hep-ph0011301/tor.tex
1: %% espcrc2.tex %%%%%%%%%%
2: %
3: % $Id: espcrc2.tex 1.1 1999/07/26 10:28:22 Simon Exp spepping $
4: %
5: \documentclass[twoside]{article}
6: \usepackage{fleqn,espcrc2}
7: 
8: % change this to the following line for use with LaTeX2.09
9: % \documentstyle[twoside,fleqn,espcrc2]{article}
10: 
11: % if you want to include PostScript figures
12: \usepackage{graphicx}
13: % if you have landscape tables
14: \usepackage[figuresright]{rotating}
15: 
16: % put your own definitions here:
17: %   \newcommand{\cZ}{\cal{Z}}
18: %   \newtheorem{def}{Definition}[section]
19: %   ...
20: \newcommand{\ttbs}{\char'134}
21: \newcommand{\AmS}{{\protect\the\textfont2
22:   A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}
23: 
24: \font\quo=cmssqi8
25: \font\bsy=cmbsy10
26: \def\ub{\underbar}
27: \def\be{\begin{equation}}
28: \def\ee{\end{equation}}
29: \def\bea{\begin{eqnarray}}
30: \def\eea{\end{eqnarray}}
31: 
32: % add words to TeX's hyphenation exception list
33: \hyphenation{author another created financial paper re-commend-ed Post-Script}
34: 
35: % declarations for front matter
36: \title{KNO scaling 30 years later}
37: 
38: \author{S\'andor Hegyi\address{Particle Physics Department \\
39:         KFKI Research Institute for Particle and Nuclear Physics \\ 
40:         H-1525 Budapest 114, P.O. Box 49. Hungary}}
41: 
42:        
43: \begin{document}
44: 
45: 
46: \begin{abstract}
47: KNO scaling, i.e. 
48: the collapse of multiplicity distributions $P_n$ onto a universal 
49: scaling curve manifests when $P_n$ is expressed as the distribution 
50: of the standardized multiplicity $(n-c)/\lambda$ with $c$ and 
51: $\lambda$ being location and scale parameters governed by leading 
52: particle effects and the growth of average multiplicity. 
53: At very high energies, strong violation of KNO scaling behavior is 
54: observed ($p\bar p$) and expected to occur ($e^+e^-$). This challenges one 
55: to introduce novel, physically well motivated and preferably simple scaling
56: rules obeyed by high-energy data. One possibility what I find useful
57: and which satisfies the above requirements is the repetition of the original 
58: scaling prescription (shifting and rescaling) in Mellin space, that is, 
59: for the multiplicity moments' rank. This scaling principle will be 
60: discussed here, illustrating its capabilities both on model predictions and 
61: on real data.
62: \end{abstract}
63: 
64: % typeset front matter (including abstract)
65: \maketitle
66: 
67: \leftline{\quo Dedicated to Wolfram Kittel}
68: \leftline{\quo on the occasion of his 60th birthday}
69: 
70: \vspace{.5cm}
71: 
72: \section{THE EARLY YEARS}
73: 
74: \subsection{Polyakov and Koba-Nielsen-Olesen}
75: 
76: In the Millennium Year we celebrated the 30{\it th\/} anniversary of 
77: the famous Eq.~(1), the basic result of Polyakov and of
78: Koba, Nielsen and Olesen concerning the asymptotic behavior of the
79: multiplicity distributions~\cite{pol,kno1,kno2}. These authors put forward 
80: the hypothesis that at very high energies~$s$ the probability
81: distributions 
82: $P_n(s)$ of producing $n$ particles in a certain collision
83: process should exhibit the scaling (homogeneity) relation
84: \begin{equation}
85:         P_n(s)=
86:         {1\over\langle n(s)\rangle}\,\psi\!
87:         \left({n\over\langle n(s)\rangle}\right)
88: \end{equation}
89: as $s\to\infty$ with
90: $\langle n(s)\rangle$ being the average multiplicity of secondaries at 
91: collision energy~$s$. 
92: This so-called KNO scaling hypothesis asserts that if we rescale $P_n(s)$
93: measured at different energies via stretching (shrinking) the vertical
94: (horizontal) axes by $\langle n(s)\rangle$, these rescaled curves will
95: coincide with each other. That is, the multiplicity distributions
96: become simple rescaled copies of the universal function $\psi(z)$ depending
97: only on the scaled multiplicity $z=n/\langle n(s)\rangle$. In the picturesque
98: terminology of Stanley~\cite{sta} the rescaled data points $P_n(s)$ measured
99: at different energies~$s$ collapse onto the unique scaling curve $\psi(z)$.
100: After Koba~\cite{kno2}, let me illustrate this schematically:
101: 
102: \begin{center}
103: \includegraphics[height=4.8cm]{fig3a.eps}
104: \begin{eqnarray*}
105: \qquad\qquad
106: {\quad\quad\Big\Downarrow\quad\mbox{\fbox{\it rescaling\thinspace}}}
107: \end{eqnarray*}
108: \includegraphics[height=4.8cm]{fig3bbb.eps}
109: \end{center}
110: 
111: \vspace{-.3cm}
112: 
113: The data collapsing behavior is a 
114: fundamental property of homogeneous functions.
115: According to Eq.~(1), at very high energies the multiplicity distributions 
116: $P_n(s)$ are homogeneous functions of degree $-1$ of $n$ and
117: $\langle n(s)\rangle$. Homogeneity rules play a central role in the
118: theory of critical phenomena~\cite{sta}. Near the critical point of a
119: physical system the thermodynamic functions exhibit homogeneous form
120: which implies the existence of a law of corresponding states: using a
121: suitably chosen scaling transformation it is possible to bring different 
122: states of the same system to coincidence and thus to compress many 
123: experimental or theoretical results into a compact form. The KNO 
124: scaling hypothesis Eq.~(1) is just such a law of corresponding states.
125: 
126: In the celebrated paper~\cite{kno1} Koba, Nielsen and Olesen showed 
127: that the validity
128: of Feynman scaling is a sufficient condition for Eq.~(1) to hold. It is 
129: worth mentioning however that the mathematical rigor of their derivation 
130: was criticized by several authors. Moreover, the precise finding of 
131: Ref.~\cite{kno1} was the energy independence of the moments
132: $C_q=\langle n^q(s)\rangle/\langle n(s)\rangle^q$ in the 
133: limit $s\to\infty$.
134: The scaling hypothesis Eq.~(1) was formulated with the additional
135: assumption that the moments $C_q=\int_0^\infty z^q\psi(z)\,dz$ determine
136: uniquely the scaling function $\psi(z)$. 
137: Polyakov arrived at the asymptotic multiplicity scaling law Eq.~(1)
138: two years earlier than Koba, Nielsen and Olesen by 
139: formulating a similarity hypothesis for strong 
140: interactions in $e^+e^-\to\mbox{\it hadrons\/}$ annihilation; see later.
141: Somehow this fundamental and elegant work was almost completely ignored 
142: in the early 70s and Eq.~(1) became known as the KNO scaling 
143: hypothesis in the high-energy physics community.
144: 
145: 
146: \subsection{Modifications}
147: 
148: 
149: The scaling relation Eq.~(1) can hold only approximately since
150: multiplicity~$n$ is a discrete random variable whose stretching or shrinking 
151: by a scale factor leaves the probabilities~$P_n$ unaltered. The proper 
152: meaning of Eq.~(1) is that with increasing collision energy~$s$ 
153: the discrete multiplicity distributions $P_n$ can be approximated
154: with increasing accuracy by a continuous probability density function
155: $f(x)$ via $P_n\approx f(x=n)$ (KNO prescription)
156: or by $P_n\approx\int_{x=n}^{x=n+1}f(x)\,dx$ (called KNO-G ``scaling'')
157: where, in the most popular modification of the original scaling rule, 
158: $f(x)$ has the generic Czy\.zewski-Rybicki form~\cite{cry,bla}
159: \be
160: 	f(x)=\frac{1}{\lambda}\,\psi\left(\frac{x-c}{\lambda}\right)
161: \ee
162: with scale parameter $\lambda>0$ and location parameter $c$. 
163: Data collapsing behavior is observed if the only $s$-dependent parameters 
164: of the approximate shape function $f(x)$ are $c$ and $\lambda$, i.e. if 
165: $f(x)$ depends on collision energy only through a change of location 
166: and scale in~$x$. The variation of $\lambda$ with increasing~$s$
167: reflects the growth of average multiplicity, whereas the $s$-dependence 
168: of $c$ is usually associated with leading particle effects (so-called
169: AKNO or KNO-$\alpha$ scaling). Other physical explanations are also possible, 
170: see e.g. the contribution of Marek P\l oszajczak on $\Delta$-scaling.
171: 
172: 
173: \section{NEW SCALING PRINCIPLE}
174: 
175: 
176: Taking into account a possible location change (shift) in 
177: multiplicity rarely proved to be sufficient to restore the data
178: collapsing behavior of $P_n$ when the original scaling law Eq.~(1) is 
179: violated at very high energies (hence we use \hbox{$c=0$}). 
180: Nevertheless, the principle of scale {\it and\/} 
181: location change in the shape of multiplicity distributions can be 
182: extremely powerful through the following generalization of the KNO scaling
183: rule. Consider that the multiplicity distributions, more precisely the 
184: continuous probability densities $f(x)$ approximating $P_n$ generate
185: $s$-dependent shifting and rescaling in Mellin space.
186: The Mellin transform of a 
187: probability density $f(x)$ is defined by 
188: ${\cal M}\{f(x);q\}=\int_0^\infty x^{q-1}f(x)\,dx$ and it provides the
189: $(q-1)${\it th\/} moment $\langle x^{q-1}\rangle$. Via the functional
190: relation
191: \begin{equation}
192: 	{\cal M}\bigg\{x^r f\big(x^\mu\big);q\bigg\}=
193: 	\frac{1}{\mu}\,{\cal M}\bigg\{f(x);{q+r\over\mu}\bigg\}
194: \end{equation}
195: one can introduce translation and dilatation in the moments' rank $q$ 
196: by performing the transformation $f(x)\to x^r f\big(x^\mu\big)$ of the 
197: probability densities approximating the shape of $P_n$. In the followings
198: I shall illustrate the utility of performing a location or scale change 
199: in Mellin space; for further details, see Refs.~\cite{hs1,hs2,hs3}.
200: 
201: 
202: 
203: \section{TRANSLATION IN {\bsy M}-SPACE}
204: 
205: \subsection{Motivation}
206: 
207: There are several possible dynamical explanations of
208: a shift in Mellin space.  The basic 
209: idea is simply the following. Recall that in Eq.~(1) the 
210: normalization of the KNO function reads
211: $
212: 	\int_0^\infty\psi(z)\,dz=\int_0^\infty z\,\psi(z)\,dz=1,
213: $
214: that is, $\langle z\rangle=1$. Assume that violation of KNO scaling 
215: is observed and we measure $s$-dependent ``scaling'' functions
216: $\psi(z)$. In order to arrive at data collapsing behavior of $P_n$
217: the simplest possibility is to rescale the function
218: $\varphi(z)=z\,\psi(z)$ by its first moment given by $C_2$.
219: In terms of $P_n$ and the moments $\langle n^q\rangle$ 
220: this corresponds to rescaling 
221: the distribution $P_{n,1}=nP_n/\langle n\rangle$ by its mean 
222: $\langle n\rangle_1=\langle n^2\rangle/\langle n\rangle$. The distribution
223: $P_{n,1}$ is known in the mathematical literature
224: as the first {\it moment distribution\/} of $P_n$. If the
225: rescaling of $P_{n,1}$ by $\langle n\rangle_1$ is not sufficient to
226: arrive at data collapsing behavior,
227: one can try the same recipe for the higher-order moment distributions
228: $P_{n,r}=n^rP_n/\langle n^r\rangle$. The generalization is 
229: straightforward. Let me show you now an example when Eq.~(1) 
230: is clearly violated, but performing a shift in Mellin space can
231: restore the collapse of different $P_n$ curves onto a single 
232: scaling curve.
233: 
234: \subsection{Intermittency and multifractality}
235: 
236: The most obvious example is intermittency. As it turned out during the 
237: past 15 years, multiparticle production often yields scale-invariant
238: density fluctuations~\cite{bp,bop,ddk}. This can be observed
239: through the power-law scaling $C_q\propto\delta^{-\varphi_q}$ of the 
240: normalized moments $C_q=\langle n^q\rangle/\langle n\rangle^q$ 
241: of $P_n(\delta)$ as the bin-size~$\delta$ in phase-space is varied (for 
242: clarity, we neglect here the influence of low count rates).
243: Assuming that the fluctuations show monofractal structure,
244: the so-called intermittency exponents 
245: $\varphi_q$ are given by~$\varphi_q=\varphi_2(q-1)$ and the anomalous 
246: fractal dimensions $D_q=\varphi_q/(q-1)$ are \hbox{$q$-independent}, 
247: $D_q=D_2$. The normalized moments $C_q$ of $P_n(\delta)$ take the form
248: \begin{equation}
249: 	C_q=A_q\,[C_2]^{\,q-1}\qquad\mbox{for}\quad q>2
250: \end{equation}
251: with coefficients $A_q$ independent of bin-size~$\delta$~\cite{ddk}. 
252: Of course Eq.~(1), with $s\to\delta$, is violated since we observe
253: $C_2\propto\delta^{-\varphi_2}$.
254: 
255: In the restoration of data collapsing behavior of $P_n$ for 
256: self-similar fluctuations the basic trick is the investigation 
257: of the higher-order moment distributions defined before. Their moments
258: are $\langle n^q\rangle_r=\langle n^{q+r}\rangle/\langle n^r\rangle$,
259: that is, the moments of the original $P_n$ are transformed out up to 
260: $r$th order via performing a shift in Mellin space, see
261: Eq.~(3). For $r=1$ the normalized moments of the 
262: first moment distribution $P_{n,1}$ are found to be
263: $
264: 	C_{q,1}=C_{q+1}/[C_2]^{\,q}
265: $
266: in terms of the original $C_q$ and comparison to Eq.~(4) yields
267: $C_{q,1}=A_{q+1}$ for monofractal multiplicity fluctuations. Since the 
268: coefficients $A_q$ are independent of bin-size~$\delta$, we see
269: that monofractality yields not only 
270: {\it power-law scaling\/} of the normalized moments of $P_n$ but also 
271: {\it data collapsing\/} behavior of the first moment distributions $P_{n,1}$
272: measured at different resolution scales $\delta$. Increasing the 
273: (possibly fractional)
274: rank~$r$ of the moment distributions allows the restoration of data
275: collapsing behavior in the presence of increasing degree of multifractality
276: of scale-invariant multiplicity fluctuations~\cite{hs1,hs2}. The effect of 
277: low multiplicities (Poisson noise) can be taken into account via the study 
278: of {\it factorial\/} moment distributions 
279: $P_{n,r}=n^{[r]}P_n/\langle n^{[r]}\rangle$
280: and their factorial moments. 
281: 
282: 
283: \subsection{Ginzburg-Landau phase transition}
284: 
285: 
286: In recent years considerable interest has been devoted to the
287: Ginzburg-Landau theory of the phase transition from quark-gluon plasma
288: to hadronic matter~\cite{hwa,hwb}. Instead of a strict power-law scaling
289: $F_q\propto\delta^{-\varphi_q}$ of the normalized factorial moments,
290: a so-called $F$-scaling behavior
291: $F_q\propto[F_2]^{\beta_q}$ is obtained with $\beta_q=(q-1)^\nu$.
292: For second-order transition $\nu=1.304$ and in case of a first-order
293: transition $\nu=1.45$. The $F$-scaling rule of the above form makes
294: difficult to identify the family of 
295: probability laws $P_n$ belonging to the model. But considering 
296: $P_{n,1}$ instead of $P_n$ the corresponding factorial moments 
297: turn out to be $F_{q,1}\propto[F_2]^{\beta_q}$ with $\beta_q=q^\nu-q$ and 
298: hence~\cite{hs1}
299: $$
300:         F_{q,1}\propto[F_{2,1}]^{\beta_q}\qquad\mbox{with}\quad
301:         \beta_q=\frac{q^\nu-q}{2^\nu-2}.
302: $$
303: This form of $F$-scaling is the familiar log-L\'evy law~\cite{lev} with
304: stability index $0<\nu\leq2$. Thus, via moment shifting, we succeeded 
305: in identifying $P_{n,1}$ for the Ginzburg-Landau formalism 
306: with a strict bound on the essential parameter~$\nu$ of the model.
307: The theoretically relevant values are within the allowed range.
308: Although this particular example is not related directly to data collapsing
309: behavior of multiplicity distributions, it illustrates the utility of the 
310: moment distributions $P_{n,r}\propto n^rP_n$ which
311: correspond to translation in Mellin space.
312: 
313: 
314: 
315: \section{RESCALING IN {\bsy M}-SPACE}
316: 
317: 
318: 
319: \subsection{Motivation}
320: 
321: The most important reason of a possible change of scale in
322: Mellin space comes from QCD. In higher-order pQCD calculations,
323: allowing a more precise account of energy conservation in the course of
324: multiple parton splittings, the natural variable of the multiplicity 
325: moments is the {\it rescaled\/} rank $q\gamma$ instead of rank $q$ 
326: itself~\cite{dok,och,dre} with
327: $\gamma(\alpha_s)$ being the multiplicity anomalous dimension of QCD.
328: Because of the running of the strong coupling constant $\alpha_s$,
329: it is inevitable to adjust an energy dependent scale factor in 
330: Mellin space if one wants to arrive at data collapsing behavior
331: of the multiplicity distributions $P_n(s)$.
332: 
333: 
334: 
335: \subsection{Running coupling}
336: 
337: 
338: In the very first paper predicting the multiplicity scaling law Eq.~(1),
339: Polyakov constructed a model for the short-distance behavior of strong
340: interactions based on the hypothesis of asymptotic scale- and conformal
341: invariance~\cite{pol}. When applied to the
342: process of $e^+e^-$ annihilation to hadrons at asymptotically high 
343: collision energies, the model predicts that multihadron production 
344: goes in a cascade way. First, a few heavy virtual objects, called jets, are
345: formed. These are repeatedly diminished in a branching process 
346: until the masses of the subsequent generation jets become at some 
347: very late stage  comparable to the hadronic masses. This cascading mechanism
348: gives rise to multiplicity distributions $P_n(s)$ satisfying Eq.~(1) and
349: the scaling function $\psi(z)$ behaves as
350: $
351: 	\psi(z)\propto a(z)\,\exp(-z^\mu)
352: $
353: with $\mu>1$ and $a(z)$ being a monomial. 
354: Up to a scale factor, $\psi(z)$ is a negative
355: binomial type scaling function, that is, a gamma distribution~\cite{cs} 
356: in the rescaled and power-transformed multiplicity~$z^\mu$.
357: 
358: 
359: It is worth emphasizing that the above result was achieved years earlier
360: than the emergence of quantum chromodynamics. Interestingly, the predictions
361: of QCD obey some important similarities to the findings of Polyakov:
362: 
363: {\it 1)\/} 
364: Hadronic jets produced in hard processes are expected in QCD to be
365: self-similar objects composed of subsequent 
366: parton branching decays~\cite{kon}. A highly
367: virtual quark or gluon decays into secondaries which are less off-shell and
368: less energetic. Each of these intermediate decay products splits into novel
369: ones. The subsequent decay chain steps form a scale-invariant branching 
370: process which continues until the virtual mass of the quanta becomes
371: approximately 1~GeV.
372: 
373: {\it 2)\/} 
374: The fragmentation of partons in QCD leads to the typical characteristics of
375: branching processes, in particular, to long-range correlations and to the
376: validity of KNO scaling~\cite{bas}. Further, it was shown~\cite{tes} that
377: KNO scaling in $e^+e^-$ annihilation is the consequence of the 
378: scale-invariance
379: (asymptotic freedom) in the presence of gluon self-coupling. It is thus
380: the manifestation of the essence of the non-abelian gauge theory of
381: strong interactions.
382: 
383: {\it 3)\/} 
384: Taking into account the higher pQCD effects responsible for
385: energy-momentum conservation in parton jets, the KNO scaling function
386: in $e^+e^-$ annihilation was shown~\cite{dok} to be identical to 
387: Polyakov's form:
388: \be
389: 	\psi(z)={\cal N}z^{\mu k-1}\exp\big(-[Dz]^\mu\big)
390: \ee
391: where $k=3/2$, ${\cal N}=\mu D^{\mu k}/\Gamma(k)$, 
392: $\mu=(1-\gamma)^{-1}\approx5/3$ and $D$ is a scale parameter
393: depending on $\gamma$, with $\gamma$ being the QCD anomalous dimension,
394: $\gamma\approx0.4$ at LEP1. Hence, the
395: conservation laws strongly reduce the width of the KNO scaling function 
396: as compared to the exponential fall-off obtained by lower-order pQCD 
397: calculations. The experimental data at $\sqrt s=91.2$~GeV (dots) confirm
398: the prediction (curve) as illustrated in the figure~a) below
399: (with the exception that $k$ turns out to be larger, $k\approx5$).
400: 
401: 
402: The crucial difference between the QCD prediction and Polyakov's result
403: lies in the fact that the scaling function $\psi(z)$  given by the 
404: Polyakov-Dokshitzer form Eq.~(5) {\it depends\/} on 
405: collision energy~$s$ in QCD. That is, QCD predicts violation of 
406: KNO scaling in $e^+e^-$ annihilation which becomes clearly visible 
407: at higher energies. This effect is due to the running
408: of the strong coupling constant $\alpha_s$, hence       
409: $\gamma\to0$ as $s\to\infty$, i.e., $\mu\to1$
410: asymptotically~\cite{dok}. Therefore the tail of $\psi(z)$ widens
411: with increasing~$s$ which gives rise to the KNO scaling violation pattern
412: shown in the figure~b) below. Asymptotically the exponential fall-off 
413: predicted by the double logarithmic approximation is recovered (solid curve) 
414: i.e. the negative binomial type scaling form of $P_n(s)$.
415: 
416: \vspace{-.2cm}
417: 
418: \begin{center}
419: \hbox{\hspace{-.2cm}\includegraphics[width=7.2cm]{logkno.fig1a.eps}}
420: \hbox{\hspace{-.2cm}\includegraphics[width=7.2cm]{logkno.fig1b.eps}}
421: \hbox{\hspace{-.2cm}\includegraphics[width=7.2cm]{logkno.fig1c.eps}}
422: \end{center}
423: 
424: \vspace{-.7cm}
425: 
426: Thus we have learned that the pre-QCD and QCD-based descriptions of $e^+e^-$
427: annihilation to hadrons make essential use of self-similar branching
428: processes, further, both predict the rescaled and
429: power-transformed multiplicity variable~$z^\mu$ 
430: to be gamma-distributed. But in the pre-QCD approach
431: the exponent $\mu$ is independent of collision energy~$s$ and
432: KNO scaling holds valid, as expected for scale-invariant branching with 
433: {\it fixed coupling\/}. On the contrary, the QCD-based calculations
434: predict violation of KNO scaling since $\mu$ varies with collision energy~$s$, 
435: due to the {\it running\/} of $\alpha_s$. Note however that
436: in the latter case data collapsing can be restored in a simple way using
437: {\it logarithmic\/} scaling variable. For the Polyakov-Dokshitzer form
438: of $\psi(z)$ given by (5) we have
439: $$
440: 	\psi(x)=\mu\exp\big(k\mu x-e^{\mu x}\big)/\Gamma(k),
441: 	\quad x=\ln(Dz).
442: $$
443: Since only the exponent $\mu$ and scale parameter $D$ of Eq.~(5) are 
444: expected to depend on collision energy~$s$ through the variation of 
445: $\gamma(\alpha_s)$, KNO scaling violated by 
446: QCD effects is recovered by plotting $\mu^{-1}\psi(\mu x)$ as displayed in 
447: the figure~c). The scale change in logarithmic multiplicity is 
448: governed by the QCD multiplicity anomalous dimension. This particular
449: type of data collapsing of the multiplicity distributions $P_n(s)$ 
450: is called log-KNO scaling~\cite{hs2,hs3} since we have the behavior
451: of type Eq.~(2) except that distributions of the logarithmic 
452: multiplicity satisfy it.
453: 
454: 
455: 
456: 
457: \subsection{Multiplicative cascades}
458: 
459: 
460: It may turn out that the log-KNO scaling law
461: has ubiquitous appearance in multiparticle dynamics.
462: Random multiplicative cascades play a distinguished role in the dynamics
463: underlying multihadron production, both in soft and hard processes.
464: The multiplicative cascade models like for example the
465: $\alpha$-model and $p$-model are in the focus of interest since the
466: pioneering work of Bia\l as and Peschanski~\cite{bp}, see also~\cite{ddk}
467: for a review. Recently, Frisch and Sornette developed a theory of 
468: extreme deviations~\cite{frs} which predicts for multiplication
469: of random variables the generic presence of stretched exponential
470: distributions $f(x)\propto\exp(-\lambda x^\mu)$ with $\mu<1$. 
471: To be specific, consider the product
472: $
473: 	X_N=m_1m_2\cdots m_N
474: $
475: of independent random variables $m_i$ distributed identically according to
476: the probability density $p(m_i)$. The above product can arise in a cascade 
477: process if primary entities (e.g. particle density) of initial size $s_0$
478: are repeatedly diminished in random proportions, so that after $N$ cascade
479: steps we have the product \hbox{$s_N=s_0m_1m_2\cdots m_N$}.
480: \begin{center} 
481: \includegraphics[height=4cm]{jet.eps}
482: \end{center}
483: In ref.~\cite{frs} it is shown that under
484: mild regularity conditions the distribution of the product variable $X_N$ 
485: exhibits the asymptotic behavior
486: \be
487: 	{\cal P}_N(X)\sim\big[p\big(X^{1/N}\big)\big]^N
488: 	\qquad\mbox{for}\quad X\to\infty
489: \ee
490: and for finite~$N$. Frisch and Sornette call attention to the intuitive
491: interpretation of (6): the tail of ${\cal P}_N(X)$ is controlled by
492: the realizations where all terms in the product are of the same order
493: hence ${\cal P}_N(X)$ is, to leading order, just the product of the $N$
494: densities $p$, each having the common argument $X^{1/N}$. When $p(x)$ is
495: chosen to be $\propto\exp(-\lambda x^\alpha)$ with $\alpha>0$,
496: then Eq.~(6) leads to stretched exponential tails
497: $\propto\exp(-\lambda Nx^{\alpha/N})$ for large $N$, with stretching
498: exponent $\alpha/N<1$. 
499: Expecting the cascade depth $N$ to be an increasing function of collision
500: energy $s$, Eq.~(1) breaks down but data collapsing can be recovered in
501: the log-KNO manner discussed previously.
502: 
503: 
504: \section{SCALING AT TEV ENERGIES}
505: 
506: 
507: The most exciting and challenging task in developing novel scaling 
508: relations is testing them on real data. In case of multiplicity 
509: distributions the KNO scaling laws Eqs.~(1-2) are known to be strongly
510: violated above ISR energies. This violation is most visible for the
511: multiplicity data measured by the E735 Collaboration~\cite{e735} see 
512: also~\cite{wal} for more details. In the followings I provide a very brief
513: summary of some preliminary results of a study of the E735 data.
514: 
515: 
516: \subsection{Analysis of E735 multiplicities}
517: 
518: The E735 Collaboration recently published the full phase-space 
519: multiplicity distributions in $pp$ and $p\bar p$ collisions
520: at c.m. energies $\sqrt s=$ 300, 546, 1000 and 1800 GeV 
521: at Tevatron~\cite{e735}. The
522:  $P_n$ curves corresponding to $\sqrt s=$ 300 and 1800 GeV are 
523: displayed in the following two figures. 
524: \begin{center}
525: \hbox{\hspace{-.4cm}\includegraphics[width=8cm]{scaling.fig1a.eps}}
526: \hbox{\hspace{-.4cm}\includegraphics[width=8cm]{scaling.fig1b.eps}}
527: \end{center}
528: 
529: 
530: It is apparent that bimodal shape of the $P_n$ curves arises at
531: TeV energies, obeying a (not too pronounced) shoulder structure like at SPS.
532: It is argued~\cite{e735,wal}
533: that the low-multiplicity regimes are affected mainly by single parton 
534: collisions and exhibit KNO scaling, 
535: whereas the \hbox{large-$n$} tail of the distributions is
536: influenced more heavily by double parton interactions and violate 
537: KNO scaling substantially. Another possible explanation of the
538: observed $P_n$ shape is the weighted superposition of a soft and
539: semihard component, the latter one being dominated by minijet production.
540: These components are responsible of the KNO scaling and KNO-violating
541: regimes of $P_n$~\cite{gu1,gu2}.
542: 
543: The solid curves in the previous two plots represent the Polyakov-Dokshitzer 
544: parametrization Eq.~(5) fitted to the KNO-violating $z\geq1$
545: component of the E735 data. The quality of fits is reasonably good 
546: (for all data sets) as can be depicted from the figures.
547: The fits correspond to $k=\frac{1}{2}$ kept fixed and $\mu$ decreasing from
548: $\mu\sim2.2$ to $\mu\sim1.7$ as one goes from 
549: $\sqrt s=$ 300 GeV to 1800 GeV. Our findings
550: indicate that the KNO violating component of the E735 multiplicity
551: distributions can be collapsed onto a single scaling curve 
552: in the log-KNO manner discussed previously. It is illustrated below.
553: \begin{center}
554: \hbox{\hspace{-.4cm}\includegraphics[width=8cm]{scaling.fig3a.eps}}
555: \hbox{\hspace{-.4cm}\includegraphics[width=8cm]{scaling.fig3b.eps}}
556: \end{center}
557: 
558: \vspace{-.3cm}
559: 
560: In the upper figure the $\sqrt s=$ 300 and 1800 GeV data sets 
561: are displayed only, 
562: whereas in the lower plot all the four data sets measured by the 
563: E735 experiment. The collapse of the data points onto a unique 
564: scaling curve is apparent. This very preliminary investigation suggests
565: that the pattern of KNO scaling violation by double parton collisions,
566: or by semihard processes dominated by minijet production, is similar
567: to that of $e^+e^-$ annihilations where log-KNO behavior is expected 
568: to arise as well at high energies (see before). 
569: But note that in parameter $k$ there 
570: is an order of magnitude difference between $e^+e^-$ and $p\bar p$ 
571: collisions, that is, the scaling functions have 
572: different shapes in the two reactions. 
573: 
574: 
575: 
576: One of the earliest reviews~\cite{ole} of Polyakov's pioneering work
577: concluded with the following speculative note:
578: {\rm ``Could it be that as in the theory of critical phenomena, while the
579: basic physics (Hamiltonians) between $e^+e^-$ annihilation and hadronic
580: multiparticle production could be entirely different, multiplicity
581: scaling may yet still be a \ub{universal} feature shared by both?''\/}
582: In the light of our findings the answer to this question is {\it yes\/}
583: but the scaling behavior foreseen 30 years ago arises in the logarithm
584: of multiplicity instead of multiplicity itself.
585: 
586: 
587: 
588: \section{SUMMARY}
589: 
590: 
591: In the past 30 years, data collapsing of multiplicity distributions $P_n$
592: according to the generic KNO scaling relation Eq.~(2) was one of the most
593: extensively studied topics in soft physics. Unfortunately, more and more 
594: experimental and theoretical results confirm that Eq.~(2) is too simple 
595: to be true. The information content of $P_n$
596: can be represented in an equally convenient manner 
597: making use of the multiplicity moments.
598: For example, validity of Eq.~(2) corresponds to constancy of the normalized
599: central moments for appropriately chosen $c$ and $\lambda$. What can we do
600: if the KNO scaling rule~(2) breaks down? How data collapsing behavior of 
601: $P_n$ can be restored, if it is~possible at all? 
602: The approach presented here is extremely
603: simple: let's perform the original scaling transformation (shifting and
604: \hbox{rescaling}) in the multiplicity moments' rank, that is, in Mellin space, 
605: rather than in  multiplicity.
606: 
607: The functional relation Eq.~(3) tells how things transform 
608: at the level of $P_n$.
609: Translation in Mellin space (moment shifting) corresponds to changing from
610: $P_n$ to the moment distributions $P_{n,r}\propto n^rP_n$. 
611: This makes possible
612: to arrive at data collapsing for intermittency phenomena: 
613: e.g. $P_{n,1}$ scales
614: for monofractal fluctuations, $P_{n,2}$ for bifractals and so on. Dilatation 
615: in Mellin space (rescaling of moments' rank) corresponds 
616: to the transformation
617: $P_n\to P_m$ with $m=n^\mu$. Therefore data collapsing arises in rescaled 
618: logarithmic multiplicity (log-KNO). Such behavior is expected to occur for
619: phenomena governed by running coupling effects or in multiplicative cascades
620: of varying length. Considering the physical significance of the mentioned
621: examples in multiparticle dynamics, it will be interesting to see how the
622: proposed scaling behaviors show up in the experiments of the Third
623: Millennium.
624: 
625: 
626: 
627: 
628: \section*{ACKNOWLEDGEMENTS}
629: 
630: 
631: I am grateful to the organizers of this workshop, in particular
632: to Alberto and Roberto for the invitation and kind hospitality
633: in Torino. 
634: 
635: The support of grants NWO-OTKA N25186 and OTKA T024094 is acknowledged.
636: 
637: 
638: \begin{thebibliography}{9}
639: \bibitem{pol} A.M. Polyakov, 
640:         Zh. Eksp. Teor. Fiz. {\bf 59}, 542 (1970).
641: \bibitem{kno1} Z. Koba, H.B. Nielsen and P. Olesen, 
642: 	Nucl. Phys. {\bf B40}, 317 (1972). 
643: \bibitem{kno2} Z. Koba,
644: 	in Proc. of the 1973 CERN School of Physics,
645: 	p.~171, CERN Yellow Report CERN-73-12 (1973).
646: \bibitem{sta} H.E. Stanley,
647: 	Introduction to Phase Transitions and Critical Phenomena
648: 	(Clarendon Press, 1971).
649: \bibitem{cry} O. Czy\.zewski and K. Rybicki,
650: 	Nucl. Phys. {\bf B47}, 633 (1972).
651: \bibitem{bla} M. Bla\v zek,
652: 	Z. Phys. {\bf C32}, 309 (1986).
653: \bibitem{hs1} S. Hegyi,
654: 	Phys. Lett. {\bf B411}, 321 (1997).
655: \bibitem{hs2} S. Hegyi,
656: 	Phys. Lett. {\bf B466}, 380 (1999).
657: \bibitem{hs3} S. Hegyi,
658: 	Phys. Lett. {\bf B467}, 126 (1999).
659: \bibitem{bp} A. Bia\l as and R. Peschanski,
660:         Nucl. Phys. {\bf B273}, 703 (1986); {\bf B308}, 857 (1988).
661: \bibitem{bop} P. Bo\.zek and M. P\l oszajczak,
662: 	Phys. Rep. {\bf 252}, 101 (1995).
663: \bibitem{ddk} E.A. De Wolf, I.M. Dremin and W. Kittel,
664: 	Phys. Rep. {\bf 270}, 1 (1996).
665: \bibitem{hwa} R.C. Hwa and M.T. Nazirov,
666:         Phys. Rev. Lett. {\bf 69}, 741 (1992).
667: \bibitem{hwb} R.C. Hwa, Phys. Rev. {\bf D47}, 2773 (1993);
668:                         Phys. Rev. {\bf C50}, 383 (1994).
669: \bibitem{lev} Ph. Brax and R. Peschanski,
670: 	Phys. Lett. {\bf B253}, 225 (1991).
671: \bibitem{cs} P. Carruthers and C.C. Shih,
672: 	Int. J. Mod. Phys. {\bf A2}, 1447 (1987).
673: \bibitem{kon} K. Konishi, A. Ukawa and G. Veneziano,
674: 	Nucl. Phys. {\bf B157}, 45 (1979).
675: \bibitem{bas} A. Bassetto, M. Ciafaloni and G. Marchesini,
676: 	Nucl. Phys. {\bf B163}, 477 (1980).
677: \bibitem{tes} K. Tesima,
678: 	Phys. Lett. {\bf B133}, 115 (1983).
679: \bibitem{dok} Yu.L. Dokshitzer,
680: 	Phys. Lett. {\bf B305}, 295 (1993); LU-TP/93-3 (1993).
681: \bibitem{och} V.A. Khoze and W. Ochs,
682: 	Int. J. Mod. Phys. {\bf A12}, 2949 (1997).
683: \bibitem{dre} I.M. Dremin and W. Gary,
684: 	hep-ph/0004215, to appear in Phys. Rep. (2000).
685: \bibitem{frs} U. Frisch and D. Sornette,
686: 	J. Phys. I France {\bf 7}, 1155 (1997).
687: \bibitem{e735} E735 Collab., T. Alexopoulos et al.,
688:         Phys. Lett. {\bf B435}, 453 (1998).
689: \bibitem{wal} S.G. Matinyan and W.D. Walker,
690:         Phys. Rev. {\bf D59}, 034022 (1999).
691: \bibitem{gu1} A. Giovannini and R. Ugoccioni,
692:         Phys. Rev. {\bf D59}, 094020 (1999).
693: \bibitem{gu2} A. Giovannini and R. Ugoccioni,
694:         Phys. Rev. {\bf D60}, 074027 (1999).
695: \bibitem{ole} P. Olesen and H.C. Tze,
696: 	``Multiplicity scaling in $e^+e^-$ annihilation and
697: 	Polyakov's bootstrap'', NBI-HE 74-9 (1974).
698: \end{thebibliography}
699: 
700: 
701: 
702: \end{document}
703: