hep-ph0507078/qcd.tex
1: \section{HADRONIC TAU DECAYS AND QCD}	
2: \label{sec:qcd}
3: 
4: Tests of Quantum Chromodynamics and the precise measurement of the 
5: strong coupling constant \as at the $\tau$ mass 
6: scale~\cite{narisonpich:1988,braaten:1989,bnp,pichledib}, 
7: carried out for the first time by the ALEPH~\cite{aleph_as} and 
8: CLEO~\cite{cleo_as} collaborations have triggered many theoretical
9: developments. They concern primarily the perturbative expansion
10: for which different optimized rules have been suggested. Among these 
11: are contour-improved (resummed) fixed-order perturbation 
12: theory~\cite{pert,pivov1,pivov2}, effective charge and minimal 
13: sensitivity schemes~\cite{grunberg1,grunberg2,dhar1,dhar2,pms}, the
14: large-$\beta_0$ expansion~\cite{beneke,altarelli,neubert}, as well as
15: combinations of these approaches. They mainly distinguish themselves
16: in how they deal with the fact that the perturbative
17: series is truncated at an order where the missing part is not
18: expected to be small. Also, in the discussion of the so-called 
19: {\em contour improved} approach we will point out the importance 
20: to fully retain the complete information from the renormalization 
21: group when computing integrals of perturbatively expanded quantities.
22: With the publication of the full vector and axial-vector spectral 
23: functions by ALEPH~\cite{aleph_vsf,aleph_asf} and OPAL~\cite{opal_vasf}
24: it became possible to directly study the nonperturbative properties of 
25: QCD through vector minus axial-vector sum rules. These analyses are 
26: described in Section~\ref{sec:sumrules}. 
27: 
28: One could wonder how $\tau$ decays may at all allow us to learn something about 
29: perturbative QCD. The hadronic decay of the $\tau$ is dominated by resonant 
30: single particle final states. The corresponding QCD interactions that bind 
31: the quarks and gluons into these hadrons necessarily involve small momentum
32: transfers, which are outside the domain of perturbation theory. On the 
33: other hand, perturbative QCD describes interactions of quarks and gluons
34: with large momentum transfer. Indeed, it is the inclusive character of 
35: the sum of all hadronic $\tau$ decays that allows us to probe fundamental
36: short distance physics. Inclusive observables like
37: the total hadronic $\tau$ decay rate \Rtau~(\ref{eq:rtau}) can be accurately 
38: predicted as function of \asm using perturbative QCD, and including
39: small nonperturbative contributions within the framework of the Operator 
40: Product Expansion (OPE)~\cite{svz}. In effect, \Rtau is a doubly inclusive 
41: observable since it is the result of an integration over all hadronic final 
42: states at a given invariant mass and further over all masses between 
43: $m_\pi$ and $m_\tau$.
44: Hadronic $\tau$ decays are more inclusive than \ee annihilation data, 
45: since vector and axial-vector contributions are of approximately
46: equal size, while the isospin-zero component in \ee data is three 
47: times smaller than the isovector part, which is (about) equal to the 
48: vector contribution in $\tau$ decays.
49: The scale $m_\tau$ lies in a compromise region 
50: where \asm is large enough so that \Rtau is sensitive to its value, 
51: yet still small enough so that the perturbative expansion converges 
52: safely and nonperturbative power terms are small.
53: 
54: If strong and electroweak radiative corrections are neglected, the 
55: theoretical parton level prediction for $SU_C(N_C)$, $N_C=3$ reads
56: \beq
57: \label{eq:parton}
58:    R_\tau =
59:      N_C\left(|V_{ud}|^2 + |V_{us}|^2\right) = 3~,
60: \eeq
61: so that we can estimate a perturbative correction to this value
62: of approximately 21\% to obtain~(\ref{eq:rtau}), assuming other sources 
63: to be small. One realizes the increase in sensitivity to \as 
64: compared to the $Z$ hadronic width, where due to the 
65: three times smaller \asZ the perturbative QCD correction
66: reaches only about 4\%.
67: 
68: The nonstrange inclusive observable \Rtau can be theoretically separated 
69: into contributions from specific quark currents, namely vector ($V$) and 
70: axial-vector ($A$) $\ubar d$ and $\ubar s$ quark currents. It is therefore 
71: appropriate to decompose
72: \beq
73: \label{eq:rtausum}
74:    \Rtau = \RtauV + \RtauA + \RtauS~,
75: \eeq
76: where for the strange hadronic width $\RtauS$ vector and axial-vector
77: contributions are not separated so far due to the lack of the
78: corresponding experimental information for the Cabibbo-suppressed modes.
79: Parton-level and perturbative terms do not distinguish vector and
80: axial-vector currents (for massless partons). Thus the corresponding 
81: predictions become 
82: $R_{\tau,V/A} = (N_C/2)|V_{ud}|^2$ and $\RtauS = N_C|V_{us}|^2$,
83: which add up to Eq.~(\ref{eq:parton}).
84: 
85: A crucial issue of the QCD analysis at the $\tau$ mass scale concerns the 
86: reliability of the theoretical description, \ie, the use of the OPE to 
87: organize the perturbative and nonperturbative expansions, and the control
88: of unknown higher-order terms in these series. A reasonable stability test 
89: is to continuously vary $m_\tau$ to lower values $\sqrt{s_0}\le m_\tau$ for 
90: both theoretical prediction and measurement, which is possible since the 
91: shape of the full $\tau$ spectral function is available. The kinematic 
92: factor that takes into account the $\tau$ phase space suppression 
93: at masses near to $m_\tau$ is correspondingly modified so that 
94: $\sqrt{s_0}$ represents the new mass of the $\tau$. 
95: 
96: %In the following subsections we proceed with a fairly detailed discussion
97: %of the various approaches to the perturbative prediction of $\Rtau$. 
98: %Improvements over the traditional fixed-order perturbative expansion 
99: %have been developed in the past
100: 
101: %
102: % ----------------------------- Tools -------------------------------
103: %
104: \subsection{Renormalization group equations}
105: \label{sec:qcd_RGEs}
106: 
107: Like in QED, the subtraction of divergences in QCD is equivalent to the 
108: renormalization of the coupling ($\as\equiv g_s^2/4\pi$), the 
109: quark masses ($m_q$), etc, and the fields in the bare  (superscript $B$) 
110: Lagrangian such as $\as^B=s^{\varepsilon}Z_{\as}\as$, 
111: $m_q^B=s^{\varepsilon}Z_mm_q$, etc.
112: Here $s$ is the renormalization scale, $\varepsilon$ the 
113: dimensional regularization parameter, and $Z$ denotes a series of 
114: renormalization constants obtained from the generating functional 
115: of the bare Green's function.
116: The renormalization procedure introduces an energy scale, $s$, 
117: which represents the point at which the subtraction to remove 
118: the divergences is actually performed.
119: This leads to the differential {\em Renormalization Group Equations} (RGE)
120: \beqn
121: \label{eq:betafun}
122: 	\frac{d a_s}{d \ln s}
123: 	&=& 
124: 	\beta(a_s) = -a_s^2\sum_n \beta_n a_s^n~, \\
125: \label{eq:gammafun}
126:    	\frac{1}{m_q}\frac{d m_q}{d \ln s}
127:      	&=& 
128:    	\gamma(\as)= -a_s\sum_n \gamma_n a_s^n~,
129: \eeqn
130: with $a_s=\as/\pi$. Expressed in the {\em modified minimal subtraction 
131: renormalization scheme} (\MSbar)~\cite{msbar1,msbar2} and for three 
132: active quark flavors at the $\tau$ mass scale, the perturbative coefficients, 
133: known to four loops~\cite{betafourloop,gammafourloop}, are\footnote
134: {
135: 	The full expressions for an arbitrary number of quark flavors ($n_f$)
136: 	are~\cite{betafourloop,gammafourloop}
137: 	\newcommand\hshere{\hspace{0.5cm}}
138: 	\beqns
139: 	   \beta_0 = \frac{1}{4}\left(11 - \frac{2}{3}n_f\right)~,  \hshere
140: 	   \beta_1 = \frac{1}{16}\left(102 - \frac{38}{3}n_f\right)~,  \hshere
141: 	   \beta_2 = \frac{1}{64}\left(\frac{2857}{2} - \frac{5033}{18}n_f 
142:                                   + \frac{325}{54}n_f^2\right)~, \\
143: 	   \beta_3 = \frac{1}{256}\left[
144: 	       \frac{149753}{6} + 3564\,\zeta_3 -
145: 	       \left(\frac{1078361}{162} + \frac{6508}{27}\,\zeta_3\right)n_f 
146: 	       + \left(\frac{50065}{162} + \frac{6472}{81}\,\zeta_3\right)n_f^2 +
147: 	       \frac{1093}{729}n_f^3\right]~,
148: 	\eeqns
149: 	and
150: 	\beqns
151: 	   \gamma_0 = 1~, \hshere
152: 	   \gamma_1 = \frac{1}{16}\left(\frac{202}{3} - \frac{20}{9}n_f\right)~,  \hshere
153: 	   \gamma_2 = \frac{1}{64}\left[1249 - 
154:                                   \left(\frac{2216}{27}+\frac{160}{3}\,\zeta_3\right)n_f 
155:                                   - \frac{140}{81}n_f^2\right]~, \\
156: 	   \gamma_3 = \frac{1}{256}
157: 	       \left[\frac{4603055}{162} + \frac{135680}{27}\,\zeta_3 - 8800\,\zeta_5
158: 	       + \left(-\frac{91723}{27} - \frac{34192}{9}\,\zeta_3 + 880\,\zeta_4 
159:                + \frac{18400}{9}\,\zeta_5\right)n_f\right.
160: 	        \\
161: 		\left.
162: 	       + \left(\frac{5242}{243} + \frac{800}{9}\,\zeta_3 
163: 			- \frac{160}{3}\,\zeta_4\right)n_f^2 +
164: 	       \left(-\frac{332}{243} + \frac{64}{27}\,\zeta_3\right)n_f^3\right]~,
165: 	\eeqns
166: 	where the $\zeta_{i=\{3,4,5\}}=\{1.2020569,\pi^4/90,1.0369278\}$ are the Riemann 
167: 	$\zeta$-functions.
168: }
169: \newcommand\hshere{\hspace{0.5cm}}
170: \beqn
171: \label{eq:beta}
172:  	\beta_0=2.25~, 			\hshere
173: 	\beta_1=4~,   			\hshere
174: 	\beta_2\approx 10.05989~,	\hshere 
175: 	\beta_3\approx 47.22804, \\
176: \label{eq:gamma}
177:  	\gamma_0=1~, 			\hshere 
178: 	\gamma_1=3.79166~, 		\hshere 
179: 	\gamma_2\approx 12.42018~, 	\hshere 
180: 	\gamma_3\approx 44.26278~.
181: \eeqn
182: The solutions for the evolutions~(\ref{eq:betafun}) and (\ref{eq:gammafun}) 
183: are obtained from integration\footnote
184: {
185: 	Note that the integration is only defined for a non-vanishing
186: 	derivative $\beta(a_s)\ne0$. Vanishing $\beta(a_s)$ may indeed 
187: 	occur for example in effective charge schemes with large $K_4$ 
188: 	or $K_5$ coefficients (see later in this section).
189: %	An iterative solution of the four-loop $\beta$ function has been
190: %	computed in~\cite{chet1}, which we give here for completeness
191: %	\beqns
192: %		a_s(s) &=&
193: %		\frac{1}{\beta_0 L} - \frac{b_1\ln L}{(\beta_0 L)^2}
194: %		+ \frac{1}{(\beta_0 L)^3}\left[b_1^2\left(\ln^2L-\ln L -1\right) + b_2\right]
195: %		\nonumber\\
196: %		&&
197: %		+\: \frac{1}{(\beta_0 L)^4}\left[b_1^3\left(-\ln^3L+\frac{5}{2}\ln^2L
198: %				+ 2\ln L - \frac{1}{2}\right)
199: %			- 3b_1 b_2\ln L + \frac{b_3}{2}\right]~,
200: %	\eeqns
201: %	where $L=\ln(s/\Lambda^2)$, $b_i=\beta_i/\beta_0$ and
202: %	terms of order $\mathcal{O}(L^{-5})$ have been neglected.
203: }~\cite{pms,chet1} 
204: \beqn
205: \label{eq:betafunint}
206: 	\ln\frac{s}{\Lambda^2} &=& 
207: 	\intl_0^{a_s(s)} da_s^\prime \left(\frac{1}{\beta(a_s^\prime)}
208: 		     + \frac{1}{\beta_0(a_s^\prime)^2+\beta_1(a_s^\prime)^3}\right)
209:          +\intl_{a_s(s)}^\infty da_s^\prime 
210: 		\frac{1}{\beta_0(a_s^\prime)^2+\beta_1(a_s^\prime)^3}~, \\
211: %	&\approx&
212: %	\frac{1}{\beta_0}\left[\frac{1}{a_s} + \frac{\beta_1}{\beta_0}\ln a_s
213: %	+ \left(\frac{\beta_2}{\beta_0}-\frac{\beta_1^2}{\beta_0^2}\right)a_s
214: %	+\left(\frac{\beta_3}{2\beta_0} - \beta_1\beta_2 
215: %		+ \frac{\beta_1^3}{2\beta_0^3}\right)a_s^2\right]
216: %	+ C~, \\
217: \label{eq:gammafunint}
218: 	\ln\frac{m_q(s)}{m_q(s_0)} &=& \intl^{a_s(s)}_{a_s(s_0)}
219: 			           d a_s^\prime \frac{\gamma(a_s^\prime)}{\beta(a_s^\prime)}~,
220: \eeqn
221: where the \MSbar asymptotic scale parameter $\Lambda$ and the RG-invariant
222: quark mass $\hat m_q$ (which appears after evaluating the 
223: integral~(\ref{eq:gammafunint})) are fixed by referring to known values 
224: $a_s(s_0)$ and $m_q(s_0)$, respectively, at scale $s_0$. 
225: % The explicit 
226: % solution of the integral~(\ref{eq:betafunint}) implies the integration 
227: % constant $C=(\beta_1/\beta_0^2)\ln\beta_0$, which is chosen to suppress 
228: % terms $\propto \ln(s/\Lambda^2)$~\cite{msbar2,furmanski} (the reader is 
229: % referred to~\cite{chet1} for a discussion of other suitable choices).
230: 
231: For most practical purposes, it is however unnecessary to explicitly
232: perform the integrations~(\ref{eq:betafunint}, \ref{eq:gammafunint}), since 
233: the evolution of the differential RGEs can be solved numerically by means 
234: of {\em single-step integration} using a modified Euler method. 
235: It consists of Taylor-developing, say, Eq.~(\ref{eq:betafun}) around the 
236: reference scale $s_0$ in powers of $\eta\equiv\ln(s/s_0)$, and reordering 
237: the perturbative series in powers of $a_s\equiv a_s(s_0)$,
238: \beqn
239: \label{eq:astaylor}
240: 	a_s(s) &=& 
241:   		a_s - 
242: 		\beta_0 \eta a_s^2 + 
243: 		\left(-\beta_1\eta + \beta_0^2 \eta^2\right) a_s^3 +
244:     		\left(-\beta_2\eta  + \frac{5}{2} \beta_0\beta_1 \eta^2 - \beta_0^3 \eta^3
245: 				\right) a_s^4 
246: 		\nonumber \\
247: 		&&   +\: 
248:     		\left(-\beta_3\eta  + \frac{3}{2} \beta_1^2\eta^2  + 3 \beta_0 \beta_2\eta^2  
249: 						- \frac{13}{3} \beta_0^2 \beta_1 \eta^3
250:                 				+ \beta_0^4 \eta^4
251: 				\right) a_s^5 
252: 		\\
253: 		&&   +\: 
254:     		\left(-\beta_4\eta  + \frac{7}{2} \beta_1 \beta_2\eta^2 
255: 						+ \frac{7}{2} \beta_0 \beta_3\eta^2  
256: 						- \frac{35}{6} \beta_0 \beta_1^2 \eta^3 
257:                 				- 6 \beta_0^2 \beta_2 \eta^3
258: 						+ \frac{77}{12} \beta_0^3 \beta_1 \eta^4
259: 						- \beta_0^5 \eta^5
260: 				\right)a_s^6 
261: 		\;+\;\mathcal{O}(a_s^7)~. \nonumber
262: \eeqn
263: For later use, we have
264: expressed the expansion up to order $a_s^6$, which involves the 
265: unknown five-loop coefficient $\beta_4$. An efficient integration 
266: of the RGEs can also be obtained using the CERNLIB routine
267: ``RKSTP'', which performs a $4^{\rm th}$ order Runge-Kutta single 
268: step approximation with excellent accuracy.
269: 
270: \subsection{Theoretical prediction of $\Rtau$}\label{sec:rtau_th}
271: 
272: According to Eq.~(\ref{eq:imv}) the absorptive parts of the vector 
273: and axial-vector two-point correlation functions 
274: $\Pi^{(J)}_{\ubar d,V/A}(s)$, with the spin $J$ of the hadronic 
275: system, are proportional to the $\tau$ hadronic \sfs\  with 
276: corresponding quantum numbers. The nonstrange ratio \RtauVpA
277: can be written as an integral of these \sfs\  over the 
278: invariant mass-squared $s$ of the final state hadrons~\cite{bnp}
279: \beq
280: \label{eq:rtauth1}
281:    \RtauVpA(s_0) =
282: 	12\pi \Sew\intl_0^{s_0}
283: 		\frac{ds}{s_0}\left(1-\frac{s}{s_0}
284:                                     \right)^{\!\!2}
285:      \left[\left(1+2\frac{s}{s_0}\right){\rm Im}\Pi^{(1)}(s+i\e)
286:       \,+\,{\rm Im}\Pi^{(0)}(s+i\e)\right]~,
287: \eeq
288: where $\Pi^{(J)}$ can be decomposed as
289: $\Pi^{(J)}=|V_{ud}|^2\left(\Pi_{ud,V}^{(J)}+\Pi_{ud,A}^{(J)}\right)$.
290: The lower integration limit is zero because the pion pole is at 
291: zero mass in the chiral limit.
292: 
293: The correlation function $\Pi^{(J)}$ is analytic in the complex $s$ plane 
294: everywhere except on the positive real axis where singularities exist.
295: Hence by Cauchy's theorem, the imaginary part of $\Pi^{(J)}$ is 
296: proportional to the discontinuity across the positive real axis
297: \begin{figure}[t]  
298:   \centerline{\epsfysize6.5cm\epsffile{figures/contour_zz.eps}}
299:   \caption[.]{\label{fig:contour}
300:               Integration contour for the r.h.s. in Eq.~(\ref{eq:contour}).}
301: \end{figure} 
302: \beq
303: \label{eq:contour}
304:    \frac{1}{\pi}\intl_0^{s_0}ds\,w(s){\rm Im}\Pi(s) =
305:    -\frac{1}{2\pi i}\hm\ointl_{|s|=s_0}\hm\hm\hm ds\,w(s)\Pi(s)~,
306: \eeq
307: where $w(s)$ is an arbitrary analytic function, and the contour 
308: integral runs counter-clockwise around the circle from $s=s_0+i\e$ to 
309: $s=s_0-i\e$ as indicated in Fig.~\ref{fig:contour}. 
310: 
311: The energy scale $s_0= m_\tau^2$ is large enough that contributions 
312: from nonperturbative effects are expected to be subdominant and the use 
313: of the OPE is appropriate. The kinematic factor $(1-s/s_0)^2$ suppresses 
314: the contribution from the region near the positive real axis where 
315: $\Pi^{(J)}(s)$ has a branch cut and the OPE validity is restricted 
316: due to large possible quark-hadron duality 
317: violations~\cite{quinnetal,braaten88}. 
318: 
319: The theoretical prediction of the vector and axial-vector
320: ratio \RtauVA can hence be written as
321: \beq
322: \label{eq:delta}
323:    \RtauVA =
324:      \frac{3}{2}|V_{ud}|^2\Sew\left(1 + \delta^{(0)} + 
325:      \delta^\prime_{\rm EW} + \delta^{(2,m_q)}_{ud,V/A} + 
326:      \hm\hm\sum_{D=4,6,\dots}\hm\hm\hm\hm\delta_{ud,V/A}^{(D)}\right)~,
327: \eeq
328: with the massless perturbative contribution $\delta^{(0)}$,
329: the residual non-logarithmic electroweak correction 
330: $\delta^\prime_{\rm EW}=0.0010$~\cite{braaten} (\cf\  the discussion 
331: on radiative corrections in Section~\ref{sec:cvc_isobreak}), and the 
332: dimension $D=2$ {\em perturbative} contribution $\delta^{(2,m_q)}_{ud,V/A}$ 
333: from quark masses, which is lower than $0.1\%$ for $u,d$ quarks. 
334: The term $\delta^{(D)}$ denotes the OPE contributions of mass
335: dimension $D$
336: \beq
337: \label{eq:ope}
338:     \delta_{ud,V/A}^{(D)} =
339:        \hm\hm\hm\sum_{{\rm dim}{\cal O}=D}\hm\hm\hm C_{ud,V/A}(s,\mu)
340:             \frac{\langle{\cal O}_{ud}(\mu)\rangle_{V/A}}
341:                  {(-\sqrt{s_0})^{D}}~,
342: \eeq
343: where the scale parameter $\mu$ separates the long-distance 
344: nonperturbative effects, absorbed into the vacuum expectation 
345: elements $\langle{\cal O}_{ud}(\mu)\rangle$, from the short-distance 
346: effects that are included in the Wilson coefficients 
347: $C_{ud,V/A}(s,\mu)$~\cite{wilson}. Note that 
348: $\delta_{ud,V+A}^{(D)}=(\delta_{ud,V}^{(D)}+\delta_{ud,A}^{(D)})/2$.
349: 
350: \subsubsection{The Perturbative Prediction}\label{sec:pert}
351: 
352: The perturbative prediction used by the experiments follows the
353: work of~\cite{pert}. The perturbative contribution is 
354: given in the chiral limit. Effects from quark masses have been 
355: calculated in~\cite{pertmass} and are found to be well below
356: 1\% for the light quarks. As a consequence, the contributions from 
357: vector and axial-vector currents coincide to any given order of 
358: perturbation theory and the results are flavor independent.
359: 
360: For the evaluation of the perturbative series, it is convenient to 
361: introduce the analytic Adler function~\cite{adler}
362: \beq
363: \label{eq:adler}
364:    D(s) \;\equiv\;
365:       - s\frac{d\Pi(s)}{ds} = 
366:             \frac{s}{\pi}\intl_0^\infty ds^\prime\,
367:             \frac{{\rm Im}\Pi(s^\prime)}{(s^\prime-s)^2}~,
368: \eeq
369: where the second identity is the dispersion relation (\cf\  Section~\ref{anomaly}). 
370: The derivative of the correlator avoids extra subtractions (renormalization) 
371: on the r.h.s. of Eq.~(\ref{eq:adler}), which are unrelated to QCD dynamics.
372: The function $D(s)$ calculated in perturbative QCD within the \MSbar 
373: renormalization scheme depends on a non-physical parameter $\mu$ 
374: occurring as $\ln(\mu^2/s)$. Furthermore it is a function of \as. 
375: On the other hand, since $D(s)$ is connected to a physical 
376: quantity, the \sf\  $\Im\Pi(s)$, it cannot depend on the subjective 
377: choice of $\mu$. This can be achieved if \as\  becomes a function of 
378: $\mu$ providing independence of $D(s)$ of the choice of $\mu$. 
379: Nevertheless, in the realistic case of a truncated series, some $\mu$ 
380: dependence remains and represents an irreducible systematic uncertainty. 
381: $D(s$) is then a function of $\mu^2/s$, $\as$ 
382: and can be understood as an effective charge (see below) that obeys the 
383: RGE. Choosing $\mu^2=s$, the perturbative series reads
384: \beq
385: \label{eq:pertAdler}
386:    D(\as(s)) = \hm\hm\hm
387:       \sum_{n=0}^{\infty}r_n a_s^{n}(s)~,
388: \eeq
389: with renormalization scheme dependent coefficients $r_n$.
390: 
391: To introduce the Adler function in Eqs.~(\ref{eq:rtauth1}) and 
392: (\ref{eq:contour}) one uses partial integration
393: \beqn
394: \label{eq:identity}
395:    \ointl_{|s|=s_0}\hm\hm\hm ds\,g(s)\Pi(s) 
396: 	&=&
397:      -\hm\hm\hm\ointl_{|s|=s_0}\hm\hm
398:      \frac{ds}{s}\left(G(s) - G(s_0)
399:                  \right)\,s\frac{d\Pi(s)}{ds}~, 
400: \eeqn
401: with the kernel $g(s)=(1-s/s_0)^2(1+2s/s_0)/s_0$ and
402: $G(s)=\int_0^s ds^\prime\,g(s^\prime)$. This gives~\cite{pert}
403: \beq
404: \label{eq:rtauadler}
405:    1+\delta^{(0)} = 
406:       -2\pi i\hm\hm\hm\ointl_{|s|=s_0}\hm\hm\frac{ds}{s}
407:        \left[1-2\frac{s}{s_0} + 2\left(\frac{s}{s_0}\right)^{\!\!3}
408:              - \left(\frac{s}{s_0}\right)^{\!\!4}\right] D(s)~.
409: \eeq
410: The perturbative expansion of the Adler function can be inferred 
411: from the third-order calculation of the \ee  inclusive cross section ratio 
412: $R_{\ee}(s)=\sigma(\ee\to{\rm hadrons}\,(\gamma))/\sigma(\ee\to\mu^+\mu^-\,(\gamma))$~\cite{adler1,adler2,2loop,loopbis,loopbisbis,pert}
413: \beq
414: \label{eq:adlerpert}
415:    D(s) =
416:      \frac{1}{4\pi^2}\sum_{n=0}^\infty \tilde{K}_n(\xi)a_s^{n}(-\xi s)~,
417: \eeq
418: with the $\tilde{K}_n(\xi)$ functions up to order $n=5$~\cite{pert}
419: \beqn
420: 	&&
421:    \tilde{K}_0(\xi) = K_0~, 	\hspace{1cm}
422:    \tilde{K}_1(\xi) = K_1~, 	\hspace{1cm}
423:    \tilde{K}_2(\xi) = K_2 + L~, \nonumber\\
424: 	&&
425:    \tilde{K}_3(\xi) = K_3 + \left(b_1 + 2\,K_2\right)L + L^2 \nonumber\\
426: 	&&
427: \label{eq:kfun}
428:    \tilde{K}_4(\xi) = K_4 + \left(b_2 + 2\,b_1 K_2 + 3\,K_3\right)L
429:  		            + \left(\frac{5}{2}\,b_1 + 3\,K_2\right)L^2 + L^3~, \\
430: 	&&
431:    \tilde{K}_5(\xi) = K_5 + \left(b_3 + 2\,b_2 K_2 + 3\,b_1 K_3 
432:                                     + 4\, K_4\right)L
433:                         + \left(\frac{3}{2}\,b_1^2 + 3\,b_2 + 7\,b_1 K_2 
434:                                 + 6\, K_3\right)L \nonumber\\
435: 	&& \hspace{1.7cm}
436:  		        +\: \left(\frac{13}{3}\,b_1 + 4\,K_2\right)L^3 + L^4~, 
437: 	\nonumber
438: \eeqn
439: where $b_i=\beta_i/\beta_0$ and $L=\beta_0\ln\xi$. The factor $\xi=s_0/s$ in 
440: Eq.~(\ref{eq:adlerpert}) represents the arbitrary renormalization 
441: scale ambiguity of which the Adler function~(\ref{eq:adler}) is independent. 
442: The $\xi$ dependence of the $K_i(\xi)$ coefficients is 
443: determined by inserting the expansion~(\ref{eq:astaylor}) into 
444: (\ref{eq:adlerpert}), and equating the two series obtained for $\xi=1$ 
445: and arbitrary $\xi$ at each order in $a_s$.\footnote
446: {
447: 	Note that this procedure corresponds to a fixed-order
448: 	perturbation theory rule since known higher order
449: 	pieces of the expansion~(\ref{eq:astaylor}) are neglected
450: 	when solving 
451: 	\beqns
452: 	\frac{\partial}{\partial\xi}\sum_{n=0}^N \tilde{K}_n(\xi)a_s^{n}(-\xi s)
453: 	=\frac{\beta(a_s^{n}(-\xi s))}{\xi}
454: 		\sum_{n=0}^N (n+1)\tilde{K}_n(\xi)a_s^{n}(-\xi s)
455: 	 +      \sum_{n=0}^N a_s^{n+1}(-\xi s)\frac{\partial \tilde{K}_n(\xi)}{\partial\xi}
456: 	 \sim \mathcal{O}(a_s^{N+1})~,
457: 	\eeqns
458: 	for a truncated series with maximum known order $N$. 
459: 	A resummation of the subleading logarithms has been proposed
460: 	in~\cite{kleiss}.
461: } The 
462: coefficients $K_n$ are known up to third order in $\as^3$. 
463: For $n\ge2$ they depend on the renormalization scheme used, while 
464: the first two coefficients are universal
465: \beq
466: \label{eq:kf1}
467:     K_0 = 1~,	\hspace{0.5cm}
468:     K_1 = 1~,	\hspace{0.5cm}
469:     K_2(\MSbar) = F_3(\MSbar)~, \hspace{0.5cm}
470:     K_3(\MSbar) = F_4(\MSbar) 
471:       \,+\, \frac{1}{3}\pi^2\beta_0^2~,
472: \eeq
473: where $F_3(\MSbar) = 1.9857 - 0.1153\,n_f$ and
474: $F_4(\MSbar) = -6.6368 - 1.2001\,n_f - 0.0052\,n_f^2$ are the 
475: coefficients of the perturbative series of $R_{\ee}$. For 
476: $n_f=3$ one has $K_2=1.640$ and $K_3=6.371$. With the
477: series~(\ref{eq:adlerpert}), inserted in the r.h.s of 
478: Eq.~(\ref{eq:rtauadler}), one obtains the perturbative expansion
479: \beq 
480: \label{eq:knan}
481:    \delta^{(0)} = 
482:        \sum_{n=1}^5 \tilde{K}_n(\xi) A^{(n)}(a_s)~,
483: \eeq
484: with the functions
485: \beqn
486: \label{eq:an}
487:    A^{(n)}(a_s) 
488: 	&=&
489:       \frac{1}{2\pi i}\hm\ointl_{|s|=s_0}\hm\hm
490:       \frac{ds}{s}
491:        \left[1-2\frac{s}{s_0} + 2\left(\frac{s}{s_0}\right)^{\!\!3}
492:              - \left(\frac{s}{s_0}\right)^{\!\!4}
493:        \right]a_s^{n}(-\xi s) \nonumber\\
494: 	&=&
495:   	\frac{1}{2\pi} \intl_{-\pi}^{\pi} d\varphi
496:        	\left[1 + 2e^{i\varphi} - 2e^{i3\varphi} - e^{i4\varphi}
497:        	\right] a_s^{n}(\xi s_0e^{i\varphi})~,
498: \eeqn
499: where $s=-s_0e^{i\varphi}$ has been substituted in the second line and
500: the integral path proceeds according to Fig.~\ref{fig:contour}. 
501: %Formally, 
502: %the integrals~(\ref{eq:an}) also obey an RGE~\cite{pert}
503: %\beq
504: %	\xi\frac{\partial A^{(n)}\left(a_s(-\xi s)\right)}
505: %                {\partial\xi} =
506: %               n\sum_{k=1}\beta_kA^{(n+k)}\left(a_s(-\xi s)\right)~.
507: %\eeq
508: 
509: \subsubsection{Fixed-order perturbation theory (FOPT)}
510: 
511: Inserting the series~(\ref{eq:astaylor}) in Eq.~(\ref{eq:knan}), 
512: evaluating the contour integral, and collecting the terms with 
513: equal powers in $a_s$ leads to the familiar expression~\cite{pert}
514: \beq 
515: \label{eq:kngn}
516:    \delta^{(0)} = 
517:        \sum_{n} \left[\tilde{K}_n(\xi) + g_n(\xi)\right]
518:        \left(\frac{\as(\xi s_0)}{\pi}\right)^{\!\!n}~,
519: \eeq
520: where the $g_n$ are functions of $\tilde{K}_{m<n}$ and $\beta_{m<n-1}$, and of
521: elementary integrals with logarithms of power $m<n$ in the integrand. Setting 
522: $\xi=1$ and replacing all known $\beta_i$ and $K_i$ coefficients
523: by their numerical values, Eq.~(\ref{eq:kngn}) simplifies to
524: \beqn
525: \label{eq:delta0exp}
526:    \delta^{(0)}
527:    &=&
528:       a_s(s_0)
529:       + (1.6398 +  3.5625)\,a_s^{2}(s_0)
530:       + (6.371  + 19.995)\,a_s^{3}(s_0) \nonumber\\
531:    &&
532:       +\: (K_4 + 78.003)\,a_s^{4}(s_0) \nonumber\\
533:    &&
534:       +\: (K_5 + 14.250\,K_4 - 391.54)\,a_s^{5}(s_0)\nonumber\\
535:    &&
536:       +\: (K_6 + 17.813\,K_5 + 45.112\,K_4 + 1.58333\,\beta_4 - 8062.1)\,a_s^{6}(s_0)~,
537: \eeqn
538: where for the purpose of later studies we have kept terms up to sixth order.
539: When only two numbers are given in the parentheses, the first number 
540: corresponds to $K_n$, and the second to $g_n$. 
541: 
542: The FOPT series is truncated at given order 
543: despite the fact that parts of the higher coefficients $g_{n>4}(\xi)$ are 
544: known to all orders and could be resummed. These known parts are the 
545: higher (up to infinite) order terms of the expansion~(\ref{eq:astaylor})
546: that are functions of $\beta_{n\le3}$ and $K_{n\le3}$ only. 
547: In effect, beyond the use of the perturbative expansion
548: of the Adler function~(\ref{eq:adlerpert}), two approximations 
549: have been used to obtain the FOPT series~(\ref{eq:delta0exp}):
550: {\em (i)} the RGE~(\ref{eq:betafun}) has been Taylor-expanded 
551: and terms higher than the given FOPT order have been truncated,
552: and {\em (ii)} this Taylor expansion is used to predict $a_s(-s)$ on
553: the entire $|s|=s_0$ contour.
554: 
555: \subsubsection{Contour-improved fixed-order perturbation theory (\FOPTCI)}
556: \label{sec:cipt}
557: 
558: A more promising approach to the solution of the contour integral~(\ref{eq:an}) 
559: is to perform a direct numerical evaluation by means of single-step
560: integration and using the solution of the RGE to four loops as 
561: input for the running $a_s(-\xi s)$ at each integration 
562: step~\cite{pivov1,pivov2,pert}.
563: It implicitly provides a partial resummation of the (known) higher 
564: order logarithmic integrals and improves the convergence of the 
565: perturbative series. While for instance the third order 
566: term in the expansion~(\ref{eq:delta0exp}) contributes with $17\%$ 
567: to the total (truncated) perturbative prediction, the corresponding 
568: term of the numerical solution amounts only to 
569: %zz $7\%$ 
570: $6\%$ 
571: (assuming $\as(m_\tau^2)=0.35$). This numerical solution of 
572: Eq.~(\ref{eq:knan}) will be referred to as {\it contour-improved} 
573: fixed-order perturbation theory (\FOPTCI) in the following.
574: 
575: \begin{figure}[t]  
576:   \centerline{\epsfysize6.3cm\epsffile{figures/taylor_test_re.eps}\hspace{0.1cm}
577:               \epsfysize6.3cm\epsffile{figures/taylor_test_im.eps}}
578:   \vspace{-0.1cm}
579:   \caption[.]{\label{fig:taylor_test}
580: 	Comparison between the iteratively evolved 
581: 	$\as(m_\tau^2e^{i\varphi})$ (exact solution) and the fourth-order
582: 	Taylor expansion~(\ref{eq:astaylor}) on the circle 
583: 	$\varphi=-\pi\dots\pi$~\cite{menke_phd}. This example
584: 	uses $\as(m_\tau^2)=0.35$.}
585: \end{figure} 
586: Single-step integration also avoids the Taylor approximation of the RGE on 
587: the entire contour, since $a_s$ is iteratively computed from the previous 
588: step using the full known RGE.\footnote
589: {
590: 	It has been tested that the single-step integration, \ie,
591: 	$a_s(\xi s_0 e^{i(\varphi+\Delta\varphi)})$ is computed from
592: 	$a_s(\xi s_0 e^{i\varphi})$, using a modified Euler method 
593: 	(that is, a Taylor expansion of $a_s$ in each integration step),
594: 	or using explicit $4^{\rm th}$ order Runge-Kutta integration,
595: 	leads to the same result, as far as the step size is chosen
596: 	to be small enough (1000 integration steps on the contour
597: 	are found to be largely sufficient).
598: }
599: Figure~\ref{fig:taylor_test} shows the comparison between 
600: the iteratively evolved $\as(m_\tau^2e^{i\varphi})$ (referred
601: to as the exact solution), and the fourth-order Taylor expansion~(\ref{eq:astaylor}) 
602: on the circle $\varphi=-\pi\dots\pi$~\cite{menke_phd}. The differences
603: appear to be significant far away from the reference value at $(\varphi=0)$.
604: However, as shown below, the numerical effect of this discrepancy
605: on the contour integral is small compared to the main difference 
606: between FOPT and \FOPTCI, which is the truncation of the perturbative 
607: series after integration.
608: 
609: \subsubsection{Effective charge perturbation theory (ECPT)}
610: \label{sec:ecpt}
611: 
612: Inspired by the pioneering work in~\cite{grunberg1,grunberg2,dhar1,dhar2,pms} 
613: the use of effective charges to approach the perturbative prediction
614: of \Rtau has become quite popular in the recent 
615: past~\cite{maxwellrs1,raczka,maxwellrs2,pivoveffch}. 
616: The advocated advantage of this technique is that the perturbative 
617: prediction of the effective charge is renormalization scheme (RS) and
618: scale invariant since it is a physical observable. The effective $\tau$ 
619: charge is defined by
620: \beq
621: 	a_\tau = \delta^{(0)}~,
622: \eeq
623: and obeys the RGE
624: \beqn
625: \label{eq:betataufun}
626: 	\frac{d a_\tau}{d \ln s}
627: 	&=& 
628: 	\beta_\tau(a_\tau) = -a_\tau^2\sum_{n=0}^\infty \beta_{\tau,n} a_\tau^n~.
629: \eeqn
630: The first two terms in the expansion of $\beta_\tau$ are universal
631: (\ie, RS-independent), while the higher order terms must be 
632: computed using perturbation theory. The prescription for this 
633: computation is obtained from the relation between the renormalization 
634: scheme and scale used to obtain the FOPT coefficients in the 
635: series~(\ref{eq:kngn}), and the requirement of RS and scale 
636: invariance. Since $a_\tau$ is scheme and scale invariant, its 
637: evolution coefficients $\beta_{\tau,i}$ have the same property.
638: ``{\em Another way of looking at this is to say that the method of 
639: effective charges involves a specific choice of renormalization
640: scheme and scale (\dots), so that the energy evolution of the 
641: observable is identical to the beta-function evolution of the 
642: coupling}''~\cite{maxwellagain}. 
643: 
644: The function $\beta_\tau$ is obtained from the equation
645: \beq
646: 	\frac{d a_\tau}{d a_s}
647: 	=
648: 	\frac{d \beta_\tau(a_\tau)}{d \beta(a_s)}~,
649: \eeq
650: which, setting $\xi=1$, can be rewritten in the form 
651: \beq
652: \label{eq:effchargerelation}
653: 	\beta_\tau(a_\tau) 
654: 	= \frac{d a_\tau}{d\ln s}
655: 	= \frac{d a_\tau}{d a_s}
656: 	  \frac{d a_s}{d\ln s}\bigg|_{a_s=a_s(a_\tau)}
657: 	= \frac{d a_\tau}{d a_s}
658: 	  \beta(a_s)\bigg|_{a_s=a_s(a_\tau)}~.
659: \eeq
660: Inserting Eq.~(\ref{eq:kngn}) for $a_\tau$ on the l.h.s. and 
661: the r.h.s. of Eq.~(\ref{eq:effchargerelation}), and equating
662: the coefficients for each order in $a_s$ leads to the coefficients\footnote
663: {
664: 	For completeness we also give the sixth-order coefficient, which 
665: 	is later needed for extrapolation tests
666: 	\beqns
667: 	\beta_{\tau,5} &=& 
668:                         \beta_5 - 4 \,\beta_4 c_2
669:                         + 2\,\beta_3 (4 \,c_2^2 - c_3)
670:                         + 4\,\beta_2 (-2 \,c_2^3 + c_2 c_3)
671: 			+ \beta_1 (-6 \,c_2^4 + 16 \,c_2^2 c_3 - 3 \,c_3^2 - 
672:    		              8 \,c_2 c_4 + 2 \,c_5) \\
673: 			&&
674: 			+\: 4\,\beta_0 (12 \,c_2^5 - 30 \,c_2^3 c_3 + 12 \,c_2 c_3^2 + 
675: 			   14 \,c_2^2 c_4 - 4 \,c_3 c_4 - 5 \,c_2 c_5 + c_6)~.
676: 	\eeqns
677: }
678: \beqn
679: 	\beta_{\tau,2} &=& \beta_2 - \beta_1 c_2 + \beta_0 (-c_2^2 + c_3)~, \\
680: 	\beta_{\tau,3} &=& \beta_3 - 2 \,\beta_2 c_2 + \beta_1 c_2^2 
681:                            + 2 \,\beta_0 \left( 2\,c_2^3 - 3 \,c_2 c_3 + c_4\right)~, \\
682: 	\beta_{\tau,4} &=& \beta_4 - 3 \,\beta_3 c_2 
683:                            + \beta_2 (4 \,c_2^2 - c_3)
684:                            + \beta_1 (-2 \,c_2 c_3 + c_4) \nonumber\\
685: 			&&
686:                            +\: \beta_0 (-14 \,c_2^4 + 28 \,c_2^2 c_3 - 5 \,c_3^2 
687:                                  - 12 \,c_2 c_4 + 3 \,c_5)~,
688: \eeqn
689: where $c_i=K_i + g_i(1)$ (\cf\  Eq.~\ref{eq:kngn}) and
690: $\beta_{\tau,0}=\beta_0$, $\beta_{\tau,1}=\beta_1$ because $K_0=K_1=1$. 
691: Inserting the known $\beta_i$ and $c_i$ coefficients, one finds~\cite{pivoveffch}
692: \beqn
693: 	\beta_{\tau,2} &=& -12.320~, \nonumber\\
694: 	\beta_{\tau,3} &=& -182.72 + 4.5\,K_4~, \nonumber\\
695: \label{eq:betataunum}
696: 	\beta_{\tau,4} &=& -236.2 + \beta_4 - 40.27\,K_4 + 6.75\,K_5~,  \\
697: 	\beta_{\tau,5} &=& 16427 - 20.80\,\beta_4 + \beta_5 - 
698:                            521.6\,K_4 - 65.79\,K_5 + 9\,K_6~,\nonumber
699: \eeqn
700: with a noticeable sign alteration for $\beta_{\tau,2}$ (and also for 
701: $\beta_{\tau,3}$ if $K_4<40$). Neglecting nonperturbative contributions, 
702: one has
703: \beq
704: 	a_\tau(s_0)=\RtauVpA(s_0)/(3\,|V_{ud}|^2)-1~, 
705: \eeq
706: which using Eq.~(\ref{eq:rtauvpa}) and the $|V_{ud}|$ value given in   
707: Section~\ref{sec:tauspecfun} yields $a_\tau(m_\tau^2)=0.2219\pm0.0055$.
708: 
709: \subsubsection{Estimating unknown higher order perturbative coefficients}
710: \label{sec:estimhigherorder}
711: 
712: Compared with the \MSbar RGE~(\ref{eq:betafun}) the $\beta_\tau(a_\tau)$ 
713: series is significantly less convergent, taking into account that 
714: $a_\tau(m_\tau^2)\simeq1.8\times a_s(m_\tau^2)$. This property has been 
715: used in the past to estimate an effective $K_4^{\rm eff}$ coefficient,
716: which was found to be $K_4^{\rm eff} \sim 27$~\cite{k4_pms}. It 
717: approximates the true value of $K_4$ in the limit of vanishing 
718: higher order contributions. In effect, the derivative of 
719: $a_\tau(m_\tau^2)$ with respect to $K_4$ is large\footnote
720: {
721: 	One can obtain the solution for $a_\tau(m_\tau^2)$ as a function 
722: 	of the $\beta_\tau$ coefficients without explicitly 
723: 	integrating the RGE~(\ref{eq:betataufun}), by using as fixed point 
724:         of the differential equation the one-loop solution at $s\to\infty$,
725: 	$\Lambda^2=s e^{-1/(\beta_0 a_\tau(s))}$, so that the four-loop
726: 	$a_\tau(\infty\to m_\tau^2)$ evolution reproduces the experimental 
727: 	value. This gives $\Lambda=0.41\gev$.
728: }
729: \beq
730: \label{eq:k4estimate}
731:  	\frac{d a_\tau(m_\tau^2)}{dK_4}\bigg|_{a_\tau(m_\tau^2)=0.22,\,K_4=25}
732: 	\simeq 0.0020~,
733: \eeq
734: so that the precise experimental error on $a_\tau(m_\tau^2)$ allows us to
735: predict $K_4$ with an accuracy of a few units, if we assume that the 
736: subleading contributions can be neglected. However, this assumption
737: is unfounded since, taking naively the unknown higher order coefficients 
738: to grow like a geometric series, \ie, $z_n\sim z_{n-1}^2/z_{n-2}$, with $z=\beta,K$, 
739: and using the estimate $K_4=25$ so that $K_5\sim98$, $\beta_{\tau,4}\sim-360$ 
740: (with $\beta_4\sim222$), one finds
741: \beq
742: \label{eq:k5estimate}
743:  	\frac{d a_\tau(m_\tau^2)}{dK_5}\bigg|_{a_\tau(m_\tau^2),K_4,K_5,\beta_4\dots}
744: 	\simeq 0.0004~,
745: \eeq
746: which is not small with respect to~(\ref{eq:k4estimate}), in particular considering
747: the coarseness of the assumptions that went into the estimate~(\ref{eq:k5estimate}).
748: It is found that even the sixth-order coefficient is not negligible. Assuming the 
749: same geometric growth, one would expect $K_6\sim395$ with a derivative of 
750: $d a_\tau(m_\tau^2)/dK_6\sim5\times10^{-5}$, which appears small. However,
751: considering that the uncertainty on $K_6$ is at least of the same size as
752: its estimated value (which is probably still optimistic) generates an 
753: uncertainty in $a_\tau(m_\tau^2)$ that is equivalent to $\sigma(K_4)\sim10$. 
754: Taking correspondingly
755: the fifth-order uncertainty of the order of the estimated value of $K_5$,
756: we conclude that the systematic uncertainty on $K_4$ obtained with the 
757: effective charge scheme is at least of the order of $\sigma(K_4)\sim23$.
758: 
759: Another method to estimate $K_4^{\rm eff}$ has been suggested 
760: in~\cite{k4_fld}. It consists of increasing the sensitivity of the \FOPTCI
761: prediction of $\delta^{(0)}$ on $K_4$ by reducing the renormalization scale
762: $\xi$ in Eq.~(\ref{eq:knan}). The result $27\pm5$ is consistent with the one 
763: obtained from ECPT. However, these methods are not independent as can be 
764: seen in the following. Using the integral~(\ref{eq:betafunint}), and taking
765: $s\to\infty$ so that $a_\tau,a_s\to 0$, we can write~\cite{pms}
766: \beqn
767: 	&&
768: 	\ln\frac{s}{\Lambda^2}-\ln\frac{s}{\Lambda_\tau^2}
769: 	=
770: 	2\ln\frac{\Lambda_\tau}{\Lambda}
771: 	= 
772: 	\frac{1}{\beta_0a_s(s)} - \frac{1}{\beta_0a_\tau(s)} + \mathcal{O}(a_s)
773: 	= 
774: 	c_2 + \mathcal{O}(a_s)~, \\
775: 	&&\hspace{1cm}
776: 	\Longleftrightarrow\hspace{1cm}
777: 	\Lambda_\tau \simeq \Lambda e^{c_2/\beta_0} \approx 3.2\Lambda~.
778: \eeqn
779: Since the asymptotic scale parameters are reciprocal to the scale,
780: transforming FOPT into ECPT is equivalent to reducing the renormalization 
781: scale, at lowest order. 
782: 
783: \begin{figure}[t]  
784:   \centerline{\epsfysize6.3cm\epsffile{figures/atau_evolution_fopt.eps}
785:               \epsfysize6.3cm\epsffile{figures/atau_evolution_cipt.eps}}
786:   \centerline{\epsfysize6.3cm\epsffile{figures/atau_evolution_xipt.eps}
787:               \epsfysize6.3cm\epsffile{figures/atau_evolution_ecpt.eps}}
788:   \vspace{-0.1cm}
789:   \caption[.]{\label{fig:atau_evolution}
790:      	Evolution of $a_\tau(s_0)$ for the different perturbative methods, 
791:         and for variations of the unknown higher order perturbative
792: 	coefficients $K_4$ (dashed) and $K_5$ (dotted). All methods 
793: 	are normalized to $a_\tau(m_\tau^2)=0.22$. The inserted frames
794: 	give the $\as(m_\tau^2)$ values derived from this $a_\tau$
795: 	value in the respective methods. The error bars show
796: 	the uncertainties, added in quadrature, from the $K_4$ and $K_5$ 
797: 	variations quoted on the plots.}	
798: \end{figure} 
799: In terms of the measurement of $\as(m_\tau^2)$ the three perturbative
800: methods FOPT, \FOPTCI and ECPT use similar information. However their 
801: dependence on the unknown higher order coefficients $K_{n\ge4}$ can be 
802: dramatically different. Figure~\ref{fig:atau_evolution} shows the evolution 
803: of $a_\tau(s_0)$ for FOPT, \FOPTCI with renormalization scales $\xi=1$ 
804: and $\xi=0.4$, and ECPT, all normalized to $a_\tau(m_\tau^2)=0.22$.
805: The solid lines correspond to fifth order perturbation theory with
806: assumed $K_4=25$, $K_5=98$, and for variations of these (dashed and
807: dotted lines). One observes relatively small dependence for FOPT and 
808: \FOPTCI with $\xi=1$, while large and very large effects are found
809: with \FOPTCI ($\xi=0.4$) and ECPT. The latter effect provides the
810: sensitivity to estimate the unknown coefficients. To better analyze this
811: sensitivity, we have compared the $\Delta K_4$ variation in fifth-order
812: ECPT to the fourth-order effect (dashed-dotted line in the lower right
813: hand plot of Fig.~\ref{fig:atau_evolution}). Including the bold estimate
814: for the next order significantly reduces the dependence on the $K_4$
815: term. This cancellation becomes obvious when looking at the numerical
816: expansion~(\ref{eq:betataunum}): the $K_4$ coefficient enters 
817: $\beta_{\tau,3}$ and $\beta_{\tau,4}$ with different signs, and 
818: with a larger coefficient at fifth order to compensate for the $a_\tau$ 
819: suppression. A similar behavior is observed for $K_5$ when turning on the 
820: sixth-order term. We conclude that as for the $K_4$ term, the strong $K_5$ 
821: dependence observed in ECPT is (by part) an artifact of the truncation of 
822: the perturbative series, and it is therefore dangerous to use it for an 
823: estimate of the unknown coefficients.
824: 
825: The embedded frames in Fig.~\ref{fig:atau_evolution} give the 
826: $\as(m_\tau^2)$ values derived from the given $a_\tau$ value in 
827: the respective methods. The error bars indicate the uncertainty, 
828: added in quadrature, from the $K_4$ and $K_5$ variations quoted
829: on the figures. One observes
830: that the difference in $\as$ between FOPT and \FOPTCI is not covered
831: by the uncertainties on $K_4$ and $K_5$. This is because the main
832: contribution to the difference in these methods is due to the 
833: truncation of the {\em known} higher order terms of the Taylor
834: expansion after the contour integration, rather than due to 
835: the unknown coefficients. Due to the good convergence of these
836: series, the dependence on the $K_5$ coefficient is marginal.
837: 
838: It should be mentioned that a large computational effort is underway
839: with the goal to calculate the $K_4$ coefficient. The large
840: number of five-loop diagrams needed to calculate the two-point current
841: correlator at this order is somewhat discouraging, however the results 
842: on two gauge invariant subsets are already available. The subset of order 
843: ${\cal O}(\alpha_s^4n_f^3)$ was evaluated long ago through the summation of 
844: renormalon chains~\cite{beneke1993}, while the much harder subset 
845: ${\cal O}(\alpha_s^4n_f^2)$ was recently calculated~\cite{baikov_a4}.
846: The known parts amount to about half of the expected $K_4$ coefficient
847: assuming a geometric growth. 
848: Using the results of~\cite{k4_pms} for different $n_f$ obtained with the
849: effective charge method, it is possible to compare the $n_f^2$ estimated 
850: term at the $\alpha_s^4$ order of the Adler function expansion 
851: ($3.64 n_f^2$), with the newly computed one ($1.875 n_f^2$). The difference
852: provides a conservative estimate for the uncertainty of the ECPT procedure
853: to predict the full $K_4$ term, as one could expect the sum of all
854: $\alpha_s^4n_f^i$ terms to be closer to the true value than 
855: the term-by-term comparison.
856: A successful test of the procedure has been performed for
857: $K_3$, where all the $\alpha_s^3n_f^i$ terms are exactly known. The two 
858: unknown remaining terms ${\cal O}(\alpha_s^4n_f^1)$ and 
859: ${\cal O}(\alpha_s^4n_f^0)$, the direct calculation of which 
860: is prevented by the present computing power available to this project, 
861: have been estimated~\cite{k4_cal} to be
862: \beq
863:      K_4 = -0.01009\, n_f^3 + 1.875\, n_f^2 - 31.8\, n_f +105.7~,
864: \eeq
865: where the first two terms are known exactly. The calculation for
866: $n_f=3$ yields $K_4\simeq 27$ with an uncertainty estimated to $\pm16$
867: in~\cite{k4_cal}.
868: 
869: \subsubsection{Infrared behavior of the effective charge}
870: \label{sec:infraredbehavior}
871: 
872: In the effective charge scheme, the physical coupling should have a finite 
873: infrared behavior and hence must not suffer from the 
874: Landau-pole that lets \MSbar (but also momentum subtraction schemes) 
875: break down below a scale $\sqrt{s}\sim1\gev$. 
876: The universal two-loop function $\beta_\tau(a_\tau)$ 
877: does however suffer from a singularity at $\sqrt{s}\sim1.1\gev$ 
878: (for $a_\tau(m_\tau^2)=0.22$). Only adding the negative three-loop coefficient
879: regularizes $\beta_\tau(a_\tau)$ 
880: and leads to a freezing $a_\tau(s\to0)\simeq0.62$. Adding higher orders does
881: not qualitatively alter this behavior, however the freezing value is 
882: very unstable against variations of the unknown coefficients (which again 
883: exhibits the mediocre convergence of the ECPT series). The lower right 
884: hand plot of Fig.~\ref{fig:atau_evolution} shows the evolution of $a_\tau(s)$ 
885: for different orders and assumptions upon the unknown higher order coefficients
886: (\cf\  the discussion in~\cite{brodsky}).
887: 
888: \subsubsection{Renormalons and the large-$\beta_0$ expansion}
889: \label{sec:renormalons}
890: 
891: Perturbative series in quantum field theory\footnote
892: {
893: 	The suppression of the constant term in the definition of 
894: 	Eq.~(\ref{eq:ren_series}) is convenient to avoid $1/\alpha$
895: 	terms in the definition of the Borel integral~(\ref{eq:borelint}).
896: 	It is without consequence for the following discussion.
897: }
898: \beq
899: \label{eq:ren_series}
900: 	R_N = \sum_{n=0}^N r_n\alpha^{n+1}~,
901: \eeq
902: be they abelian or not, are divergent at $N\to\infty$ for any 
903: value of $\alpha$.
904: Because the number of Feynman graphs increases at each order $n$, 
905: one expects a factorial growth of the perturbative coefficients of 
906: the corresponding $S$-matrix series
907: % \footnote
908: % {
909: % 	\underline{Digression:}
910: % 	A well-known example for an asymptotic series is the expansion of the
911: % 	exponential integral
912: % 	\beqns
913: % 		{\rm E}_{\rm i}(x)=-\int_{-x}^\infty dt \frac{e^{-t}}{t}
914: % 		~~~\approx~~~		
915: % 		\frac{e^x}{x}\sum_{n=0}^N \frac{n!}{x^n}=
916: % 		\tilde{\rm E}_{\rm i}(x,N) ~,
917: % 	\eeqns
918: % 	which is monotonous for $x>0$ and alternating for $x<0$.
919: % 	(For practical purpose it is useful to compute 
920: % 	$\tilde{\rm E}_{\rm i}(x,N)$ by means of the expression 
921: % 	$-\Gamma(0,-x)-(-1)^{N}\Gamma(2+N)\Gamma(-1-N,-x)$,
922: % 	where $\Gamma(\dots)$ ($\Gamma(\dots,\dots)$) is the complete
923: % 	(incomplete) gamma function.) The series $\tilde{\rm E}_{\rm i}(x,N)$ 
924: % 	diverges for $N\to\infty$, for all values of $x$. As for the Adler 
925: % 	function, the coefficients of $\tilde{\rm E}_{\rm i}$ exhibit a 
926: % 	factorial growth. The smaller $|x|$, the worse is the convergence,
927: % 	and the smaller is the order ($N_{\rm min}$) where minimal sensitivity is  
928: % 	reached. For example, for $x=2$ one finds $N_{\rm min}=1$ with a relative
929: % 	error 
930: % 	$|\tilde{\rm E}_{\rm i}(2,1)/{\rm E}_{\rm i}(x=2)-1|=12\%$.
931: % 	For $x=4$ one finds $N_{\rm min}=3$ and a relative error of $2\%$, and
932: % 	finally for $x=10$ one has $N_{\rm min}=9$ with $0.01\%$ error. One notices
933: % 	that $N_{\rm min}$ is the first order {\em after} the crossing of the
934: % 	exact solution ${\rm E}_{\rm i}(x)$. \\\indent
935: % 	{\em FOPT vs. \FOPTCI}: it is instructive and technically easy to 
936: % 	develop $\tilde{\rm E}_{\rm i}(x,N)$ into a Taylor series 
937: %         $T[\tilde{\rm E}_{\rm i}]$, 
938: % 	and test the perturbative schemes by means of solving the Cauchy integral
939: % 	$(2\pi)^{-1}\int_{-\pi}^\pi T[\tilde{\rm E}_{\rm i}
940: % 	 (x+\varepsilon e^{i\varphi},N)]d\varphi$.
941: % 	While \FOPTCI (single-step solution of the integral
942: % 	and resummation of all known orders) %(trivially)
943: %         reproduces 
944: % 	$T[\tilde{\rm E}_{\rm i}(x,N)]$, the FOPT integration is less accurate.
945: % }
946: \beq
947: \label{eq:rnfactgrowth}
948: 	r_n \sim K a^n n! n^\gamma~,
949: \eeq
950: with constant $K$, $a$, $\gamma$. The divergent behavior of perturbative
951: expansions often has physical significance: it indicates the 
952: nonperturbative structure of the vacuum and its excitations.
953: To be a useful approximation of the true value $R$, the 
954: series~(\ref{eq:ren_series}) must be {\em asymptotic to $R$} for 
955: given $\alpha$, \ie, it exists a number $K_N$ so that\footnote
956: {
957: 	Our discussion of renormalons has greatly benefited from
958: 	the comprehensive review given in~\cite{benekephysrp}.
959: }
960: $	|R - R_N| < K_{N+1}\alpha^{N+2}$.
961: If $r_n\sim K_n$, the best approximation of $R$ is achieved when
962: the series is truncated at its minimal (absolute) term and the 
963: truncation error can be roughly estimated by the size of the minimal 
964: term. The accuracy in the point of minimal sensitivity depends on
965: the strength of the perturbative expansion parameter (the coupling).
966: The larger it is, the lower the order when minimal sensitivity is 
967: reached, and the worse is the approximation. Note that as long as 
968: the computed order is below the order of minimal sensitivity, there
969: is no actual difference between a divergent or a convergent series.
970: 
971: A convenient way to deal with this divergence is to consider the Borel 
972: transform 
973: \beq
974:    B[R](t) \equiv
975:       \sum_{n=0}^\infty\frac{r_n}{n!}t^n~,
976: \eeq
977: which is believed to have a finite radius of convergence in the $t$-plane~\cite{mueller}.
978: The $n^{\rm th}$ fixed order perturbation coefficient is then generated by the 
979: $n^{\rm th}$ derivative 
980: \beq
981: \label{eq:generator}
982:    r_n = \frac{d^nB[R](t)}{dt^n}\bigg|_{t=0}~.
983: \eeq
984: The explicit factor $n!$ in the denominator of $B[R](t)$ makes the 
985: Borel-transformed series much better behaved. Summing up all orders 
986: leads to the integral representation\footnote
987: {
988: 	The Borel integral is easily derived by using the integral
989: 	representation $n!=\int_0^\infty dv\exp(-v)v^n$, and exchanging
990: 	summation and integration in Eq.~(\ref{eq:ren_series})
991: 	(see, \eg, \cite{maxwellwhoelse}). Note however that this interchange
992: 	is formally not justified in presence of the Landau pole for the 
993: 	coupling constant.
994:  }
995: \beq
996: \label{eq:borelint}
997:    R(\alpha) = \intl_0^\infty dt\,e^{-t/\alpha} B[R](t)~.
998: \eeq
999: If the integral exists, it gives the Borel sum of the original divergent
1000: series. Existence requires that $B[R](t)$ 
1001: has no singularities in the integration range. However, for a factorial
1002: behavior akin to Eq.~(\ref{eq:rnfactgrowth}) (non-alternating 
1003: series), the Borel transform is given by
1004: $B[R](t) \propto (1-a t)^{-1-\gamma}$. Hence, if $\gamma$ is positive 
1005: there exists a singularity $t=1/a$ so that the integral~(\ref{eq:borelint}) 
1006: does not exist. Indeed, ``{\em the divergent behavior of the original series 
1007: is encoded in the singularities of its Borel transform}''~\cite{benekephysrp}.
1008: One notices that the larger $a$, \ie, the faster diverging the series is, 
1009: the closer to zero moves the singularity. 
1010: 
1011: \begin{figure}[t]
1012:   \epsfxsize8cm
1013:   \centerline{\epsffile{figures/renormalons.eps}}
1014:   \vspace{-0.1cm}
1015:   \caption[.]{\label{fig:renormalons}
1016:                Multi-fermion loop insertion (renormalons) into a
1017:                fermion anti-fermion vacuum polarization diagram.}
1018: \end{figure} 
1019: Let us consider a renormalized fermion-loop insertion into a quark-antiquark
1020: vacuum polarization diagram (\cf\  one bubble in Fig.~\ref{fig:renormalons}).
1021: Assuming dominance of the $(\beta_0 a_s(-s))^n$ term (denoted as the 
1022: {\em large-$\beta_0$ expansion}), one can resum this chain with a 
1023: large number of $n$ bubbles. If one neglets higher order terms of the 
1024: $\beta$-function, it is thus possible to derive estimates for the FOPT 
1025: coefficients of a given perturbative series at all orders. This 
1026: assumption is supported empirically by the observation that 
1027: the $\beta_0$ term dominates second order radiative corrections for many 
1028: observables in the \MSbar  scheme~\cite{beneke}. (We shall however 
1029: anticipate here that the accuracy of these estimates is insufficient 
1030: for the purpose of a precise perturbative prediction of $\Rtau$, and 
1031: difficult to control.) The procedure provides a {\it naive non-abelianization} 
1032: of the theory since lowest order radiative corrections do not include gluon 
1033: self-coupling. 
1034: %At sufficiently large orders of $n$ vacuum polarization 
1035: %bubbles, the coefficients diverge as 
1036: %\beq
1037: %\label{eq:appr}
1038: %	r_n \sim C_k n!\,n^{\gamma_k}(\beta_0/k)^n~,
1039: %\eeq
1040: %with constant $C_k$, $\gamma_k$, $k$.
1041: 
1042: Following this line, the Adler function can be expanded as~\cite{benekebraun}
1043: \beq
1044: \label{eq:adlerlargebeta}
1045:     D(s) = 1 + a_s\sum_{n=0}^N a_s^n\left(d_n\beta_0^{n} + \delta_n\right)~,
1046: \eeq
1047: where the coefficients $d_n$ are computed in terms of fermion bubble 
1048: diagrams\footnote
1049: {
1050: 	One identifies the coefficients $d_n$ in Eq.~(\ref{eq:adlerlargebeta})
1051: 	with their leading-$n_f$ pieces $d_n^{[n]}$ in the expression
1052: 	$d_n = d_n^{[n]}n_f^k + \dots + d_k^{[0]}$. It has been shown 
1053: 	in~\cite{broadhurst} that the $d_n^{[n]}$ in \MSbar
1054: 	are given by ($s=\mu^2$) 
1055: 	\beqns
1056: 		d_n^{[n]} = 2^{1-n} n! 
1057:                 \sum_{k=0}^n \left(-\frac{5}{9}\right)^{\!\!k}
1058:                              \frac{\Psi_{n+2-k}^{[n+2-k]}}
1059:                                   {k!(n-k)!}~,
1060: 	\eeqns
1061: 	with the generating function
1062: 	\beqns
1063: 		\Psi_m^{[m]}=
1064: 			\frac{3^{2-m}}{2}\left(\frac{d}{dx}\right)^{\!\!m-2}
1065:   			\!\!\!\!\!P(x)\bigg|_{x=1}~,
1066: 	\hspace{0.5cm}\mbox{and}\hspace{0.5cm}
1067: 		P(x) = \frac{32}{3(1+x)}
1068:                        \sum_{k=2}^{\infty}\frac{(-1)^kk}{(k^2-x^2)^2}~.
1069: 	\eeqns
1070: } 
1071: (see Fig.~\ref{fig:renormalons}). The
1072: \MSbar Borel transform of Eq.~(\ref{eq:adlerlargebeta}) is given 
1073: by~\cite{broadhurst,beneke1993}
1074: \beq
1075: \label{eq:boreladler}
1076:       B[D](u) = \frac{32\pi}{3}e^{\frac{5}{3}u}\frac{u}{1-(1-u)^2}
1077:                 \sum_{k=2}^\infty\frac{(-1)^kk}{(k^2-(1-u)^2)^2}~,
1078: \eeq
1079: where $u=-\beta_0 t$, and where we have chosen the renormalization scale 
1080: to be equal to the momentum scale. Neglecting the corrections
1081: $\delta_n$, the series~(\ref{eq:adlerlargebeta}) leads to the
1082: large-$\beta_0$ expansion of $D(s)$. The first elements of the 
1083: series are
1084: \beqn
1085: \label{eq:dnlargebeta}
1086:       d_0          = 1~,\hspace{1cm}
1087:       d_1\beta_0   = 0.496~,\hspace{1cm}
1088:       d_2\beta_0^2 = 15.71~,\hspace{1cm} \nonumber\\
1089:       d_3\beta_0^3 = 24.83~,\hspace{1cm}
1090:       d_4\beta_0^4 = 787.8~,\hspace{1cm}
1091:       d_5\beta_0^5 = -1991 ~,\hspace{1cm}
1092: \eeqn
1093: which compare only coarsely with the exact terms~(\ref{eq:kf1})
1094: where these are known.
1095: The Borel transform $B[D](u)$ has singularities on the real axis, which  
1096: can be distinguished in infrared (IR renormalons) for small virtuality,
1097: and ultraviolet singularities (UV renormalons) for high virtuality of 
1098: the exchanged gluon. One finds an infinite sequence of IR (UV) renormalons
1099: at positive (negative) integer $u$, with the exception of $u=1$
1100: (see Fig.~\ref{fig:singularities}).
1101: 
1102: The integral~(\ref{eq:borelint}) may still be solvable in presence of
1103: renormalons by acquiring an imaginary part from moving the integration 
1104: contour above or below the real axis. The associated ambiguity is 
1105: exponentially small in the expansion parameter $\alpha$, and is 
1106: therefore of nonperturbative origin. Correspondingly, the IR poles 
1107: give rise to ambiguities when one uses the 
1108: generators~(\ref{eq:generator}) to recompute $D(s)$ from 
1109: its Borel transform. Using in first order 
1110: \begin{figure}[t]
1111:   \epsfxsize10.6cm
1112:   \centerline{\epsffile{figures/singularities.eps}}
1113:   \vspace{-0.3cm}
1114:   \caption[.]{\label{fig:singularities}
1115:               Ultraviolet (UV) and infrared (IR) renormalons
1116:               in the Borel plane of the Adler function.}
1117: \end{figure} 
1118: \beq
1119: \label{eq:alphaslog}
1120:    a_s(s) = \frac{1}{\beta_0{\rm ln}(s/\Lambda^2)}~,
1121: \eeq
1122: one obtains for the associated ambiguity for $u=2,3,\dots$
1123: \beqn
1124:    \Delta D(s) 
1125:    \sim
1126:       e^{-u/(\beta_0a_s)} =
1127:       \left(\frac{\Lambda^2}{s}\right)^{\!\!u}~.
1128: \eeqn
1129: These IR renormalons proportional to $(\Lambda^2/s)^u$ are reabsorbed 
1130: into the nonperturbative terms of the OPE.
1131: %\footnote
1132: %{
1133: %	Infrared renormalons are not connected to the existence
1134: %	of the Landau pole. While the former always exist
1135: %	(their location does not depend on the higher $\beta_n$
1136: %	coefficients~\cite{beneke}), the Landau pole depends on the $\beta$
1137: %	function. For instance, we have seen in Section~\ref{sec:infraredbehavior} 
1138: %	that the ECPT scheme is IR finite.
1139: %}. 
1140: The absence of a $u=1$ IR 
1141: renormalon is thereby related to the impossibility to build a gauge 
1142: invariant operator of dimension $D=2$, which is an important property 
1143: that is exploited for the determination of $\as$ from $\Rtau$.
1144: The lowest IR renormalon is hence of the order $(\Lambda^2/s)^2$.
1145: On the contrary to the UV renormalons, this term is scale independent
1146: and cannot be decreased~\cite{keq1}. It should therefore have physical 
1147: meaning and can be identified with the gluon condensate 
1148: $\langle 0|G_{\mu\nu}G^{\mu\nu}|0\rangle$ as the leading infrared 
1149: contribution to the Adler function. However, while IR renormalons
1150: lead to nonperturbative operators for power corrections, it does
1151: not necessarily contain all of them. For example, at $D=4$ one 
1152: misses the quark condensate $\langle0|q \qbar|0\rangle$, which is
1153: generated by chiral symmetry breaking that does not occur in 
1154: perturbation theory.
1155: 
1156: Due to asymptotic freedom, governed by the negative sign in front
1157: of the RGE $\beta$-function, the UV renormalons arise on the negative 
1158: real axis\footnote
1159: {
1160: 	In QCD, the presence of IR renormalons implies that 
1161: 	nonperturbative corrections should be added to define the theory 
1162: 	unambiguously. The same is true for UV renormalons in QED.
1163: 	This corresponds to the observation of large coupling 
1164: 	constants in the respective energy regimes.
1165: } 
1166: (on the contrary to QED where they occur on the positive 
1167: side) so that they are outside the integration range of~(\ref{eq:borelint}) 
1168: and insofar harmless, that is, Borel-summable~\cite{mueller}. Let us 
1169: return for a moment to the series~(\ref{eq:ren_series}) in the 
1170: large-$\beta_0$ expansion. For large $n$, the factorial growth of the 
1171: perturbation series is dominated by the contribution of the leading UV 
1172: renormalon $k=-1$ with alternating coefficients 
1173: $r_n \sim n! n^\gamma(-\beta_0)^n$. A guess at which 
1174: order $N_{\rm min}$ minimal sensitivity is achieved can be obtained from the 
1175: argument that the series is convergent, \ie, reliable at order $n+1$ 
1176: if $\alpha (r_{n+1}/r_n<1)$. Considering leading IR and UV renormalons only, 
1177: one finds from the asymptotic behavior
1178: $r_{n+1}/r_n\sim n\beta_0/2$ (IR) and $r_{n+1}/r_n\sim -n\beta_0$ (UV).
1179: Hence the breakdown of convergence is first caused by the UV renormalon
1180: and $N_{\rm min}\sim1/(\beta_0 a_s)$ is of order one. A series truncated 
1181: at finite order 
1182: brings an intrinsic limitation of accuracy along with it. The associated 
1183: error is reasonably estimated with the magnitude of the order 
1184: $N=N_{\rm min}$ term of the perturbative series. That gives using 
1185: for the leading UV renormalon 
1186: \beq
1187:   |r_{N}|a_s^{N} 
1188:    \sim
1189:       N!N^\gamma(-\beta_0 a_s)^{N} \nonumber \\
1190: %   &\simeq&
1191: %      N^{N} e^{-N}
1192: %     \sqrt{2\pi N}N^\gamma (-\beta_0a_s)^{N}~~~~~{\rm (Stirling~formula)}
1193: %     \nonumber \\
1194: %   &\sim& 
1195: %     e^{-1/(\beta_0a_s)}(\beta_0a_s)^{-\gamma}
1196: %     \sqrt{2\pi/(\beta_0\as)} \nonumber \\
1197: %   &\sim&
1198: 	\sim
1199:      \frac{\Lambda^2}{s} \times {\rm logarithms}~,
1200: \eeq
1201: where the r.h.s. is obtained with the use of the Stirling formula, 
1202: and where we have fixed the renormalization scale at $\mu^2=s$.	
1203: The truncation of the perturbative series is accompanied
1204: by an uncertainty which scales like $1/s$, \ie, with apparent 
1205: dimension $D=2$. This contribution is however not to be confounded
1206: with nonperturbative IR renormalon ambiguities\footnote
1207: {
1208: 	It has been shown in~\cite{keq1} that the truncation 
1209: 	uncertainty at $N_{\rm min}$ scales actually as $\Lambda^2s/\mu^4$ 
1210: 	once the renormalization scheme dependence of $\Lambda$ is taken into
1211: 	account.
1212: }. 
1213: Similarly to Eq.~(\ref{eq:adlerlargebeta}) the $\Rtau$ FOPT 
1214: series~(\ref{eq:kngn}) can be written as
1215: \beq
1216: \label{eq:rtaulargebeta}
1217:     \delta^{(0)}(s) = 1 + a_s\sum_{n=0}^Na_s^n\left(d_n^\tau\beta_0^{n} 
1218:                                                  + \delta_n^\tau\right)~,
1219: \eeq
1220: where $d_n^\tau\beta_0^{n}+\delta_n^\tau=K_{n+1}+g_{n+1}(1)$.
1221: Inserting Eq.~(\ref{eq:boreladler}) into (\ref{eq:rtauadler}), and taking 
1222: advantage of the factorized $s$ dependence in the large-$\beta_0$ expansion, 
1223: the Borel transform of~(\ref{eq:rtaulargebeta}) is~\cite{beneke1993}
1224: \beq
1225:       B[\Rtau](u)=B[D](u)\frac{\sin(\pi u)}{\pi}
1226:                  \left(\frac{1}{u} + \frac{2}{1-u} + \frac{3}{3-u} 
1227:                        \frac{1}{4-u}\right)~.
1228: \eeq
1229: The first elements of the series are
1230: \beqn
1231: \label{eq:dntau}
1232:       d_0^\tau          = 1~,\hspace{1cm}
1233:       d_1^\tau\beta_0   = 5.119~,\hspace{1cm}
1234:       d_2^\tau\beta_0^2 = 28.78~,\hspace{1cm} \nonumber\\
1235:       d_3^\tau\beta_0^3 = 156.6~,\hspace{1cm}
1236:       d_4^\tau\beta_0^4 = 900.9~,\hspace{1cm}
1237:       d_5^\tau\beta_0^5 = 4867 ~.\hspace{1cm}
1238: \eeqn
1239: They compare reasonably well with the exact terms~(\ref{eq:delta0exp})
1240: where these are known.
1241: In this approach, the unknown fixed-order fourth-order coefficient
1242: is estimated to be $K_4\sim79$, which is significantly larger
1243: than the estimate based on ECPT, and also larger than the one 
1244: obtained from the large-$\beta_0$ expansion of the Adler 
1245: function~(\ref{eq:dnlargebeta}).
1246: 
1247: Apart from the conceptual interest in the study of renormalons, the
1248: main question regarding the concrete use of the large-$\beta_0$ expansion
1249: certainly is: how large are the uncertainties? From Eq.~(\ref{eq:dntau})
1250: we see that the series is positive and that minimal sensitivity is 
1251: reached at $n\sim5$. On the other hand, the resummation of the known $g_n(\xi)$
1252: coefficients~(\ref{eq:delta0exp}) shows that large negative coefficients
1253: occur at higher order FOPT so that one would need a far larger than
1254: a geometric growth of the $K_n$ coefficients to keep the series 
1255: positive. A direct comparison between the large-$\beta_0$ expansion 
1256: coefficients for the Adler function~(\ref{eq:dnlargebeta}) with the 
1257: exact calculation~(\ref{eq:kf1}) reveals
1258: an oscillating behavior of the large-$\beta_0$ series, \ie, small $d_n$ 
1259: coefficients for $n$-even and large $d_n$ for $n$-odd. For example,
1260: the large-$\beta_0$ $d_3$ coefficient is substantially overestimated.
1261: One also has to worry about important contributions from the $d_5$ 
1262: term, which is found to be $788$. As a consequence, our conclusion 
1263: is that the uncertainties associated with the large-$\beta_0$ expansion 
1264: are not under sufficient control to improve the perturbative prediction 
1265: of $\Rtau$. It is also an unreliable estimator of the uncertainty of
1266: the perturbative series.
1267: 
1268: \subsubsection{Comparison of the perturbative methods}
1269: \label{sec:pert_comp}
1270: 
1271: Some comparisons between the fixed-order PT methods have already been 
1272: given for the discussion of the estimate of the unknown higher order 
1273: coefficients in Section~\ref{sec:estimhigherorder}. The slow convergence 
1274: of ECPT (and CIPT at small renormalization scales, $\xi\ll1$) has been
1275: pointed out in this respect. To further study the convergence of the 
1276: perturbative series, we give in Table~\ref{tab:intsol} the contributions 
1277: of the different orders in PT to $\delta^{(0)}$ for the various approaches
1278: and using $\as(m_\tau^2)=0.35$. We assumed here $K_4=25$ and a geometric
1279: growth of all other unknown PT and RGE coefficients. In the case of CIPT 
1280: the results are given for the various techniques used to evolve 
1281: $\as(s_0e^{i\varphi})$: the truncated Taylor expansion~(\ref{eq:astaylor}),
1282: Runge-Kutta integration of the RGE with $\xi=1$, and Runge-Kutta integration
1283: with $\xi=0.4$. 
1284: 
1285: \begin{table}[t]
1286:   \caption[.]{\label{tab:intsol}
1287:               Massless perturbative contribution to $\Rtau(m_\tau^2)$ 
1288: 	      for the various methods considered, and at orders 
1289:               $n\ge1$ with $\asm=0.35$. The value of $K_4$ is 
1290:               set to 25, while all unknown higher order $K_{n>4}$ and 
1291: 	      $\beta_{n>3}$ coefficients are assumed to follow a geometric 
1292: 	      growth.}
1293: \begin{center}
1294: \setlength{\tabcolsep}{0.0pc}
1295: \begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lrrrrrrrr} 
1296: \hline\noalign{\smallskip}
1297:   	& \mc{7}{c}{$\delta^0$} \\
1298: Pert. Method	
1299:         & $n=1$	 & $n=2$  & $n=3$   & $(n=4)$ & $(n=5)$ & $(n=6)$   & $\sum_{n=1}^4$ & $\sum_{n=1}^6$ \\
1300: \noalign{\smallskip}\hline\noalign{\smallskip}
1301: FOPT ($\xi=1$)
1302: 	& 0.1114 & 0.0646 & 0.0365  & 0.0159  & 0.0010  & $-0.0086$ & 0.2283      & 0.2208 \\
1303: \FOPTCI (Taylor RGE, $\xi=1$)         	
1304: 	& 0.1573 & 0.0317 & 0.0126  & 0.0042  & 0.0011  & 0.0001    & 0.2058      & 0.2070 \\
1305: \FOPTCI (full RGE, $\xi=1$)           	
1306: 	& 0.1524 & 0.0311 & 0.0129  & 0.0046  & 0.0013  & 0.0002    & 0.2009      & 0.2025 \\
1307: \FOPTCI (full RGE, $\xi=0.4$)           	
1308: 	& 0.2166 &$-0.0133$&0.0006  &$-0.0007$& 0.0010  &$-0.0007$  & 0.2032      & 0.2048 \\
1309: ECPT
1310: 	& 0.1442 & 0.2187 &$-0.1195$&$-0.0344$&$-0.0160$&$-0.0120$  & 0.2090      & 0.1810 \\
1311: Large-$\beta_0$ expansion
1312: 	& 0.1114 & 0.0635 & 0.0398  & 0.0241  & 0.0155  & 0.0093    & 0.2388      & 0.2636 \\
1313: \noalign{\smallskip}\hline
1314: \end{tabular*}
1315:   \end{center}
1316: \end{table}
1317: As advocated in~\cite{pert}, faster convergence is observed for CIPT
1318: compared to FOPT yielding a significantly smaller error associated with 
1319: the renormalization scale ambiguity (while, somewhat counter intuitive, 
1320: the uncertainty due to the unknown $K_4$ and higher order coefficients 
1321: is similar in both approaches). Our coarse extrapolation
1322: of the higher order coefficients could indicate that minimal sensitivity 
1323: is reached at $n\sim5$ for FOPT, while the series further converges
1324: for CIPT. Although the Taylor expansion in the CIPT integral exhibits 
1325: significant deviations from the exact solution on the integration circle 
1326: (\cf\  Fig.~\ref{fig:taylor_test}), the actual numerical effect from this 
1327: on $\delta^{(0)}$ is small (\cf\  second and third column in 
1328: Table~\ref{tab:intsol}). As discussed in Section~\ref{sec:ecpt}, the 
1329: convergence of the ECPT series is much worse than for FOPT and CIPT.
1330: Consequently, the difference between truncation at $n=4$ and $n=6$ is 
1331: expected to be significant. A similar behavior is observed for the 
1332: large-$\beta_0$ expansion.
1333: 
1334: The CIPT series is found to be better behaved than FOPT (as well as ECPT)
1335: and is therefore to be preferred for the numerical analysis of the $\tau$
1336: hadronic width. As a matter
1337: of fact, the difference in the result observed when using a Taylor 
1338: expansion and truncating the perturbative series after integrating
1339: along the contour (FOPT) with the exact result at given order (CIPT) 
1340: is exhibiting the incompleteness of the perturbative series. However, 
1341: it is even worse than that since large known coefficients are neglected
1342: in FOPT so that the difference between CIPT and FOPT is actually overstating
1343: the perturbative truncation uncertainty. This can be verified by studying 
1344: the behavior of this difference for the various orders in perturbation 
1345: theory given in Table~\ref{tab:intsol}. The CIPT-vs.-FOPT discrepancy 
1346: increases with the addition of each order, up to order four where a 
1347: maximum is reached. Adding the fifth order does not reduce the effect,
1348: and only beyond fifth order the two evaluations become asymptotic
1349: to each other.
1350: As a consequence varying the unknown higher order coefficients {\em and} using 
1351: the difference between FOPT and CIPT as indicator of the theoretical 
1352: uncertainties overemphasizes the truncation effect. 
1353: 
1354: \subsubsection{Quark-mass and nonperturbative contributions}\label{sec:nonpert}
1355: 
1356: Following SVZ~\cite{svz}, the first contribution to \Rtau\  beyond the 
1357: $D=0$ perturbative expansion is the non-dynamical quark mass correction 
1358: of dimension\footnote
1359: {
1360: 	Corrections of dimension $D=2$ refer to the mass dependence
1361: 	of the perturbative OPE coefficients of the unit operator.
1362: } 
1363: $D=2$, \ie, corrections in powers of $1/s_0$. 
1364: The leading $D=2$ corrections induced by the light-quark masses are
1365: computed using the running quark masses evaluated at the two-loop level.
1366: Evaluation of the contour integral in FOPT leads to~\cite{bnp}
1367: \beqn
1368:    \delta^{(2-\rm mass)}_{ud,V/A} 
1369:   &=&
1370:       -\,8\left(1+\frac{16}{3}a_s(s_0)\right)
1371:       \frac{m_u^2(s_0) + m_d^2(s_0)}{s_0} \nonumber \\
1372:   & & \pm\,
1373:        4\left(1+\frac{25}{3}a_s(s_0)\right)
1374:       \frac{m_u(s_0) m_d(s_0)}{s_0}~,
1375: \eeqn
1376: where $m_i(s_0)$ are the running quark masses evaluated at the scale 
1377: $s_0$ using the RGE $\gamma$-function~(\ref{eq:gammafun}).
1378: The following values are used by ALEPH for the renormalization group 
1379: invariant quark mass parameters $\hat{m}_i$ defined in~(\ref{eq:gammafunint})
1380: \beq
1381: \label{eq:qmasses}	
1382: 	\hat{m}_u = (8.7\pm1.5)\mev~,~~ \hat{m}_d = (15.4\pm1.5)\mev~,
1383: 	~~\hat{m}_s = (270\pm30)\mev~.
1384: \eeq
1385: 
1386: The dimension $D=4$ operators have dynamical contributions from the gluon 
1387: condensate $\GG$ and light $u,d$ quark condensates 
1388: $\langle m_iq_iq_j\rangle$, which are the matrix elements of the gluon field 
1389: strength-squared and the scalar quark densities, respectively. Remaining $D=4$ 
1390: operators are running quark masses to the fourth power. Inserting the Wilson 
1391: coefficients of these operators~\cite{becchi,generalis,broadhurst1981,gluonterm}
1392: in the contour integral one obtains~\cite{bnp}
1393: \beqn
1394: \label{eq:nondel4}
1395:    \delta_{ud,V/A}^{(4)} 
1396:    &=&
1397:        \frac{11}{4}\pi^2 a_s^2(s_0)
1398:        \frac{\GG}{s_0^2} \nonumber\\
1399:    & & +\, 16\pi^2
1400:            \left[1
1401:              +\frac{9}{2} a_s^2(s_0)
1402:            \right]\frac{\langle(m_u \mp m_d)
1403:                         \left(\ubar u \mp \dbar d\right)\rangle}
1404:                        {s_0^2} \nonumber \\
1405:    & & -\, 18\pi^2 a_s^2(s_0)
1406:            \left[\frac{\langle m_u\ubar u + m_d\dbar d\rangle}
1407:                       {s_0^2}
1408:                  + \frac{4}{9}\sum_{k=u,d,s}
1409:                  \frac{\langle m_k\qbar_kq_k\rangle}
1410:                       {s_0^2}
1411:            \right] \nonumber \\
1412:    & & -\, \left[\frac{48}{7}\frac{1}{a_s(s_0)}
1413:                  - \frac{22}{7}
1414:            \right]\frac{\left(m_u(s_0) \mp m_d(s_0)\right)
1415:                         \left(m_u^3(s_0) \mp m_d^3(s_0)\right)}
1416:                        {s_0^2} \nonumber \\
1417:    & & \pm\, 6\frac{m_u(s_0)m_d(s_0)\left(m_u(s_0) \mp m_d(s_0)\right)^2}
1418:                   {s_0^2}
1419:              +36\frac{m_u^2(s_0)m_d^2(s_0)}{s_0^2}~.
1420: \eeqn
1421: Where two signs are given the upper (lower) one is for $V$ ($A$). 
1422: The gluon condensate vanishes in first order \assz. However, there appear 
1423: second order terms in the Wilson coefficients due to the $s$ dependence of 
1424: $a_s$, which after integration becomes $a_s^2$.
1425: 
1426: The contributions from dimension $D=6$ operators are complex. 
1427: The most important operators arise from four-quark dynamical effects of 
1428: the form $\qbar_i\Gamma_1q_j\qbar_k\Gamma_2q_l$. Other operators, such 
1429: as the triple gluon condensate whose Wilson coefficient vanishes at order 
1430: \as, or those which are suppressed by powers of quark masses, are neglected 
1431: in the evaluation of the contour integrals performed in~\cite{bnp}.
1432: The large number of independent operators of the four-quark type occurring 
1433: in the $D=6$ term can be reduced by means of the {\em factorization} 
1434: (or {\em vacuum saturation}) assumption~\cite{svz} to leading order \as. 
1435: The operators are then expressed as products of scale dependent two-quark 
1436: condensates $\as(\mu)\langle\qbar_iq_i(\mu)\rangle\langle\qbar_jq_j(\mu)\rangle$.
1437: To take into account possible deviations from the vacuum saturation assumption, 
1438: one can introduce an effective scale independent operator 
1439: $\rho \as\langle\qbar q\rangle^2$ that replaces the above product. The 
1440: effective $D=6$ term obtained in this way is~\cite{bnp}
1441: \beqn
1442: \label{eq:nondel6}
1443:    \delta_{ud,V/A}^{(6)} 
1444:    &\simeq& \bigg(\!\begin{array}{c}7\\[-0.2cm]-11\end{array}\!\bigg)
1445:    \frac{256\pi^4}{27}\frac{\rho\as\langle\qbar q\rangle^2}
1446:                            {s_0^3}~,
1447: \eeqn
1448: predicting a different sign and hence a partial cancellation between 
1449: vector and axial-vector contributions\footnote
1450: {
1451: 	It has been pointed out in~\cite{pertmass} that 
1452: 	Eq.~(\ref{eq:nondel6}) assumes that the same $\rho$
1453: 	parameter can be used for both vector and axial-vector 
1454: 	contributions. 
1455: }
1456: 
1457: The dimension $D=8$ contribution has a structure of nontrivial quark-quark, 
1458: quark-gluon and four-gluon condensates whose explicit form is given 
1459: in~\cite{dimeight}. For the theoretical prediction of \Rtau\  it is custom
1460: to absorb the complete long and short distance part into the scale 
1461: invariant phenomenological $D=8$ operator $\langle{\cal O}_8\rangle$,
1462: which is fit simultaneously with \as\  and the other unknown 
1463: nonperturbative operators to date.
1464: 
1465: Higher order contributions from $D\ge10$ operators are expected to be small 
1466: as, equivalent to the gluon condensate, constant terms and terms in leading 
1467: order \as vanish in Eq.~(\ref{eq:contour}) after integrating over the contour.
1468: 
1469: \subsection{Results} \label{sec:moment}
1470: %
1471: It was shown in~\cite{pichledib} that one can exploit the shape of the 
1472: \sfs\  to obtain additional constraints on \assz and---more 
1473: importantly---on the nonperturbative effective operators. The 
1474: {\em $\tau$ spectral moments} at $s_0=m_\tau^2$ are defined by
1475: \beq
1476: \label{eq:moments}
1477:    R_{\tau,V/A}^{k\l} =
1478:        \intl_0^{m_\tau^2} ds\,\left(1-\frac{s}{m_\tau^2}\right)^{\!\!k}\!
1479:                               \left(\frac{s}{m_\tau^2}\right)^{\!\!\l}
1480:        \frac{dR_{\tau,V/A}}{ds}~,
1481: \eeq
1482: where $R_{\tau,V/A}^{00}=R_{\tau,V/A}$. The factor $(1-s/m_\tau^2)^k$ 
1483: suppresses the integrand at the crossing of the positive real axis where the 
1484: validity of the OPE less certain~\cite{braaten88} and the experimental accuracy 
1485: is statistically limited. Its counterpart $(s/m_\tau^2)^\l$ projects upon
1486: higher energies. The spectral information is used to fit simultaneously 
1487: \asm  and the effective operators $\GG$, 
1488: $\rho\as\langle\qbar q\rangle^2$ and $\langle{\cal O}_D\rangle$ for dimension
1489: $D=4$, $6$ and $8$, respectively. Due to the intrinsic experimental correlations 
1490: only five moments are used as input to the fit.
1491: 
1492: In analogy to \Rtau, the contributions to the moments originating from 
1493: perturbative and nonperturbative QCD are decomposed through the OPE
1494: \beq
1495: \label{eq:rtaumom}
1496:    R_{\tau,V/A}^{k\l} = 
1497:        \frac{3}{2}|V_{ud}|^2\Sew
1498:        \left(1 + \delta^{(0,k\l)} + 
1499:              \delta^\prime_{\rm EW} + 
1500:              \delta^{(2-{\rm mass},k\l)}_{ud,V/A} + 
1501:              \hm\hm\sum_{D=4,6,\dots}\hm\hm\delta_{ud,V/A}^{(D,k\l)}
1502:        \right)~.
1503: \eeq
1504: The prediction of the perturbative contribution takes the form
1505: \beq 
1506:    \delta^{(0,k\l)} = 
1507:        \sum_{n=1}^3 \tilde{K}_n(\xi) A^{(n,k\l)}(a_s)~,
1508: \eeq
1509: with the functions
1510: \beqn
1511: \label{eq:anmom}
1512:    A^{(n,k\l)}(a_s) 
1513:    &=&
1514:       \frac{1}{2\pi i}\hm\ointl_{|s|=m_\tau^2}\hm\hm
1515:       \frac{ds}{s}
1516:        \Bigg[2\Gamma(3 + k)
1517:              \left(\frac{\Gamma(1 + \l)}{\Gamma(4 + k + \l)} +
1518:                    2\frac{\Gamma(2 + \l)}{\Gamma(5 + k + \l)} 
1519:              \right) \nonumber\\[0.2cm]
1520:    &&\hspace{1.4cm}
1521:              -\; I\left(\frac{s}{s_0},1+\l,3+k\right) 
1522:              - 2 I\left(\frac{s}{s_0},2+\l,3+k\right)
1523:             \Bigg]
1524: 	     a_s^n(-\xi s)~,
1525: \eeqn
1526: which make use of the elementary integrals 
1527: $I(\gamma,a,b)=\int_0^\gamma t^{a-1}(1-t)^{b-1}dt$.
1528: The contour integrals are numerically solved for the running 
1529: $a_s(-\xi s)$ (CIPT).
1530: 
1531: In the chiral limit\footnote
1532: {
1533: 	In the chiral limit, vector and axial-vector currents are 
1534: 	conserved so that $s\Pi_V^{(0)}=s\Pi_A^{(0)}=0$.		
1535: } 
1536: and neglecting the small logarithmic $s$ dependence of the 
1537: Wilson coefficients, the dimension $D$ nonperturbative contributions 
1538: in Expression~(\ref{eq:rtaumom}) read
1539: \beq
1540: \label{eq:array}
1541:    \delta_{ud,V/A}^{(D,k\l)} =
1542:       8 \pi^2 \left({\scriptsize\begin{array}{cccccc} 
1543:                (D=2) & (D=4) & (D=6) & (D=8) & (D=10)& (k,\l) \\
1544:                   1  &  0    & -3    &  -2   &   0   & (0,0) \\
1545:                   1  &  1    & -3    &  -5   &  -2   & (1,0) \\
1546:                   0  & -1    & -1    &   3   &   5   & (1,1) \\
1547:                   0  &  0    &  1    &   1   &  -3   & (1,2) \\
1548:                   0  &  0    &  0    &  -1   &  -1   & (1,3)
1549:               \end{array}}\right)
1550:       \sum_{{\rm dim}{\cal O}=D}\hm\hm C^{(1+0)}(\mu)
1551:             \frac{\langle{\cal O}_D(\mu)\rangle_{V/A}}
1552:                  {m_\tau^D}~,
1553: \eeq
1554: where the matrix is defined by the choice of the coefficients for the 
1555: moments $k=1$, $\l=0,1,2,3$ and the corresponding dimension $D$. For 
1556: completeness, we also give the coefficients for dimension $D=10$, which 
1557: is not considered by the experiments since they do not contribute to 
1558: $R_{\tau,V/A}^{00}$ at leading order. With increasing weight $\l$ the 
1559: contributions from low dimensional operators vanish. For example, the 
1560: only nonperturbative contribution to the moment $R_{\tau,V/A}^{13}$ stems 
1561: from dimension $D=8$ and beyond. Hence, in a fit using spectral 
1562: moments, $D=8$ will be determined by $R_{\tau,V/A}^{13}$. This 
1563: observation is in some sense a ``paradox'', as higher moments project 
1564: upon higher masses on the contrary to \Rtau\ and the spirit of the OPE, 
1565: where the higher dimension terms are enhanced at small $s_0$. 
1566: 
1567: For practical purpose it is more convenient to define moments 
1568: that are normalized to the corresponding \RtauVA\ in order 
1569: to decouple the normalization from the shape of the $\tau$ \sfs
1570: \beq
1571: \label{eq:dkl}
1572:    D_{\tau,V/A}^{k\l} =
1573:      \frac{R_{\tau,V/A}^{k\l}}{R_{\tau,V/A}}~.
1574: \eeq
1575: The two sets of experimentally almost uncorrelated 
1576: observables---\RtauVA\  on one hand and the spectral moments on the other 
1577: hand---yield independent constraints on \asm and thus provide an important 
1578: test of consistency. The correlation between these observables is completely
1579: negligible in the $V+A$ case where \RtauVpA\ is calculated from the 
1580: difference $R_\tau-R_{\tau,S}$, which has no correlation with the hadronic 
1581: invariant mass spectrum. One experimentally obtains the $D_{\tau,V/A}^{k\l}$ 
1582: by integrating the weighed normalized invariant mass-squared spectrum. 
1583: The corresponding theoretical predictions are easily modified. 
1584: 
1585: %
1586: % ----------------- Fit pf alpha_s ==== results ------------------------
1587: %
1588: 
1589: \subsubsection{The ALEPH determination of $\as(m_\tau^2)$ and 
1590:                nonperturbative contributions}
1591: \label{sec:qcd_as_fit}
1592: 
1593: Combined fits to experimental spectral moments and the extraction of
1594: $\as(m_\tau^2)$ together with the leading nonperturbative operators have 
1595: been performed by ALEPH, CLEO and 
1596: OPAL~\cite{aleph_as,aleph_asf,aleph_taubr,cleo_as,opal_vasf} using 
1597: similar strategies and inputs.
1598: Let us follow in this report the most recent analysis performed by the
1599: ALEPH collaboration~\cite{aleph_taubr}.
1600: 
1601: Since the determination of $\as$ should be model-independent, the experiments
1602: proceed by fitting simultaneously the nonperturbative operators, which 
1603: is possible since the correlations between these and $\as$ turn out to be 
1604: small enough. The theoretical framework and its intrinsic uncertainties were 
1605: the subject of the previous section. Minor additional contributions originate 
1606: from the CKM matrix element $|V_{ud}|$, the electroweak radiative correction 
1607: factor $\Sew$, the light quark masses $m_{u,d}$ and the quark condensates 
1608: (Section~\ref{sec:nonpert}). The largest contributions
1609: have their origin in the truncation of the perturbative expansion.
1610: Although it introduces some double-counting for the systematic error, the 
1611: procedure used in~\cite{aleph_taubr} considers separate variations of 
1612: the unknown higher order coefficient $K_4$ and the renormalization scale. 
1613: The renormalization scale is varied around $m_\tau$ from 1.1 to $2.5\gev$
1614: with the variation over half of the range taken as systematic uncertainty.
1615: Taking advantage of the new theoretical developments discussed above, 
1616: the value $K_4 = 25 \pm 25$ is used in \cite{aleph_taubr}.
1617: 
1618: \begin{table}[t]
1619:   \caption[.]{\label{tab_asresults}
1620: 	Fit results~\cite{aleph_taubr} for \asm and the nonperturbative 
1621: 	contributions for vector, axial-vector and $V+A$ combined
1622: 	fits using the corresponding experimental spectral moments 
1623: 	as input parameters. Where two errors are given the first is
1624: 	experimental and the second theoretical. The $\delta^{(2)}$ term 
1625: 	is theoretical only with quark masses varying within their 
1626: 	allowed ranges (see text). The quark condensates in the $\delta^{(4)}$ 
1627: 	term are obtained from PCAC, while the gluon condensate is 
1628: 	determined by the fit. The total nonperturbative contribution is 
1629: 	the sum $\delta_{\rm NP}=\delta^{(4)}+\delta^{(6)}+\delta^{(8)}$.
1630: 	Full results are listed only for the \FOPTCI\
1631: 	perturbative prescriptions, except for \asm\ where both \FOPTCI\
1632:         and FOPT results are given.}
1633:   \begin{center}
1634: {\small
1635: \setlength{\tabcolsep}{0.0pc}
1636: \begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lccc} 
1637: \hline\noalign{\smallskip}
1638:   Parameter     &Vector ($V$) &Axial-Vector ($A$)&  $V\,+\,A$\\
1639: \noalign{\smallskip}\hline\noalign{\smallskip}
1640:  \asm  (\FOPTCI) &  $0.355\pm0.008\pm0.009$  
1641:                  &  $0.333\pm0.009\pm0.009$   
1642:                  &  $0.350\pm0.005\pm0.009$   
1643: \\
1644:  \asm  (FOPT)    &  $0.331\pm0.006\pm0.012$  
1645:                  &  $0.327\pm0.007\pm0.012$   
1646:                  &  $0.331\pm0.004\pm0.012$ \\
1647: \noalign{\smallskip}\hline\noalign{\smallskip}
1648:   $\delta^{(2)}$ (\FOPTCI)    & $(-3.3\pm3.0) \times 10^{-4}$
1649:                               & $(-5.1\pm3.0) \times 10^{-4}$
1650:                               & $(-4.4\pm2.0) \times 10^{-4}$
1651:  \\
1652: %  $\delta^{(2)}$ (FOPT)       & $(-3.0\pm3.0) \times 10^{-4}$
1653: %                              & $(-5.0\pm3.0) \times 10^{-4}$
1654: %                              & $(-4.0\pm2.0) \times 10^{-4}$ \\
1655: \noalign{\smallskip}\hline\noalign{\smallskip}
1656:  $\GG$ ($\gev^4$) (\FOPTCI)   &  $(0.4\pm0.3) \times 10^{-2}$  
1657:                                 &  $(-1.3\pm0.4) \times 10^{-2}$   
1658:                                 &  $(-0.5\pm0.3) \times 10^{-2}$   
1659: \\ 
1660: % $\GG$ ($\gev^4$) (FOPT)    &  $(1.5\pm0.3) \times 10^{-2}$  
1661: %                              &  $(-0.2\pm0.4) \times 10^{-2}$   
1662: %                              &  $(\ph{-}0.6\pm0.2) \times 10^{-2}$   \\
1663: \noalign{\smallskip}\hline\noalign{\smallskip}
1664:   $\delta^{(4)}$ (\FOPTCI)    & $(4.1\pm1.2) \times 10^{-4}$
1665:                               & $(-5.7\pm0.1) \times 10^{-3}$
1666:                               & $(-2.7\pm0.1) \times 10^{-3}$
1667:  \\      
1668: %  $\delta^{(4)}$ (FOPT)       & $(6.8\pm1.0) \times 10^{-4}$
1669: %                              & $(-5.3\pm0.1) \times 10^{-3}$
1670: %                              & $(-2.4\pm0.1) \times 10^{-3}$ \\
1671: \noalign{\smallskip}\hline\noalign{\smallskip}
1672:  $\delta^{(6)}$ (\FOPTCI) &  $(2.85\pm0.22) \times 10^{-2}$  
1673:                           &  $(-3.23\pm0.26) \times 10^{-2}$   
1674:                           &  $(-2.1\pm2.2) \times 10^{-3}$   
1675: \\
1676: % $\delta^{(6)}$ (FOPT)    &  $(2.70\pm0.25) \times 10^{-2}$  
1677: %                          &  $(-2.96\pm0.31) \times 10^{-2}$   
1678: %                          &  $(-1.6\pm2.5) \times 10^{-3}$     \\
1679: \noalign{\smallskip}\hline\noalign{\smallskip}
1680:  $\delta^{(8)}$ (\FOPTCI) &  $(-9.0\pm0.5) \times 10^{-3}$  
1681:                           &  $(8.9\pm0.6) \times 10^{-3}$   
1682:                           &  $(-0.3\pm4.8) \times 10^{-4}$   
1683: \\
1684: % $\delta^{(8)}$ (FOPT)    &  $(-8.6\pm0.6) \times 10^{-3}$  
1685: %                          &  $(8.6\pm0.6) \times 10^{-3}$   
1686: %                          &  $(\ph{-}1.2\pm5.2) \times 10^{-4}$     \\
1687: \noalign{\smallskip}\hline\noalign{\smallskip}
1688:   Total $\delta_{\rm NP}$ (\FOPTCI)    & $(1.99\pm0.27) \times 10^{-2}$  
1689:                                        & $(-2.91\pm0.20) \times 10^{-2}$
1690:                                        & $(-4.8\pm1.7) \times 10^{-3}$
1691:  \\
1692: %  Total $\delta_{\rm NP}$ (FOPT)       & $(1.91\pm0.31) \times 10^{-2}$  
1693: %                                       & $(-2.63\pm0.25) \times 10^{-2}$
1694: %                                       & $(-3.9\pm2.0) \times 10^{-3}$ \\
1695: \noalign{\smallskip}\hline\noalign{\smallskip}
1696:  $\chi^2/$DF (\FOPTCI)       & 0.52            & 4.97            & 3.66    
1697:  \\ 
1698: % $\chi^2/$DF (FOPT)          & 0.01            & 0.63            & 0.11  \\
1699: \noalign{\smallskip}\hline
1700:   \end{tabular*}
1701: }
1702:   \end{center}
1703: \end{table}
1704: 
1705: %\begin{table}[t]
1706: %  \caption[.]{\label{tab_rescorr}
1707: %        Correlation matrices for the fits given in Table~\ref{tab_asresults} 
1708: %	for vector (left table), axial-vector (middle) and $V+A$ (right table) 
1709: %	using \FOPTCI~\cite{aleph_taubr}. }
1710: %{\footnotesize
1711: %\begin{center}
1712: %\setlength{\tabcolsep}{0.6pc}
1713: %\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}l|ccc|ccc|ccc} 
1714: %\hline\noalign{\smallskip}
1715: %Parameter &  
1716: %$\GG_V$ &  $\delta^{(6)}_V$ & $\delta^{(8)}_V$ &
1717: %$\GG_A$ &  $\delta^{(6)}_A$ & $\delta^{(8)}_A$ &
1718: %$\GG_{V+A}$
1719: %                     &  $\delta^{(6)}_{V+A}$ & $\delta^{(8)}_{V+A}$ \\[0.15cm]
1720: %\hline
1721: %&&&&&&&&& \\[-0.3cm]
1722: %\asm          & $-0.39$  & $-0.28$  & $-0.34$
1723: %              & $-0.57$  & $\ph{-}0.52$  & $-0.55$   
1724: %              & $-0.37$  & $\ph{-}0.38$  & $-0.45$ \\
1725: %$\GG$
1726: %              &  $   1$  & $\ph{-}0.44$  &  $\ph{-}0.46$
1727: %              &  $   1$  & $-0.81$  &  $\ph{-}0.80$   
1728: %              &  $   1$  & $-0.65$  &  $\ph{-}0.65$ \\
1729: %
1730: %$\delta^{(6)}$ 
1731: %              &     --   &  $   1$  & $-0.98$
1732: %              &     --   &  $   1$  & $-0.99$   
1733: %              &     --   &  $   1$  & $-0.98$ \\
1734: %$\delta^{(8)}$ 
1735: %              &     --   &     --   &  $   1$
1736: %              &     --   &     --   &  $   1$
1737: %              &     --   &     --   &  $   1$ \\
1738: %\noalign{\smallskip}\hline
1739: %  \end{tabular*}
1740: %\end{center}
1741: %}
1742: %\end{table}
1743: The fit minimizes the $\chi^2$ of the differences between measured 
1744: and adjusted quantities contracted with the inverse of the sum of the 
1745: experimental and theoretical covariance matrices. 
1746: The results of~\cite{aleph_taubr} are listed in Table~\ref{tab_asresults}. 
1747: % Table~\ref{tab_rescorr} gives the corresponding correlation matrices
1748: % of the fit parameters. 
1749: %The limited number of observables
1750: %and the strong correlations between the spectral moments explain the
1751: %large correlations observed, especially between the
1752: %nonperturbative operators, which are solely determined by the shapes 
1753: %of the spectral functions. 
1754: The precision of \asm  obtained with the 
1755: two perturbative methods employed is comparable, however their central
1756: values differ by  0.01--0.02, which is larger than the theoretical errors
1757: assigned. The $\delta^{(2)}$ term is theoretical only with quark masses 
1758: varying within their allowed ranges~(\ref{eq:qmasses}). The quark 
1759: condensates in the $\delta^{(4)}$ term are obtained from PCAC, 
1760: while the gluon condensate is determined by the fit. The main contributions
1761: to the theoretical uncertainty on \asm are listed in Table~\ref{tab_theoerr}.
1762: The fact that the $\chi^2$ value of the axial-vector fit is better
1763: with FOPT than with CIPT should not---in our view--- affect the 
1764: conclusion in favor of the latter method, reached in 
1765: Sections~\ref{sec:pert_comp} and \ref{sec:cipt} on the basis of
1766: a better perturbative behavior, but it may require further studies.
1767: \begin{table}[t]
1768:   \caption[.]{\label{tab_theoerr}
1769:               Sources of theoretical uncertainties and their impact 
1770:               on $R_{\tau,V+A}$ and \asm~\cite{aleph_taubr}.}
1771:   \begin{center}
1772: \setlength{\tabcolsep}{0.0pc}
1773: \begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lccccc} 
1774: \hline\noalign{\smallskip}
1775:  &     & \mc{2}{c}{$\Delta^{\rm th} R_{\tau,V+A}$} 
1776:        & \mc{2}{c}{$\Delta^{\rm th}$\asm} \\
1777: \rs{Error source}  & \rs{Value\pms$\Delta^{\rm th}$}
1778:        & \FOPTCI & FOPT
1779:        & \FOPTCI & FOPT \\ 
1780: \noalign{\smallskip}\hline\noalign{\smallskip}
1781:  $\Sew$          & $1.0198\pm0.0006$              & \mc{2}{c}{0.002}
1782:                                                     & \mc{2}{c}{0.001} \\
1783:  $|V_{ud}|$      & $0.9745\pm0.0004$              & \mc{2}{c}{0.002} 
1784:                                                     & \mc{2}{c}{0.001} \\
1785:  $K_4$           & $25\pm25$                      & 0.015 & 0.009
1786:                                                     & 0.007 & 0.003 \\
1787:  R-scale $\mu$ & $m_\tau^2 \pm 2\gev^2$           & 0.012 & 0.035
1788:                                                     & 0.006 & 0.011 \\
1789: \noalign{\smallskip}\hline\noalign{\smallskip}
1790:  \mc{2}{c}{Total errors}                        & 0.019 & 0.036
1791:                                                       & 0.009 & 0.012\\ 
1792: \noalign{\smallskip}\hline
1793:   \end{tabular*}
1794:   \end{center}
1795: \end{table}
1796: 
1797: The final result on \asm  given by~\cite{aleph_taubr} is taken as the arithmetic 
1798: average of the \FOPTCI\  and FOPT values given in Table~\ref{tab_asresults}, 
1799: with half of their difference added as additional theoretical error
1800: $\as(m_\tau^2) = 0.340 \pm 0.005_{\rm exp} \pm 0.014_{\rm th}$.
1801: The first error accounts for the experimental uncertainty and the 
1802: second error gives the uncertainty related to the theoretical prediction 
1803: of the spectral moments including the ambiguity between \FOPTCI\  and FOPT.
1804: 
1805: There is a remarkable agreement within statistical errors between the 
1806: \asm  determinations using the vector and axial-vector data.
1807: This provides an important consistency check of the results, since
1808: the two corresponding spectral functions are experimentally independent and
1809: manifest a quite different resonant behavior. 
1810: %The results are displayed
1811: %in Fig.~\ref{alphas_res}.
1812: %\begin{figure}[t]
1813: %   \centerline{\psfig{file=figures/alphas_res.eps,width=78mm}}
1814: %  \caption[.]{\label{alphas_res}
1815: %              Results for \asm from the fits of $R_{\tau,V,A,V+A}$ and 
1816: %              the moments $D^{k\l}_{V,A,V+A}$ using the \FOPTCI and FOPT
1817: %              perturbative expansions. The measurements are correlated due 
1818: %              to the theoretical errors (see Table~\ref{tab_asresults}).}
1819: %\end{figure}
1820: 
1821: The advantage of 
1822: separating the vector and axial-vector channels and comparing to
1823: the inclusive $V+A$ fit becomes obvious in the adjustment of the 
1824: leading nonperturbative contributions of dimension $D=6$ and $D=8$, 
1825: which approximately cancel in the inclusive sum. This cancellation of the 
1826: nonperturbative terms increases the confidence in the \asm determination 
1827: from the inclusive $V+A$ observables. The gluon condensate is 
1828: determined by the first $k=1$, $\l=0,1$ moments, which receive 
1829: lowest order contributions. It is observed that the values obtained in 
1830: the $V$ and $A$ fits are not very consistent, which could indicate 
1831: problems in the validity of the OPE approach used once the 
1832: nonperturbative terms become significant. Taking the value obtained in 
1833: the $V+A$ fit, where nonperturbative effects are small, 
1834: and adding as systematic uncertainties half of the difference between 
1835: the vector and axial-vector fits as well as between 
1836: the \FOPTCI\ and FOPT results, ALEPH measures the gluon condensate to be
1837: \beq
1838: \label{eq:gluon_cond}
1839: 	\GG=(0.001\pm0.012)\gev^4.
1840: \eeq
1841: This result does not provide evidence for a nonzero gluon condensate, but it 
1842: is consistent with and has comparable accuracy to the independent value 
1843: obtained using charmonium sum rules and \ee data in the charm 
1844: region, $(0.011\pm0.009)\gev^4$ in a combined determination with 
1845: the $c$ quark mass~\cite{ioffe}. 
1846: 
1847: The $D=6,8$ nonperturbative contributions are obtained after averaging the
1848: FOPT and \FOPTCI\ values:
1849: \beq
1850: \begin{array}{rclrcl}
1851:   \delta^{(6)}_V	&=& (2.8 \pm 0.3) \times 10^{-2}~, &
1852:   \delta^{(8)}_V 	&=& (-8.8 \pm 0.6) \times 10^{-3}~,\\
1853:   \delta^{(6)}_A 	&=& (-3.1 \pm 0.3) \times 10^{-2}~, & 
1854:   \delta^{(8)}_A 	&=& (8.7 \pm 0.6) \times 10^{-3}~,  \\
1855:   \delta^{(6)}_{V+A} 	&=& (-1.8 \pm 2.4) \times 10^{-3}~,  &
1856:   \delta^{(8)}_{V+A} 	&=& (0.5 \pm 5.1) \times 10^{-4}~. 
1857: \end{array}
1858: \eeq
1859: The remarkable feature is the approximate cancellation of these contributions
1860: in the $V+A$ case, both for $D=6$ and $D=8$. This property was 
1861: predicted~\cite{bnp} for $D=6$ using the simplifying assumption of vacuum
1862: saturation for the matrix elements of four-quark operators, yielding
1863: $\delta^{(6)}_V /\delta^{(6)}_A =-7/11=-0.64$, in fair agreement with
1864: the above results which reads $-0.90\pm0.18$. The estimate~\cite{bnp} for 
1865: $\delta^{(6)}_V = (2.5 \pm 1.3) \times 10^{-2}$ agrees well with 
1866: the experimental result. 
1867: 
1868: The total nonperturbative $V+A$ correction, 
1869: $\delta_{{\rm NP},V+A}=(-4.3\pm1.9) \times 10^{-3}$, is an
1870: order-of-magnitude smaller than the corresponding values in the $V$ and 
1871: $A$ components, $\delta_{{\rm NP},V}=(2.0\pm0.3) \times 10^{-2}$ and 
1872: $\delta_{{\rm NP},A}=(-2.8\pm0.3) \times 10^{-2}$.
1873: 
1874: \subsubsection{The OPAL results}
1875: 
1876: The OPAL analysis~\cite{opal_vasf} makes use of the spectral functions 
1877: shown in Fig.~\ref{fig:vasf} and provides also a complete description 
1878: in terms of QCD of their vector and axial-vector data. It proceeds 
1879: along very similar lines as the ALEPH analysis, with the difference 
1880: that OPAL has chosen to quote their results separately for three 
1881: prescriptions for the perturbative expansion, respectively
1882: \FOPTCI, FOPT and with renormalon chains, {\it i.e.}
1883: \beq
1884: \label{qcd_opal}
1885: \asm =\left\{\begin{array}{ll}
1886: 0.348\pm 0.009_{\rm exp}\pm 0.019_{\rm th} & \hspace{5mm}{\rm CIPT}~, \\
1887: 0.324\pm 0.006_{\rm exp}\pm 0.013_{\rm th} & \hspace{5mm}{\rm FOPT}~, \\
1888: 0.306\pm 0.005_{\rm exp}\pm 0.011_{\rm th} & \hspace{5mm}{\rm renormalon}~.
1889: \end{array}\right.
1890: \eeq
1891: 
1892: The OPAL results using \FOPTCI and FOPT agree with those from ALEPH.
1893: The quoted theoretical uncertainties are consistently treated between 
1894: the two analyses , with the important difference that the more recent
1895: ALEPH analysis~\cite{aleph_taubr} uses a reduced uncertainty range for 
1896: the $K_4$ parameter ($K_4=25\pm25$) as compared to $K_4=50\pm50$ in the 1998 
1897: analysis~\cite{aleph_asf} and $K_4=25\pm50$ in OPAL. The justification
1898: of the smaller $K_4$ uncertainty was provided in Section~\ref{sec:pert}.
1899: As for the value obtained with the renormalon chains, we refer to our
1900: discussion in Sections~\ref{sec:renormalons} and \ref{sec:pert_comp} 
1901: where we concluded that however interesting this approach is, its
1902: performance for a reliable and precise determination of \asm is
1903: not adequate. OPAL also derives values for the gluon condensate: averaging
1904: the results of the \FOPTCI and FOPT analyses, one obtains
1905: $ \GG=(0.005\pm0.017)\gev^4$, in agreement with ALEPH.
1906: 
1907: \subsubsection{Running of $\as(s)$ below $m_\tau^2$}
1908: %\subsubsection{Measuring the running of $\as(\mu^2)$ for $\mu<m_\tau$}
1909: \label{sec:qcd_as_running}
1910: %
1911: Using the \sfs, one can simulate the physics of a hypothetical 
1912: $\tau$ lepton with a mass $\sqrt{s_0}$ smaller than $m_\tau$
1913: through Eq.~(\ref{eq:rtauth1}). Assuming quark-hadron duality, 
1914: the evolution of $R_\tau(s_0)$ provides a direct test of the running 
1915: of $\as(s_0)$, governed by the RGE $\beta$-function. On the other 
1916: hand, it is also a test of the stability of the OPE approach 
1917: in $\tau$ decays. The studies performed in this section employ only 
1918: \FOPTCI. Results obtained with FOPT are similar\footnote
1919: {
1920: 	As already indicated by the smaller $\chi^2$ value of the 
1921: 	spectral moments fit (see Table~\ref{tab_asresults}), the 
1922: 	agreement between the FOPT evolution computation and the 
1923: 	data is better than it is when using \FOPTCI.
1924: } 
1925: and differ mostly in the central \asm value.
1926: \begin{figure}[t]
1927:    \centerline{
1928:         \epsfxsize8.3cm\epsffile{figures/va_runrtau.eps}
1929:         \epsfxsize8.3cm\epsffile{figures/va_runas.eps}}
1930:   \vspace{-0.2cm}
1931:   \caption[.]{\label{vpa_runrtauas}
1932:         \underline{Left:} 
1933: 	The ratio $R_{\tau,V+A}$ versus the square ``$\tau$ mass'' $s_0$.
1934:       	The curves are plotted as error bands to emphasize their 
1935:        	strong point-to-point correlations in $s_0$. Also 
1936:       	shown is the theoretical prediction using \FOPTCI\ and
1937:        	the results for $R_{\tau,V+A}$ and the nonperturbative 
1938:        	terms from Table~\rm\ref{tab_asresults}.
1939:         \underline{Right:} 
1940: 	The running of $\as(s_0)$ obtained from the 
1941:        	fit of the theoretical prediction to $R_{\tau,V+A}(s_0)$ using
1942:        	CIPT. The shaded band shows the data including only experimental
1943:        	errors. The curve gives the four-loop RGE evolution 
1944:        	for three flavors.}
1945: \end{figure}
1946: 
1947: The functional dependence of $R_{\tau,V+A}(s_0)$ is plotted in the left
1948: hand plot of Fig.~\ref{vpa_runrtauas} together with the theoretical 
1949: prediction using the results of Table~\ref{tab_asresults}. 
1950: The spreads due to uncertainties are shown as error bands. 
1951: The correlations between two adjacent points in $s_0$ are large as 
1952: the only new information is provided by the small mass difference 
1953: between the two points and the slightly modified weight functions. 
1954: Moreover they are reinforced by the original experimental and theoretical 
1955: correlations. Below $1\gev^2$ the error of the theoretical 
1956: prediction of $R_{\tau,V+A}(s_0)$ starts to blow up due to the 
1957: increasing uncertainty from the unknown $K_4$ perturbative term;
1958: errors of the nonperturbative contributions are {\it not} contained 
1959: in the theoretical error band. Figure~\ref{vpa_runrtauas} (right) shows the plot 
1960: corresponding to Fig.~\ref{vpa_runrtauas} (left), translated into the running 
1961: of $\as(s_0)$, \ie, the experimental value for $\as(s_0)$ 
1962: has been individually determined at every $s_0$ from the comparison 
1963: of data and theory. Also plotted is the four-loop RGE evolution using 
1964: three quark flavors.
1965: 
1966: It is remarkable that the theoretical prediction using the parameters 
1967: determined at the $\tau$ mass and $R_{\tau,V+A}(s_0)$ extracted from 
1968: the measured
1969: $V+A$ \sf\ agree down to $s_0 \sim 0.8\gev^2$. The agreement is good to
1970: about 2\% at $1\gev^2$. This result, even more directly illustrated by the
1971: right hand plot of Fig.~\ref{vpa_runrtauas}, demonstrates the validity of 
1972: the perturbative approach
1973: down to masses around $1\gev$, well below the $\tau$ mass scale. The
1974: agreement with the expected scale evolution between 1 and $1.8\gev$ is an
1975: interesting result, considering the relatively low mass range, where
1976: $\as$ is seen to decrease by a factor of 1.6 and reaches rather 
1977: large values $\sim 0.55$ at the lowest masses. This behavior provides
1978: confidence that the \asm  measurement is on solid phenomenological ground.
1979: 
1980: \subsubsection{Final assessment on the $\as(m_\tau^2)$ determination}
1981: \label{sec:assess}
1982: 
1983: Although the recent ALEPH evaluation of $\as(m_\tau^2)$ represents the
1984: state-of-the art, several remarks can be made:
1985: \begin{itemize}
1986: 
1987: \item 	The analysis is based on the ALEPH spectral functions and branching
1988: 	fractions, ensuring a good consistency between all the observables, 
1989:         but not exploiting the full experimental information currently 
1990:         available from other experiments. Since the result on $\as(m_\tau^2)$ 
1991:         is limited by theoretical uncertainties, one should expect only a 
1992:         small improvement of the final error in this way, however it can 
1993:         influence the central value.
1994: 
1995: \item 	One example for this is the evaluation of the strange component. Some 
1996: 	discrepancy is observed between the ALEPH measurement of the 
1997: 	$(K \pi \pi)^- \nu$ mode and the CLEO and OPAL results, as discussed 
1998: 	in Section~\ref{sec:brs_k}. Although this could still be the
1999: 	result of a statistical fluctuation, their average provides a 
2000: 	significant shift in the central value compared to using the ALEPH 
2001: 	number alone.
2002: 	Another improvement is the substitution of the measured branching
2003: 	fraction for the $\Km\nu$ mode by the more precise value predicted from
2004: 	$\tau$--$\mu$ universality (see Eq.~(\ref{k_uni})). Both operations 
2005: 	have the effect to increase the strange $\RtauS$ ratio from  
2006: 	$0.1603 \pm 0.0064$, as obtained by ALEPH, to $0.1686 \pm 0.0047$
2007: 	for the world average.
2008: 
2009: \item 	One can likewise substitute the world average value for the 
2010: 	universality-consolidated value of the electronic branching fraction 
2011:         given in Eq.~(\ref{eq:uni_be}), $\BR_e^{\rm uni}=(17.818 \pm 0.032)\%$,
2012:         to the corresponding
2013: 	ALEPH result, $\BR_e=(17.810 \pm 0.039)\%$, with little effect
2014: 	on the central value, but some improvement in the precision.
2015: 
2016: \item 	Most of the theoretical uncertainty originates from the limited 
2017: 	knowledge of the perturbative expansion, only predicted to third order.
2018: 	This problem was discussed in detail in Section~\ref{sec:pert} and, 
2019: 	among the standard FOPT and \FOPTCI\  approaches, strong arguments 
2020: 	have been presented in favor of the latter one. 
2021:         We therefore take the result from the CIPT expansion, 
2022:         not introducing any additional 
2023: 	uncertainty spanning the difference between FOPT and CIPT results. The 
2024: 	dominant theoretical errors are from the uncertainty in $K_4$ and from 
2025: 	the renormalization scale dependence, both covering the effect of 
2026: 	truncating the series after the estimated fourth order.
2027: 
2028: \end{itemize}
2029: 
2030: From this analysis, one finds the new value for the nonstrange ratio,
2031: \beqn
2032: \label{eq:rtau_new}
2033:   R_{\tau,V+A} 	&=& R_\tau - R_{\tau,S} \nonumber \\
2034:        	        &=& (3.640 \pm 0.010)~-~(0.1686 \pm 0.0047) \nonumber \\
2035:                 &=& 3.471 \pm 0.011~, 
2036: \eeqn
2037: to be compared to the ALEPH value of $3.482 \pm 0.014$. The 
2038: result~(\ref{eq:rtau_new}) translates into an updated determination 
2039: of $\as(m_\tau^2)$ from the inclusive $V+A$ component using the \FOPTCI\  
2040: approach
2041: \beq
2042: \label{astau:final}
2043:   \as(m_\tau^2) = 0.345 \pm 0.004_{\rm exp} \pm 0.009_{\rm th}~,
2044: \eeq
2045: with improved experimental and theoretical precision over the ALEPH
2046: result.
2047: 
2048: \subsubsection{Evolution to $M_Z^2$}
2049: \label{sec:qcd_as_evolution}
2050: 
2051: It is customary to compare \as values, obtained at different 
2052: renormalization scales, at the scale of the $Z$-boson mass. 
2053: To do this, one evolves the \as result using the RGE~(\ref{eq:betafun}).
2054: A difficulty arises from the quark thresholds: 
2055: in the MS scheme, or any of its modifications such as \MSbar,
2056: the beta function governing the running of the strong coupling constant
2057: is independent of quark masses.
2058: This simplifies the calculation of QCD corrections beyond the one-loop 
2059: level. On the other hand, decoupling of heavy quarks~\cite{decoupling} 
2060: is not manifest in each order of perturbation theory~\cite{bernr}. 
2061: A heavy quark decoupling is realized in momentum-space 
2062: dependent renormalization schemes (MO). The RGE in an MO scheme 
2063: involves the scaling-function 
2064: $\beta_{\rm MO}=\beta(\as^\prime,m_i^\prime,a^\prime)$
2065: that depends on the coupling $\as^\prime(\mu^2)$, the quark masses 
2066: $m_i^\prime(\mu^2)$ and a gauge parameter $a^\prime(\mu^2)$, where the 
2067: primes stand for the scheme dependence of the parameters. Quark-loop 
2068: calculations in this scheme appear rather complicated, however the 
2069: mass dependence of the $\beta$-function provides a suppression of 
2070: heavy-quark effects at scales much smaller than the masses of these 
2071: quarks, \ie, it decouples heavy quarks from the light-particle Green's
2072: function to each order in perturbation theory.
2073: 
2074: To obtain decoupling in MS schemes, one builds in the decoupling region, 
2075: $\mu\ll \mbar_q^{(n_f)}(\mu^2)$, where $\mbar^{(n_f)}(\mu^2)$ is the 
2076: running \MSbar mass of the 
2077: heavy quark with flavor $f$, an effective field theory that behaves as 
2078: if only light quarks up to flavor $n_f-1$ were present. Matching conditions
2079: connect the parameters of the low energy effective Lagrangian to the 
2080: full theory. The coupling constant of the effective theory can then 
2081: be developed in a power series of the coupling constant of the full 
2082: theory with coefficients that depend on $x=\ln(\mu^2/\mbar_q^2)$.
2083: Doing so one obtains for the matching of $a_s$ and the light quark 
2084: masses between the ``light'' flavor $n_f-1$ and the heavy-quark flavor 
2085: $n_f$~\cite{betafourloop,chet1,chet2,wetzel1,wetzel2,pichsanta}
2086: \beqn
2087:    a_s^{(n_f)}(\mu^2) 
2088:      &=&
2089:         a_s^{(n_f-1)}(\mu^2)
2090:         \left[1 + \sum_{k=1}^\infty C_k(x)\left(a_{s}^{(n_f-1)}(\mu^2)\right)^k\right]~, \nonumber\\
2091:    \mbar_q^{(n_f)}(\mu^2) 
2092:      &=&
2093: \label{eq:CHfunctions}
2094:         \mbar_q^{(n_f-1)}(\mu^2)
2095:         \left[1 + \sum_{k=1}^\infty H_k(x)\left(a_{s}^{(n_f-1)}(\mu^2)\right)^k\right]~.
2096: \eeqn
2097: There is no straightforward choice of the matching scale $\mu$. However,
2098: to acquire good convergence of the perturbative expansion, one should
2099: satisfy $\mu/\mbar_q\sim \mathcal{O}(1)$. Since vacuum polarization of
2100: quark pairs modifies the coupling constant, we believe that the scale
2101: $\mu=2\mbar_q$ is quite meaningful. The effect from this scale ambiguity 
2102: is part of the systematic error assigned to the RGE evolution (see below).
2103: 
2104: The functions $C_k(x)$ and $H_k(x)$ are
2105: derived by inserting the RGE-based Taylor expansions (Eq.~(\ref{eq:astaylor}) 
2106: and equivalent for the quark masses)
2107: into Eq.~(\ref{eq:CHfunctions}), and 
2108: solving for each order in $a_s^{(n_f-1)}$ the corresponding coupled
2109: differential equations~\cite{pichsanta}
2110: \beqn
2111: 	&&
2112: 	C_1=\frac{1}{6}\,x~,\hspace{0.5cm
2113: 	C_2 = c_{2,0} + \frac{19}{24}\,x + \frac{1}{36}\,x^2~,} \\
2114: 	&&
2115: 	C_3 = c_{3,0} + \left[\frac{241}{54} + \frac{13}{4}\,c_{2,0}
2116: 		- \left(\frac{325}{1728} + \frac{1}{6}\,c_{2,0}\right)n_f\right]x
2117: 		+ \frac{511}{576}\,x^2 + \frac{1}{216}\,x^3~,\\
2118: 	&&
2119: 	H_1 = 0~,\hspace{0.5cm}
2120: 	H_2 = d_{2,0} + \frac{5}{36}\,x - \frac{1}{12}\,x^2~, \\
2121: 	&&
2122: 	H_3 = d_{3,0} + \left[\frac{1627}{1296} - c_{2,0} + \frac{35}{6}\,d_{2,0}
2123: 		+ \left(\frac{35}{648} - \frac{1}{3}\,d_{2,0}\right)n_f
2124: 		+ \frac{5}{6}\,\zeta_3\right]x \nonumber\\
2125: 	&&\hspace{2.4cm}
2126: 		-\: \frac{299}{432}\,x^2 
2127: 		- \left(\frac{37}{216} - \frac{1}{108}\,n_f\right)x^3~,
2128: \eeqn
2129: where the integration coefficients $c_{i,0}$, $d_{i,0}$ are computed
2130: in the \MSbar scheme at the scale of the quark masses 
2131: (\ie, $m_q=\mbar_q(m_q)$)~\cite{wetzel1,wetzel2,chet1,pichsanta}
2132: \beq
2133: 	c_{1,0}=d_{1,0}=0~,\hspace{0.3cm}
2134: 	c_{2,0}=-\frac{11}{72}~, \hspace{0.3cm}
2135: 	c_{3,0}= \frac{82043}{27648}\,\zeta_3 - \frac{575263}{124416}
2136: 			+ \frac{2633}{31104}\,n_f~, \hspace{0.3cm}
2137: 	d_{2,0}=-\frac{89}{432}~.
2138: \eeq
2139: 
2140: The evolution of the \asm measurement from the inclusive $V+A$ 
2141: observables given in Eq.~(\ref{astau:final}), based on Runge-Kutta 
2142: integration of the RGE~(\ref{eq:betafun}) to N$^3$LO 
2143: (see Section~\ref{sec:qcd_RGEs}), and three-loop quark-flavor 
2144: matching~(\ref{eq:CHfunctions}), gives
2145: \beqn
2146: \label{eq:asres_mz}
2147:    \as(M_Z^2) &=& 0.1215 \pm 0.0004_{\rm exp} 
2148:                                        \pm 0.0010_{\rm th} 
2149:                                        \pm 0.0005_{\rm evol}~, \nonumber \\
2150:                          &=& 0.1215 \pm 0.0012~.
2151: \eeqn
2152: The first two errors originate from the \asm determination given
2153: in Eq.~(\ref{astau:final}), and the last error stands for
2154: ambiguities in the evolution due to uncertainties in the 
2155: matching scales of the quark thresholds~\cite{pichsanta}. This
2156: evolution error receives contributions from the uncertainties in 
2157: the $c$-quark mass (0.00020, $m_c$ varied by $\pm0.1\gev$) 
2158: and the $b$-quark mass (0.00005, $m_b$ varied by $\pm0.1\gev$), the 
2159: matching scale (0.00023, $\mu$ varied between $0.7\,m_q$ 
2160: and $3.0\,m_q$), the three-loop truncation in the matching 
2161: expansion (0.00026) and the four-loop truncation in the RGE 
2162: equation (0.00031), where we used for the last two errors 
2163: the size of the highest known perturbative term as systematic 
2164: uncertainty. These errors have been added in quadrature.
2165: The result~(\ref{eq:asres_mz}) is a determination of the strong coupling
2166: at the $Z$ mass scale with a precision of 1\%.
2167: 
2168: The evolution path of \asm is shown in the upper plot 
2169: of Fig.~\ref{fig:evolution}.
2170: The two discontinuities are due to the quark-flavor matching
2171: at $\mu=2\mbar_q$. One could prefer to avoid the discontinuities by 
2172: choosing $\mu=\mbar_q$ so that the logarithms in Eq.~(\ref{eq:CHfunctions})
2173: vanish, and the matching becomes (almost) smooth. However, in this case,
2174: one must first  evolve from $m_\tau$ down to $\mbar_c$ to match the 
2175: $c$-quark flavor, before evolving to $\mbar_b$. 
2176: The effect on \asZ from this ambiguity is within the assigned 
2177: systematic uncertainty for the evolution.
2178: \begin{figure}[t]
2179: %
2180: %...kumac is in: ~hoecker/alpha_s/evolution/eval.kumac
2181:   \centerline{\epsfxsize12.0cm\epsffile{figures/evolution_4.eps}}
2182:   \vspace{-0.7cm}
2183:   \caption[.]{\label{fig:evolution}
2184:       	\underline{Top}: The evolution of \asm~(\ref{astau:final}) to 
2185: 	higher scales $\mu$ using the four-loop RGE and the 3-loop 
2186: 	matching conditions applied at the heavy quark-pair thresholds 
2187: 	(hence the discontinuities at $2m_c$ and $2m_b$). The
2188:       	evolution is compared with other independent measurements (see
2189:       	text) covering scales varying over more than two orders magnitude. 
2190: 	The experimental values are briefly discussed in the text; they are 
2191: 	mostly taken from the compilation~\cite{bethke04}.
2192:         \underline{Bottom}: The corresponding extrapolated $\as$ values
2193:        	at $M_Z$. The shaded band displays the $\tau$ decay result within 
2194: 	errors.
2195: 	}
2196: \end{figure} 
2197: 
2198: \subsubsection{Comparison with other determinations of $\as(M_Z^2)$}
2199: \label{sec:qcd_as_comp}
2200: 
2201: The evolution of \asm in Fig.~\ref{fig:evolution}
2202: is compared with other independent measurements compiled 
2203: in~\cite{bethke04}. Listed below are some outstanding measurements made 
2204: over a large energy range in a variety of processes with different 
2205: experimental and theoretical precisions:
2206: \begin{description}
2207: 
2208: \item[DIS (Bj-SR):] $\as(1.58\gev)=0.375^{+0.062}_{-0.081}$~\cite{ek95}
2209:      was obtained using the polarized structure function data 
2210:      for protons $g^p_1(x)$ and neutrons $g^n_1(x)$ 
2211:      from the CERN SMC~\cite{aea94} and SLAC E142~\cite{e143:95} experiments,
2212:      based on the Bjorken sum rule~\cite{bj70} at 
2213:      N$^2$LO~\cite{ltv91,lv91}
2214:      \begin{equation}
2215:         \intl^1_0\left[g^p_1(x)-g^n_1(x)\right]dx
2216: 	=\frac{1}{3}\left|\frac{g_A}{g_V}\right|
2217: 		\left[1-\frac{\as}{\pi}
2218: 			-3.58\left(\frac{\as}{\pi}\right)^{\!2}
2219: 			-20.2\left(\frac{\as}{\pi}\right)^{\!3}
2220: 		\right]\,.
2221:      \end{equation}
2222:      Here $g_V$ and $g_A$ are the vector and axial-vector coupling 
2223:      constants of the neutron decay, $|g_A/g_V|=-1.2573\pm 0.0028$ (determined
2224:      assuming CVC), and the last two coefficients are
2225:      calculated for three active flavors. 
2226:      No explicit corrections for nonperturbative 
2227:      higher twist effects\footnote
2228:      {
2229:          Interactions involving more than one parton,
2230:          like secondary interactions of quarks with the target remnant,
2231:          give rise to additional contributions to the structure functions.
2232:          These ``higher twist'' terms are classified according to their $Q^2$
2233:          dependence $1/Q^{2n}$, where $n=0$ is leading twist (twist two),
2234:          $n=1$ is twist four, etc.
2235:      } 
2236:      were applied to
2237:      derive this result. However an estimate of the size of 
2238:      the perturbative ${\cal O}(\alpha^4_s)$ term was taken into account.
2239: 
2240: \item[DIS (GLS-SR):] 
2241:      $\as(1.73\gev)=0.280\pm 0.061_{\rm exp}\left.^{+0.035}_{-0.030}\right|_{\rm th}$ is an update of 
2242:      the analysis using structure function $F_3$ data in the $Q^2$ range 
2243:      from $1$ to $15.5\gev^2$ performed by the CCFR 
2244:      collaboration~\cite{ccfr98}. It uses the Gross-Llewellyn-Smith (GLS) 
2245:      sum rule~\cite{gls69} at N$^2$LO~\cite{ck92,ltv91,lv91}
2246:      \begin{equation}
2247:         \intl^1_0 F_3(x,Q^2)dx
2248: 		\equiv3\left[1-\frac{\as}{\pi}
2249: 				-3.58\left(\frac{\as}{\pi}\right)^{\!2}
2250: 				-19.0\left(\frac{\as}{\pi}\right)^{\!3}
2251: 			\right]\,,
2252:      \end{equation}
2253:       given for three active flavors.
2254:       The systematic uncertainty is dominated by the extrapolation of 
2255:       the GLS integral to the region $x<0.01$ where no measurements exist, 
2256:       and to $x>0.5$, which is substituted by the structure function $F_2$ from 
2257:       SLAC data~\cite{withlow:1992}. 
2258:       The theoretical error is dominated by uncertainties 
2259:       in the higher twist corrections.
2260: 
2261: \item[\boldmath $\Upsilon$ sum rules:] 
2262:      $\as(m_b)|_{m_b=4.75\gev}=0.217\pm 0.021$
2263:      was found together with the bottom quark pole mass $m_b$ 
2264:      from sum rules for the $\Upsilon$ system in N$^2$LO, resumming all 
2265:      ${\cal O}\left(\as^2,\as v, v^2\right)$ terms with $v$ being 
2266:      the velocity of the heavy quark~\cite{pp98}. The precision of 
2267:      this result on $\as(m_b)$ was not confirmed by a similar 
2268:      analysis~\cite{Hoang}, where the theoretical parameter space
2269:      has been scanned, which leads to a more conservative estimate 
2270:      of the theory uncertainty.
2271: 
2272: \item[\boldmath $e^+e^- (\sigma_{\rm had})$:] 
2273:      $\as(10.52\gev)=0.20\pm 0.06$
2274:      was determined in N$^2$LO by CLEO~\cite{cleo98} using the total 
2275:      $e^+e^-$ hadronic cross section according to
2276:      \begin{equation}
2277:      R=\frac{\sigma(e^+e^-\to {\rm hadrons})}
2278:      	{\sigma(e^+e^-\to \mu^+\mu^-)}=3\sum_i Q^2_i
2279:      		\left[1+\frac{\as}{\pi}
2280: 			+1.52\left(\frac{\as}{\pi}\right)^2
2281: 			-11.5\left(\frac{\as}{\pi}\right)^3
2282: 		\right]~,
2283:      \end{equation}
2284:      where $Q_i$ are the electric charges of quark flavors $i$ 
2285:      that are produced at center-of-mass energy just below the 
2286:      $\FourS$ resonance in the $e^+e^-$ continuum.
2287:      Another measurement $\as(42.4\gev)=0.175\pm 0.028$ 
2288:      was obtained with the use of data from PETRA and TRISTAN in the 
2289:      center-of-mass energy range from $20$ to $65\gev$~\cite{dh95,bethke00}. 
2290: 
2291: \item[\boldmath $e^+e^-$ (jet \& event-shape):] Using data from PETRA, 
2292:      TRISTAN, SLC and LEP1,2 over a large energy range,
2293:      several determinations of $\as$ were obtained
2294:      based on the jet rate and event-shape variables that are sensitive 
2295:      to gluon radiation governed by the strong coupling constant.
2296:      The observables are calculated in resummed NLO (\ie, NLO matched 
2297:      with next-to-leading order logarithms). 
2298:      In all these determinations, the theoretical uncertainty from the
2299:      renormalization scale uncertainty is the dominant
2300:      error source. At low PETRA energy the hadronization uncertainty 
2301:      is also important. See references in~\cite{bethke00}.
2302: 
2303: \item[\boldmath $p\overline{p}\to b\bbar X$:]
2304:      $\as(20\gev)=0.145\left.^{+0.012}_{-0.010}\right|_{\rm
2305:      exp}\left.^{+0.013}_{-0.016}\right|_{\rm th}$ was obtained by UA1~\cite{ua1:96}
2306:      from a measurement of the cross section of the process 
2307:      $p\overline{p}\to b\bbar X$ for which NLO QCD predictions 
2308:      exist. The theoretical error includes uncertainties due to 
2309:      different sets of structure functions, renormalization/factorization 
2310:      scale uncertainties and the $b$-quark mass.
2311: 
2312: \item[\boldmath $p\overline{p},pp\to \gamma X$:]
2313:      $\as(24.3\gev)=0.135\pm0.006_{\rm exp}\left.^{+0.011}_{-0.005}\right|_{\rm th}$ 
2314:      was determined by UA6~\cite{ua6:99} in NLO from a measurement of 
2315:      the cross section difference 
2316:      $\sigma(p\overline{p}\to \gamma X)-\sigma(pp\to \gamma X)$
2317:      such that the poorly known contributions of the sea quarks and
2318:      gluon distributions in the proton cancel. 
2319:      The theoretical error includes uncertainties from the scale choice 
2320:      and from the variation of the parton distribution functions.
2321: 
2322: \item[\boldmath $e^+e^-$ ($Z$ width):]
2323:      $\as(M_Z)_{Z\,{\rm width}}=0.1186\pm 0.0027$ was determined from
2324:      the hadronic width at the $Z^0$ resonance by a global fit to all 
2325:      electroweak data in N$^2$LO~\cite{ewfit}.
2326: 
2327: \end{description}
2328: 
2329: A comparison is best achieved by extrapolating these measurements 
2330: to $M_Z$ as shown in the lower plot in Fig.~\ref{fig:evolution}.
2331: The most precise result at the $M_Z$ scale stems from $\tau$ 
2332: decays~(\ref{eq:asres_mz}), which is a 1.1\% determination limited in 
2333: accuracy by theoretical uncertainties in the perturbative expansion. 
2334: The significant improvement in precision compared to the previous ALEPH 
2335: result~\cite{aleph_asf}, 
2336: $0.1202 \pm 0.0008_{\rm exp} \pm 0.0024_{\rm th} \pm 0.0010_{\rm evol}$, 
2337: is due to the higher statistics and the more detailed experimental analysis,
2338: but mostly because of the smaller theory uncertainty assigned to 
2339: the perturbative expansion and our conclusion to lift the 
2340: ambiguity between \FOPTCI\  and FOPT.
2341: 
2342: Another precise determination (DIS ($e/\mu; F_2$) [$1.9-15.2$]\gev),
2343: $\as(M_Z^2)=0.1166\pm 0.0009_{\rm stat}\pm 0.0020_{\rm syst}$~\cite{sy01,bethke03},
2344: obtained using the structure function $F_2$ from deep inelastic electron 
2345: ($e$) and muon ($\mu$) scattering (DIS) data in the $Q^2$ range between 
2346: $3.5\gev^2$ and $230\gev^2$ is also shown. The analyses use N$^2$LO
2347: calculations wherever available. The systematic error is mostly theoretical
2348: including higher twist effects and an estimate of the N$^3$LO corrections, 
2349: and is doubled to account for missing contributions associated with the
2350: renormalization scale and scheme uncertainties.
2351: 
2352: In addition to the measurements shown in Fig.~\ref{fig:evolution}, 
2353: there is now a rather precise determination from lattice 
2354: QCD simulations~\cite{davis04}: $\alpha_s(M^2_Z)=0.121\pm 0.003$. This is
2355: the first lattice determination with realistic quark vacuum polarization
2356: and an ${\cal O}(a^2)$ improved staggered-quark discretization of the
2357: light-quark action.
2358: 
2359: \subsubsection{A measure of asymptotic freedom between $m_\tau^2$ and $M_Z^2$}
2360: \label{sec:qcd_as_runningness}
2361: 
2362: The $\tau$-decay and $Z$-width determinations have comparable accuracies, 
2363: which are however very different in nature. The $\tau$ value is 
2364: dominated by theoretical uncertainties, whereas the determination 
2365: at the $Z$ resonance, benefiting from the much larger energy scale 
2366: and the correspondingly small uncertainties from the truncated 
2367: perturbative expansion, is limited by the experimental precision 
2368: on the electroweak observables, essentially the ratio of leptonic 
2369: to hadronic peak cross sections. The consistency between the two results
2370: provides the most 
2371: powerful present test of the evolution of the strong interaction 
2372: coupling, over a range of $s$ spanning more than three 
2373: orders of magnitude, as it is predicted by the nonabelian nature of 
2374: the QCD gauge theory. The difference between the extrapolated
2375: $\tau$-decay value and the measurement at the $Z$ is:
2376: \beq
2377:  \as^\tau(M_Z^2)-\as^Z(M_Z^2) = 0.0029 \pm 0.0010_\tau \pm 0.0027_Z~,
2378: \eeq
2379: which agrees with zero with a relative precision of 2.4\%.
2380: 
2381: In fact, the comparison of these two values is valuable since they are among
2382: the most precise single measurements and they are widely spaced in energy 
2383: scale. Thus it allows one to perform an accurate test of asymptotic freedom.
2384: Let us consider the following evolution estimator~\cite{delphi_run} 
2385: for the inverse of $\as(s)$,
2386: \beq
2387:  r(s_1,s_2) = 2\cdot\frac {\as^{-1}(s_1)-\as^{-1}(s_2)} 
2388:                            {\ln s_1 -\ln s_2}~,
2389: \eeq
2390: which reduces to the logarithmic derivative of $\as^{-1}(s)$ 
2391: when $s_1 \rightarrow s_2$, 
2392: \beqn
2393: \frac {d\as^{-1}}{d\ln \sqrt{s}} &=& -\frac {2\pi\beta(s)}{\as^2}~,\nonumber \\
2394:  &=&  \frac {2\beta_0}{\pi}~(1 + \frac {\beta_1} {\beta_0} \frac {\as}{\pi} +\cdots~)~, 
2395: \eeqn
2396: with the notations of Eq.~(\ref{eq:betafun}). 
2397: At first order, the logarithmic derivative is driven by $\beta_0$.
2398: 
2399: The $\tau$ and $Z$ experimental determinations of $\as(s)$ yield the value
2400: \beq
2401:  r_{\rm exp}(m_\tau^2,M_Z^2) = 1.405 \pm 0.053~,
2402: \eeq
2403: which agrees with the QCD prediction using the RGE to N$^3$LO, 
2404: and three-loop quark-flavor matching~(\ref{eq:CHfunctions}), as discussed 
2405: in Section~\ref{sec:qcd_as_evolution},
2406: \beq
2407:  r_{\rm QCD}(m_\tau^2,M_Z^2) = 1.353 \pm 0.006~.
2408: \eeq
2409: This is to our knowledge the most precise experimental test of the
2410: asymptotic freedom property of QCD at present. It can be compared to 
2411: an independent  determination~\cite{delphi_run}, using an analysis of event 
2412: shape observables at LEP between the $Z$ energy and 207~\gev,
2413: $r(M_Z^2,(207~\gev)^2) = 1.11 \pm 0.21$, for a QCD expectation
2414: of 1.27.
2415: