:con2}
1443: \end{eqnarray}
1444: for fixed $Q^{2}$. Furthermore, we can implement these constraints
1445: by defining our fitting parameters $A_{n}$ in terms of a new set of
1446: parameters and then minimizing $\chi^{2}$ with respect to these new
1447: parameters.
1448: 
1449: 
1450: 
1451: We begin by picking some value of $Q_{0}^{2}$ at which to implement
1452: the conditions. We then take the last moment used in the analysis
1453: ($n=20$) and rewrite the constraint in Eq.~(\ref{eq:6:con1}) as,
1454: \begin{eqnarray}
1455: \MM(20;Q_{0}^{2})&=&\lb(\hA_{20}\rb)^{2},\label{eq:6:con1a}
1456: \end{eqnarray}
1457: where $\hA_{20}$ is a real number. The constraints in Eq.~(\ref{eq:6:con2})
1458: can also be rewritten as,
1459: \begin{eqnarray}
1460:   \MM(n;Q_{0}^{2})&=&\MM(n+1;Q_{0}^{2})+\lb(\hA_{n}\rb)^{2},\label{eq:6:con2a}
1461: \end{eqnarray}
1462: for $1\leq n<20$, where $\hA_{n}$ are all real numbers. The LHSs of
1463: Eqs.~(\ref{eq:6:con1a}) and (\ref{eq:6:con2a}) are simply a fitting
1464: parameter times a number. For example, in the case of $n=2$ and
1465: $Q^{2}_{0}=12.6\;\GeVV$ we have,
1466: \begin{eqnarray}
1467:   \MM(2;12.6\;\GeVV)&=&0.3932\,A_{2}.
1468: \label{eq:6:momnum}
1469: \end{eqnarray}
1470: From this, and equivalent expressions for the rest of $A_{n}$, we
1471: can obtain an expression for each $A_{n}$ in terms of the parameters
1472: $\hA_{1}$ - $\hA_{20}$. This means that we can replace the
1473: parameters $A_{1}$ - $A_{20}$ with $\hA_{1}$ - $\hA_{20}$ in the
1474: $\chi^{2}$ function. By doing this and then minimizing with respect
1475: to the $\hA_{n}$ parameters, we can find a minimum in $\chi^{2}$ for
1476: which the constraints in Eqs.~(\ref{eq:6:con1}) and
1477: (\ref{eq:6:con2}) are automatically satisfied. In effect, the
1478: reparameterization embedded in Eqs.~(\ref{eq:6:con1a}) and
1479: (\ref{eq:6:con2a}) restricts the parameter space to exclude
1480: solutions for which the constraints are not satisfied.
1481: 
1482: To implement this reparameterization we must choose a value of
1483: $Q_{0}^{2}$ at which to impose the constraints, whereas in reality
1484: they must be satisfied for all $Q^{2}$. Because of this, we perform
1485: the analysis for several different values of $Q_{0}^{2}$ and check
1486: that the results remain stable.
1487: 
1488: \section{Results of fitting to the data}\vspace{-\parskip}
1489: \label{s:Res}
1490: 
1491: We focus principally on the results from the CORGI analysis in which
1492: both odd and even moments are included and in which we include
1493: target mass corrections. This analysis results in a prediction for
1494: the  QCD scale parameter of,
1495: \begin{eqnarray}
1496: \LMS^{(5)}&=&219.11_{-23.6}^{+22.1}\;{\rm MeV}.
1497: \end{eqnarray}
1498: The errors on this value can be broken down into four different
1499: sources,
1500: \begin{eqnarray}
1501: \LMS^{(5)}&=&219.11\;_{-16.57}^{+18.36}\;({\rm
1502: stat.})\quad_{-8.17}^{+8.36}\;({\rm
1503: sys.})\quad_{-13.74}^{+14.47}\;({\rm mod.})\quad\pm8.97\;({\rm
1504: HT})\;{\rm MeV}. \label{eq:6:errbrea}
1505: \end{eqnarray}
1506: We have used method {\bf I} to obtain experimental values of the
1507: averages. The deviation between the results obtained using methods
1508: {\bf I} and {\bf IV} is used to evaluate the modelling error since
1509: these are the two methods which exhibit the largest deviation.
1510: 
1511: This result for $\LMS$ corresponds to a value of the strong coupling
1512: constant (evaluated at the mass of the $Z$ particle) of,
1513: \begin{eqnarray}
1514:   \alpha_{s}(M_{Z})&=&0.1189_{-0.0019}^{+0.0019}
1515: \end{eqnarray}
1516: These values are in excellent agreement with the current global
1517: averages of $\LMS^{(5)}=207.2\;\rmGeV$ and $\alpha_{s}=0.118\pm.002$
1518: \cite{r16}. There is also good agreement with the result obtained
1519: from fits using the Jacobi polynomial method  \cite{r9C} which yield
1520: ${\alpha}_{s}(M_Z)={0.119}^{+0.004}_{-0.004}$. This result is based
1521: on fits using odd moments only and includes a contribution to the
1522: error from scale dependence. Whilst close to the global average, the value of $\alpha_s(M_Z)$ we obtain
1523: is significantly larger than the values found in a recent analysis
1524: of the $F_2(x,Q^2)$ structure function \cite{r16c}, or from fits of
1525: parton distibution functions where DIS and Drell-Yan data are
1526: combined \cite{r16d}, it is also larger than the value which
1527: minimizes the $\chi^2$ in global parton distribution function fits
1528: such as Ref.~\cite{r16e}. 
1529: 
1530: 
1531: The $\chi^{2}/$d.o.f.~for our CORGI result is as follows,
1532: \begin{eqnarray}
1533: \frac{\chi^{2}}{\rm d.o.f.}&=&\frac{20.37}{271-(20+1)}\nn\\
1534: &=&0.0815.\label{eq:6:chi}
1535: \end{eqnarray}
1536: Here `$271$' refers to the number of experimental Bernstein average
1537: points used in the fits. Although this value is an order of
1538: magnitude larger than the $\chi^{2}/$d.o.f. obtained in
1539: Ref.~\cite{r3}, it is still significantly smaller than one would
1540: expect, suggesting the errors on the Bernstein averages have been
1541: over estimated. However, as discussed previously, the $\chi^{2}$
1542: function we are using does not take into account the correlation
1543: between data points. Indeed, if correlation was taken into account,
1544: one might expect that a more reasonable value of $\chi^{2}$ would be
1545: obtained.
1546: 
1547: Furthermore, in the $\chi^{2}$ function we eventually used, the
1548: errors on the Bernstein averages were rescaled in order to take into
1549: account the `over-counting of degrees of freedom'. As a result, the
1550: number of Bernstein averages is not representative of the true
1551: number of degrees of freedom in this particular $\chi^{2}$ function.
1552: Indeed, the Bernstein averages in the plots in Fig.~\ref{fig:BAfits}
1553: can be constructed from just 58 different moments at different
1554: values of $Q^{2}$ (via Eq.~(\ref{eq:BAsMoms})). Similarly, the
1555: modified Bernstein averages in Fig.~\ref{fig:xBA fits} can be built
1556: from 53 different moments. Hence,
1557: \begin{eqnarray}
1558: \frac{\chi^{2}}{\rm d.o.f.}\;=\;\frac{20.37}{111-21}\;=\;0.226,
1559: \label{eq:6:newchi}
1560: \end{eqnarray}
1561: is more representative of the true value of $\chi^{2}/$d.o.f.~in
1562: this approach. This is a more acceptable value, however we stress
1563: that the {\it true} minimum in $\chi^{2}$ can only be determined by
1564: taking correlation fully into account.
1565: 
1566: In Fig.~\ref{fig:BAfits} we plot the CORGI predictions for the
1567: Bernstein averages (with TMCs included) fitted to the experimental
1568: values. Figure \ref{fig:xBA fits} shows equivalent plots for the
1569: modified averages.
1570: \begin{figure}
1571: \begin{center}
1572: \hspace*{4pt}\begin{tabular}{c c} \psfrag{Fb}{$F_{nk}$}
1573: \psfrag{F70}{\textcolor{magenta}{\small{$F_{70}$}}}
1574: \psfrag{F50}{\textcolor{blue}{\small{$F_{50}$}}}
1575: \psfrag{F30}{\textcolor{mygreen}{\small{$F_{30}$}}}
1576: \psfrag{F10}{\textcolor{red}{\small{$F_{10}$}}}
1577: \hspace{-.04\textwidth}
1578: \includegraphics[width=0.5\textwidth]{b1.eps}& \psfrag{Fb}{$F_{nk}$}
1579: \psfrag{F80}{\textcolor{myred}{\small{$F_{80}$}}}
1580: \psfrag{F60}{\textcolor{mygold}{\small{$F_{60}$}}}
1581: \psfrag{F40}{\textcolor{mycyan}{\small{$F_{40}$}}}
1582: \psfrag{F20}{\textcolor{magenta}{\small{$F_{20}$}}}
1583: \includegraphics[width=0.5\textwidth]{b2.eps}\\
1584: \footnotesize{$Q^{2}/\GeVV$}&\footnotesize{$Q^{2}/\GeVV$}\\\\
1585: \psfrag{Fb}{$F_{nk}$}
1586: \psfrag{F81}{\textcolor{myred}{\small{$F_{81}$}}}
1587: \psfrag{F71}{\textcolor{mygold}{\small{$F_{71}$}}}
1588: \psfrag{F61}{\textcolor{mycyan}{\small{$F_{61}$}}}
1589: \psfrag{F51}{\textcolor{magenta}{\small{$F_{51}$}}}
1590: \psfrag{F41}{\textcolor{blue}{\small{$F_{41}$}}}
1591: \psfrag{F31}{\textcolor{mygreen}{\small{$F_{31}$}}}
1592: \psfrag{F21}{\textcolor{red}{\small{$F_{21}$}}}
1593: \hspace{-.04\textwidth}\includegraphics[width=0.5\textwidth]{b3.eps}&
1594: \psfrag{Fb}{$F_{nk}$}
1595: \psfrag{F82}{\textcolor{myred}{\small{$F_{82}$}}}
1596: \psfrag{F72}{\textcolor{mygold}{\small{$F_{72}$}}}
1597: \psfrag{F62}{\textcolor{mycyan}{\small{$F_{62}$}}}
1598: \psfrag{F52}{\textcolor{magenta}{\small{$F_{52}$}}}
1599: \psfrag{F42}{\textcolor{blue}{\small{$F_{42}$}}}
1600: \psfrag{F32}{\textcolor{mygreen}{\small{$F_{32}$}}}
1601: \psfrag{F74}{\textcolor{red}{\small{$F_{74}$}}}
1602: \includegraphics[width=0.5\textwidth]{b4.eps}\\
1603: \footnotesize{$Q^{2}/\GeVV$}&\footnotesize{$Q^{2}/\GeVV$}\\\\
1604: \multicolumn{2}{c}{\psfrag{Fb}{$F_{nk}$}
1605: \psfrag{F83}{\textcolor{mycyan}{\small{$F_{83}$}}}
1606: \psfrag{F73}{\textcolor{magenta}{\small{$F_{73}$}}}
1607: \psfrag{F63}{\textcolor{blue}{\small{$F_{63}$}}}
1608: \psfrag{F84}{\textcolor{mygreen}{\small{$F_{84}$}}}
1609: \psfrag{F53}{\textcolor{red}{\small{$F_{53}$}}}
1610: \includegraphics[width=0.5\textwidth]{b5.eps}}\\
1611: \multicolumn{2}{c}{\footnotesize{$Q^{2}/\GeVV$}}
1612: \end{tabular}
1613: \end{center}
1614: \caption{CORGI fits for the Bernstein averages, with TMCs included.}
1615: \label{fig:BAfits}
1616: \end{figure}%INCORRECT
1617: 
1618: \begin{figure}[t]
1619: \begin{center}
1620: \begin{tabular}{c c} \psfrag{Fb}{$\tF_{nk}$}
1621: \psfrag{F90}{\textcolor{mygreen}{\small{$\tF_{80}$}}}
1622: \psfrag{F80}{\textcolor{myred}{\small{$\tF_{80}$}}}
1623: \psfrag{F70}{\textcolor{myred}{\small{$\tF_{70}$}}}
1624: \psfrag{F60}{\textcolor{mygold}{\small{$\tF_{60}$}}}
1625: \psfrag{F50}{\textcolor{mycyan}{\small{$\tF_{50}$}}}
1626: \psfrag{F40}{\textcolor{magenta}{\small{$\tF_{40}$}}}
1627: \psfrag{F30}{\textcolor{blue}{\small{$\tF_{30}$}}}
1628: \psfrag{F20}{\textcolor{mygreen}{\small{$\tF_{20}$}}}
1629: \psfrag{F10}{\textcolor{red}{\small{$\tF_{10}$}}}
1630: \hspace{-.04\textwidth}
1631: \includegraphics[width=0.5\textwidth]{mb1.eps}&
1632: \psfrag{Fb}{$\tF_{nk}$}
1633: \psfrag{F81}{\textcolor{myred}{\small{$\tF_{81}$}}}
1634: \psfrag{F71}{\textcolor{mygold}{\small{$\tF_{71}$}}}
1635: \psfrag{F61}{\textcolor{mycyan}{\small{$\tF_{61}$}}}
1636: \psfrag{F51}{\textcolor{magenta}{\small{$\tF_{51}$}}}
1637: \psfrag{F41}{\textcolor{blue}{\small{$\tF_{41}$}}}
1638: \psfrag{F31}{\textcolor{mygreen}{\small{$\tF_{31}$}}}
1639: \psfrag{F21}{\textcolor{red}{\small{$\tF_{21}$}}}
1640:  \includegraphics[width=0.5\textwidth]{mb2.eps}\\
1641: \footnotesize{$Q^{2}/\GeVV$}&\footnotesize{$Q^{2}/\GeVV$}
1642: \\\\
1643: \psfrag{Fb}{$\tF_{nk}$}
1644: \psfrag{F82}{\textcolor{myred}{\small{$\tF_{82}$}}}
1645: \psfrag{F72}{\textcolor{mygold}{\small{$\tF_{72}$}}}
1646: \psfrag{F62}{\textcolor{mycyan}{\small{$\tF_{62}$}}}
1647: \psfrag{F52}{\textcolor{magenta}{\small{$\tF_{52}$}}}
1648: \psfrag{F42}{\textcolor{blue}{\small{$\tF_{42}$}}}
1649: \psfrag{F84}{\textcolor{mygreen}{\small{$\tF_{84}$}}}
1650: \psfrag{F74}{\textcolor{red}{\small{$\tF_{74}$}}}
1651: \hspace{-.04\textwidth}  \includegraphics[width=0.5\textwidth]{mb3.eps}&
1652: \psfrag{Fb}{$\tF_{nk}$}
1653: \psfrag{F83}{\textcolor{magenta}{\small{$\tF_{83}$}}}
1654: \psfrag{F73}{\textcolor{blue}{\small{$\tF_{73}$}}}
1655: \psfrag{F63}{\textcolor{mygreen}{\small{$\tF_{63}$}}}
1656: \psfrag{F53}{\textcolor{red}{\small{$\tF_{53}$}}}
1657: \includegraphics[width=0.5\textwidth]{mb4.eps}\\
1658: \footnotesize{$Q^{2}/\GeVV$}&\footnotesize{$Q^{2}/\GeVV$}
1659: \end{tabular}
1660: \end{center}
1661: \caption{ CORGI fits for the  modified Bernstein averages, with TMCs
1662: included.} \label{fig:xBA fits}
1663: \end{figure}
1664: 
1665: 
1666: 
1667: \begin{table}
1668: \begin{center}
1669: \begin{tabular}{|c||c|c|c|}
1670: \hline&&&\\[-10pt] &$\LMS^{(5)}
1671: (\textrm{MeV})$&$\alpha_{s}(M_{Z})$&$\chi^{2}/{\rm
1672: d.o.f.}$\\[.1cm]\hhline{|=#=|=|=|}
1673: 
1674: 
1675: All moments&219.1\pmpm{+23.6}{-22.1}&
1676: 0.1189\pmpm{+0.0019}{-0.0019}&$20.37/(271-(20+1))$\\\hhline{-|---|}%done
1677: 
1678: 
1679: Odd moments&210.5\pmpm{+35.0}{-32.7}&
1680: 0.1182\pmpm{+0.0029}{-0.0030}&$10.94/(130-(10+1))$\\\hhline{-|---|}%done
1681: 
1682: 
1683: Even moments&229.5\pmpm{+64.7}{-62.1}&
1684: 0.1198\pmpm{+0.0048}{-0.0048}&$9.24/(141-(10+1))$\\\hhline{-|---|}%done
1685: 
1686: 
1687: All moments: $Q^{2}>m_{b}^{2}$ only&232.4\pmpm{+34.9}{-33.2}&
1688: 0.1200\pmpm{+0.0027}{-0.0027}&$15.59/(228-(20+1))$\\\hline%done
1689: \end{tabular}
1690: \end{center}
1691: \caption{In this table we present the result of the analysis
1692: performed using
1693:   the {\bf CORGI} approach to perturbation theory with target mass corrections
1694:   included. We compare the results obtained when we include
1695:   all moments (up to $n=20$) with those obtained when we restrict the
1696:   analysis to even or odd moments only. We also show the results from
1697:   performing the analysis with only data points for which
1698:   $Q^{2}>m_{b}^{2}$ ($\nf=5$) included.}
1699: \label{t:1}
1700: \end{table}
1701: In table \ref{t:1} we present the full set of results from the CORGI
1702: analysis. This table shows the results obtained when we include both
1703: odd and even moments (standard and modified Bernstein averages)
1704: together and also when we restrict the analysis to odd and even or
1705: odd moments only. These results allow us to check consistency
1706: between the odd, even and `All' analyses.
1707: 
1708: We can also use the results of the `odd moments' analysis to check
1709: consistency with previous analyses. In the PS analysis of
1710: Ref.~\cite{r3} (in which only the $n=1,3,5,7,9,11$ and 13 moments
1711: were included) a value of $\LMS^{(4)}=255\pm72$MeV was found,
1712: corresponding to $\LMS^{(5)}=178_{-55}^{+57}$MeV, with a value of
1713: $\chi^{2}/{\rm d.o.f.}=0.007$. Using the same set of moments, the
1714: CORGI analysis of Ref.~\cite{r4} found a value of
1715: $\LMS^{(5)}=228_{-36}^{+35}$MeV, although it must be noted that this
1716: analysis used incorrect values of the coefficients $X_{2}$, and
1717: therefore the result must be regarded as unreliable. Both of those
1718: results are indeed consistent with the `odd moments' analysis
1719: performed here.
1720: 
1721: 
1722: We also perform an analysis in which we restrict the CCFR data to
1723: $Q^{2}>m_{b}^{2}$ only, as a check on our method of evolving through
1724: the $b$ quark threshold. These results are also included in table
1725: \ref{t:1}. In table 3 we give the fitted CORGI values for the $A_n$
1726: non-perturbative coefficients for $n=1-20$, together with the values
1727: of the corresponding moments at $Q^2={8.75}\;{\rm{GeV}}^{2}$ and
1728: ${12.6}\;{\rm{GeV}}^{2}$.
1729: 
1730: 
1731: 
1732: 
1733: 
1734: 
1735: In table \ref{t:2} we compare the CORGI results with those obtained
1736: using the PS and EC approaches. We also present results obtained
1737: from performing these analyses with and without target mass
1738: corrections. The fact that the number of d.o.f. for the CORGI fits
1739: is $271$ ($272$), as opposed to $273$ for PS and EC, reflects the
1740: fact that for the smallest energy bin $Q^2={7.9}\;{\rm{GeV}}^2$, the
1741: ${\Lambda}_{\MM}^{2}$ appearing in the CORGI coupling exceeds $Q^2$
1742: for the highest $n=19,20$ moments, and hence one is below the Landau
1743: pole in the CORGI coupling of Eq.~(\ref{eq:6tHooft}).
1744: Correspondingly $x_{\rmt{CORGI}}$ is significantly less than unity
1745: (see table 1). We simply omit the two affected Bernstein average
1746: points from the CORGI fit.
1747: \begin{table}
1748: \begin{center}
1749: \begin{tabular}{|c|c|c|c|}\hline
1750: $n$&$A_{n}^{(4)}$&$\MM(n;8.75\,\GeVV)$&$\MM(n;12.6\,\GeVV)$\\
1751: \hline 1&2.346&2.494& 2.525$$\\\hline
1752:  2&0.8814& 0.3557&0.3466\\\hline
1753:  3 &0.4133&0.1002&9.545$\times 10^{-2}$\\\hline
1754:  4&0.2217&3.835$\times 10^{-2}$&3.584$\times 10^{-2}$\\\hline
1755:  5&0.1292&1.744$\times 10^{-2}$&1.603$\times 10^{-2}$\\\hline
1756:  6&8.134$\times 10^{-2}$&9.048$\times 10^{-3}$&8.191$\times 10^{-3}$\\\hline
1757:  7&5.241$\times 10^{-2}$&4.988$\times 10^{-3}$&4.452$\times 10^{-3}$\\\hline
1758:  8&3.639$\times 10^{-2}$&3.044$\times 10^{-3}$&2.681$\times 10^{-3}$\\\hline
1759:  9&2.434$\times 10^{-2}$&1.826$\times 10^{-3}$&1.588$\times 10^{-3}$\\\hline
1760:  10&1.822$\times 10^{-2}$&1.246$\times 10^{-3}$&1.07$\times 10^{-3}$\\\hline
1761:  11&1.202$\times 10^{-2}$&7.588$\times 10^{-4}$&6.438$\times 10^{-4}$\\\hline
1762:  12&9.64$\times 10^{-3}$&5.677$\times 10^{-4}$&4.76$\times 10^{-4}$\\\hline
1763:  13&5.935$\times 10^{-3}$&3.289$\times 10^{-4}$&2.726$\times 10^{-4}$\\\hline
1764:  14&5.119$\times 10^{-3}$&2.69$\times 10^{-4}$&2.204$\times 10^{-4}$\\\hline
1765:  15&2.702$\times 10^{-3}$&1.355$\times 10^{-4}$&1.098$\times 10^{-4}$\\\hline
1766:  16&2.535$\times 10^{-3}$&1.22$\times 10^{-4}$&9.768$\times 10^{-5}$\\\hline
1767:  17&9.362$\times 10^{-4}$&4.343$\times 10^{-5}$&3.44$\times 10^{-5}$\\\hline
1768:  18&9.739$\times 10^{-4}$&4.376$\times 10^{-5}$&3.426$\times 10^{-5}$\\\hline
1769:  19&9.807$\times 10^{-10}$&4.284$\times 10^{-11}$&3.317$\times 10^{-11}$\\\hline
1770:  20&7.69$\times 10^{-10}$&3.277$\times 10^{-11}$&2.509$\times 10^{-11}$\\\hline
1771: \end{tabular}
1772: \end{center}
1773: \caption{Fitted values of $A_n$ (in the $\nf=4$ region) together
1774: with the moments evaluated at
1775:   $Q^{2}=8.75$ and $12.6\,\GeVV$, for the CORGI approach.}
1776: \label{t:An}
1777: \end{table}
1778: \begin{table}
1779: \begin{center}
1780: \begin{tabular}{|c|c||c|c|c|}
1781: \hline \multicolumn{2}{|c||}{}
1782: &$\LMS^{(5)}(\textrm{MeV})$&$\alpha_{s}(M_{Z})$&$\chi^{2}/{\rm
1783: d.o.f.}$\\\hhline{|=|=#=|=|=|}
1784: 
1785: \multirow{2}{2cm}{CORGI} &with
1786: TMC&219.1\pmpm{+23.6}{-22.1}&0.1189\pmpm{+0.0019}{-0.0019}
1787: &$20.37/(271-(20+1))$\\\hhline{~|----|}%done
1788: 
1789: &no TMC &280.3\pmpmc{+24.6}{-23.4}&0.1235\pmpm{+0.0017}{-0.0017}
1790: &$24.76/(272-(18+1))$\\\hhline{|=|=#=|=|=|}%done
1791: 
1792: 
1793: \multirow{2}{2cm}{PS} &with TMC&
1794: 200.4\pmpmc{+25.8}{-24.8}&0.1173\pmpm{+0.0023}{-0.0022}
1795: &$21.73/(273-(20+1))$\\\hhline{|~|----|}%done
1796: 
1797: &no TMC &257.5\pmpmc{+27.8}{-27.6}&0.1219\pmpm{+0.0020}{-0.0020}
1798: &$25.20/(273-(18+1))$\\\hhline{|=|=#=|=|=|}%done
1799: 
1800: 
1801: \multirow{2}{2cm}{EC} &with TMC
1802: &204.5\pmpmc{+19.9}{-18.9}&0.1177\pmpm{+0.0017}{-0.0017}
1803: &$22.71/(273-(20+1))$\\\hhline{|~|----|}%done
1804: 
1805: &no TMC &261.5\pmpmc{+19.0}{-18.4}&0.1222\pmpm{+0.0014}{-0.0014}\
1806: &$26.04/(273-(18+1))$\\\hline%done
1807: \end{tabular}
1808: \end{center}
1809: \caption{In this table we compare the results of the analysis performed with
1810:   the three different approaches to perturbation theory described in section 3,
1811:   CORGI, PS and EC. We also show the results from these analyses  performed  with and without
1812:   target mass corrections.}
1813: \label{t:2}
1814: \end{table}
1815: 
1816: 
1817: 
1818: \section{Discussion and conclusions}\vspace{-\parskip}
1819: \label{s:6Conc} In this paper we have used three different
1820: approaches to perturbation theory to perform a phenomenological
1821: analysis of moments of $F_{3}$ using the method of Bernstein
1822: averages. The three approaches differ in how they deal with the FRS
1823: dependence.  In the CORGI approach, we allow the FRS invariant
1824: quantity $X_{0}(Q)$ to determine the relationship between $M$, $\mu$
1825: and $Q$ for each moment. In so doing, we automatically resum the
1826: subset of terms present in the full perturbative expansion which are
1827: RG-predictable at NNLO. In the physical scale approach we set
1828: $M=\mu=Q$ and adopt the \MSbar~scheme for the subtractions in the
1829: renormalization {\it and} factorization procedures. In the effective
1830: charge approach, we set $M=\mu$ and apply the CORGI approach to the
1831: resulting single-scale effective charge. We described how
1832: predictions are derived in these three approaches and corrected
1833: errors in the CORGI method which were present in Refs.~\cite{r4,r5}.
1834: 
1835: 
1836: We described how target mass and higher twist corrections affect
1837: these theoretical predictions and also how we evolve expressions for
1838: the moments through the $b$-quark threshold. We explained how the
1839: Bernstein averages method eliminates any potential dependence of the
1840: analysis on missing data regions in $x$ and $Q^{2}$, and we also
1841: described how this method is generalized to treat both odd and even
1842: moments. We described the fitting procedure used to extract the
1843: optimal values of the QCD scale parameter and how we can implement
1844: various constraints which ensure that the results of this fitting
1845: are consistent with the structure functions being positive definite
1846: functions. We also presented an alternative, and slightly easier
1847: method for deriving the FRS invariant quantities $X_{i}$.
1848: 
1849: The results of the CORGI analysis presented in table \ref{t:1} show
1850: excellent agreement with the current global average for the strong
1851: coupling evaluated at $Q^{2}=M_{Z}^{2}$ \cite{r16}, and are also in
1852: good agreement with fits based on the Jacobi polynomial approach
1853: \cite{r9C}. From this we conclude that CORGI perturbation theory
1854: performs well when applied to the analysis of moments. The analyses
1855: in which we include only odd or  even moments are consistent with
1856: each other and with the full (all moments) analysis. Furthermore, in
1857: the analysis in which we include all moments, the errors are greatly
1858: reduced. This improvement in the analysis is made possible by the
1859: availability of the full NNLO anomalous dimension calculation and
1860: represents significant improvement on previous analyses.
1861: 
1862: 
1863: Excluding data points for which $Q^{2}<m_{b}^{2}$ leads to no
1864: significant change in the results and from this we conclude that the
1865: quark mass threshold method we have applied is suitable to the
1866: moment analysis. The error associated with the exclusion of
1867: higher-twist effects, given in Eq.~(\ref{eq:6:errbrea}), is
1868: relatively small, signifying that these effects are not particularly
1869: important at scales $Q^{2}>7.6\,\GeVV$.
1870: 
1871: 
1872: We include in the analysis positivity constraints on the moments
1873: (Eqs.~(\ref{eq:6:con1}) and (\ref{eq:6:con2})), via the
1874: parameter redefinitions defined in section 4. We find that this
1875: implementation has little effect on the prediction of $\LMS$ ($\sim10$ Mev),
1876: but does make a difference to the values of $A_{n}$.
1877: 
1878: 
1879: 
1880: The CORGI predictions for the Bernstein averages (with TMCs
1881: included) are plotted in Figs.~\ref{fig:BAfits} and \ref{fig:xBA
1882: fits} and show excellent agreement with experimental values. This is
1883: reflected by the low value of $\chi^{2}/$d.o.f.~associated with this
1884: fitting, given in Eq.~(\ref{eq:6:chi}). However, as we noted, ideally
1885: the full covariance matrix should be used in constructing $\chi^2$ to
1886: account for correlations between the Bernstein averages used in the fits.
1887: Unfortunately, we found that this matrix is ill-conditioned, having some
1888: eigenvalues close to zero, and so it proved to be numerically intractable
1889: to invert the matrix to construct the true $\chi^2$. We therefore
1890: resorted to the same approximate rescaling of errors employed in
1891: Refs.\cite{r2,r3} to try to compensate for possible correlations.
1892: 
1893: The results also show consistency between CORGI, PS and EC. The PS
1894: and EC analyses lead to values of $\LMS$ and ${\alpha}_{s}$ slightly
1895: lower than in the CORGI analysis. However, this variation is well
1896: within the error bars on the associated quantities. Inclusion of HT
1897: effects generally results in a small shift in $\LMS$ of about 10
1898: MeV. However, when target mass corrections are included, we see a
1899: shift of approximately $60$ MeV in the predicted value of $\LMS$ and
1900: from this we conclude that these contributions are significant in
1901: the case of $F_{3}$.
1902: 
1903: 
1904: An obvious further study would be to apply the same fitting procedure to the recently
1905: released NuTev data \cite{r1a}. In future work we also hope to report on similar fits to
1906: data for the $F_2$ structure function \cite{r18}. This analysis is considerably more complicated due
1907: to the presence of an additional singlet component.
1908: 
1909: 
1910: %% \sectionm{Summary}
1911: %% \label{s:6Summ}
1912: \section*{Note Added in Proof}
1913: At around the same time as our paper was completed a related analysis
1914: of the CCFR data for $xF_3$ also employing Bernstein averages appeared
1915: \cite{r19}.
1916: 
1917: \section*{Acknowledgements}\vspace{-\parskip}
1918: We would like to thank Andreas Vogt for providing us with computer
1919: code implementing the NNLO anomalous dimension coefficients of
1920: Refs.~\cite{r7,r8}, and also for helpful discussions on quark mass
1921: thresholds. We thank Jeff Forshaw for pointing out the potential importance of taking
1922: correlations between the Bernstein averages into account. 
1923: Andrei Kataev is also thanked for useful discussions. P.M.B.
1924: gratefully acknowledges the receipt of a PPARC UK studentship.
1925: 
1926: 
1927: 
1928: 
1929: 
1930: \begin{thebibliography}{99}
1931: \bibitem{r1} W.Seligman et al. (CCFR Collaboration), Phys.Rev.Lett.{\bf 79} (1997) 1213; W.Seligman, Ph.D. thesis (Columbia University), Nevis Report 292.
1932: \bibitem{r1a} M. Tzanov et al. (NuTev Collaboration), Int. J. Mod. Phys. {\bf A20} (2005) 3759.
1933: \bibitem{r2} J.Santiago and F.J.Yndurain, Nucl. Phys. {\bf B563} (1999) 45. [hep-ph/9904344].
1934: \bibitem{r3} J.Santiago and F.J.Yndurain, Nucl. Phys. {\bf B611} (2001) 447. [hep-ph/0102247].
1935: \bibitem{r4} C.J. Maxwell and A. Mirjalili, Nucl. Phys. {\bf B645} (2002) 298. [hep-ph/0207069].
1936: \bibitem{r5} C.J. Maxwell [hep-ph/9908463];
1937: C.J. Maxwell and A. Mirjalili, Nucl. Phys. {\bf B577} (2000) 209.
1938: [hep-ph/0002204].
1939: \bibitem{r6} A. Retey and J.A.M. Vermaseren, Nucl. Phys. {\bf B604} (2001) 281. [hep-ph/0007294].
1940: \bibitem{r7} A. Vogt, S. Moch, J.A.M. Vermaseren, Nucl. Phys. {\bf B688} (2004) 101. [hep-ph/0403192].
1941: \bibitem{r8} A. Vogt, S. Moch, J.A.M. Vermaseren, Nucl. Phys. {\bf B691} (2004) 129. [hep-ph 0404111].
1942: \bibitem{r9} G. Grunberg, Phys. Rev. {\bf D29} (1984) 2315.
1943: \bibitem{r9A} A.L. Kataev, A.V. Kotikov, G. Parente and A.V. Sidorov, Phys. Lett. {\bf B417} (1998) 374. [hep-ph/9706534].
1944: \bibitem{r9B} A.L. Kataev, G. Parente and A.V. Sidorov, Nucl. Phys. {\bf B573} (2000) 405. [hep-ph/9905310].
1945: \bibitem{r9C} A.L. Kataev, G. Parente, A.V. Sidorov, Phys. Part. Nucl. {\bf 34} (2003) 20. [hep-ph/0106221].
1946: \bibitem{r9a} P.M. Stevenson, Phys. Rev. {\bf D23}, (1981) 2916.
1947: \bibitem{r10} H. David Politzer, Nucl. Phys. {\bf B194} (1982) 493.
1948: \bibitem{r11} P.M. Stevenson. H. David Politzer, Nucl. Phys. {\bf B277} (1986) 758.
1949: \bibitem{r11a}  G 't Hooft, in Deeper Pathways in High Energy Physics, proceedings of
1950: Orbis Scientiae, 1977, Coral Gables, Florida, edited by A. Perlmutter and L.F. Scott
1951: (Plenum, New York, 1977).
1952: \bibitem{r12}  E. Gardi, G. Grunberg and M. Karliner, JHEP {\bf 9807} (1998)
1953:   007, [hep-ph/9806462].
1954: \bibitem{r13}  M.A. Magradze, Int. J. Mod. Phys. {\bf A15},  (2000) 2715, [hep-ph/9911456].
1955: \bibitem{r14}  R.M.~Corless,~G.H.~Gonnet,~D.E.G~Hare,~D.J.~Jeffrey and D.E. Knuth, ``On~the
1956: ~Lambert~$W$~function'', Advances~in~Computational~Mathematics~{\bf 5}~(1996)~329,
1957: available from {\tt http://www.apmaths.uwo.ca/$\sim$djeffrey/offprints.html.}
1958: \bibitem{r14a} A.J. Buras, E.G. Floratos, D.A. Ross, C.T. Sachrajda, Nucl. Phys. {\bf B131} (1977) 308;
1959: W.A. Bardeen, A.J. Buras, D.W. Duke, T. Muta, Phys. Rev. {\bf D18} (1978) 3998.
1960: \bibitem{r15} O.~Nachtmann,  Nucl. Phys. {\bf B63} (1973) 237.
1961: \bibitem{r15a} H.~Georgi and H. David
1962:  Politzer, Phys. Rev.{\bf D14} (1976) 1829.
1963: \bibitem{r16}
1964: W.-M. Yao {\it et al.}  [Particle Data Group], J. Phys. {\bf G33} (2006) 1.
1965:   %``Review of particle physics,''
1966: \bibitem{r16a}  W. Bernreuther and W Wetzel, Nucl. Phys. \textbf{B197} (1982) 228; Erratum {\it ibid.} {\bf B513} (1998) 758.
1967: \bibitem{r16b} R.J. Barlow, Statistics: A guide to the use of statistical methods in the Physical Sciences, Manchester Physics Series (Wiley 1989).
1968: \bibitem{r16c} J. Bl\"umlein, Helmut B\"ottcher, and Alberto Guffanti, [hep-ph/0607200].  
1969: \bibitem{r16d} S. Alekhin, K. Melnikov and F. Petriello, Phys. Rev. {\bf D74} (2006) 054033. [hep-ph/0606237] 
1970: \bibitem{r16e} A.D. Martin, R.G. Roberts, W.J. Stirling and R.S. Thorne, Phys. Lett. {\bf B531} (2002) 216. [hep-ph/0201127]
1971: \bibitem{r17} K. Adel, F. Barreiro and F.J. Yndurain, Nucl. Phys. {\bf B495} (1997)  221.
1972: \bibitem{r18} P.M. Brooks and C.J. Maxwell, in preparation.
1973: \bibitem{r19} A. Khorramian and S. Atashbar Tehrani, JHEP03 (2007) 051.
1974: [hep-ph/0610136].
1975: \end{thebibliography}
1976: \end{document}
1977: