1: \documentclass[tbtags,12pt,a4paper]{article}
2: \usepackage{a4wide}
3: \usepackage{fancyhdr}
4: \usepackage{graphicx}
5: \usepackage{float}
6: \usepackage{enumerate}
7: \usepackage{amsmath}
8: \usepackage{amssymb}
9: \usepackage{amsfonts}
10: \usepackage{amsthm}
11: \usepackage{amssymb}
12: \usepackage{mathrsfs}
13: %\usepackage{subeqn}
14: \usepackage{equation}
15:
16: \begin{document}
17: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
18: \def\gsim{\:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\:}
19: \def\lsim{\:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\:}
20:
21: \begin{titlepage}
22:
23: \begin{flushright}
24: RIKEN--TH 86 \\[-0.1cm]
25: October 2006\\[5mm]
26: \end{flushright}
27:
28: \vspace{0.5cm}
29:
30: \begin{center}
31: {\Large \bf QCD Effects in the Decays of TeV Black Holes}\\[1.cm]
32: {\large Christian Alig$^{1}$, Manuel Drees$^1$, and Kin-ya Oda$^2$}
33: \end{center}
34:
35: \vskip 0.5cm
36:
37: {\small
38: \begin{center}
39:
40: $^1$ {\it Physikalisches Institut, Universit\"at Bonn, Nussallee 12, D53115
41: Bonn, Germany} \\[2mm]
42: $^2$ {\it Theoretical Physics Laboratory, RIKEN, Saitama 351-0198, Japan}
43:
44: \end{center}
45: }
46:
47: \vspace{1.cm}
48: \begin{abstract}
49:
50: \noindent In models with ``large'' and/or warped extra dimensions, the
51: higher--dimensional Planck scale may be as low as a TeV. In that case black
52: holes with masses of a few TeV are expected to be produced copiously in
53: multi--TeV collisions, in particular at the LHC. These black holes decay
54: through Hawking radiation into typically ${\cal O}(20)$ Standard Model
55: particles. Most of these particles would be strongly interacting. Naively this
56: would lead to a final state containing 10 or so hadronic jets. However, it
57: has been argued that the density of strongly interacting particles would be so
58: large that they thermalize, forming a ``chromosphere'' rather than
59: well--defined jets. In order to investigate this, we perform a QCD simulation
60: which includes parton--parton scattering in addition to parton showering. We
61: find the effects of parton scattering to remain small for all cases we
62: studied, leading to the conclusion that the decays of black holes with masses
63: within the reach of the LHC will {\em not} lead to the formation of
64: chromospheres.
65:
66: \end{abstract}
67: \vskip 0.5cm
68:
69: \end{titlepage}
70:
71: \section{Introduction}
72:
73: The Standard Model (SM) of particle physics is very successful in reproducing
74: experimental data. However, from the theoretical point of view it has several
75: unsatisfactory features. Its perhaps biggest problem stems from the very large
76: hierarchy between the electroweak mass scale $M_{\rm weak} \sim 100$ GeV and
77: the (reduced) Planck mass $M_{\rm P} \simeq 2.4 \cdot 10^{18}$ GeV. The origin
78: of this hierarchy has no explanation within the SM; worse, it tends to be
79: destroyed by quantum corrections.
80:
81: In the late 1990's it has been suggested \cite{Arkani-Hamed:1998rs,
82: Randall:1999ee} that this problem could be solved by introducing additional
83: spatial dimensions. They could be relatively large if only gravity is allowed
84: to propagate in them, i.e. if all (other) SM fields are bound to a brane with
85: three spatial dimensions. The strength of the gravitational interactions at
86: lengths larger than the radii of these additional dimensions would then be
87: diluted by a factor which is essentially given by the ratio of the volume of
88: these extra dimensions and the appropriate power of the higher--dimensional
89: Planck length. This ratio can be very large even if the higher--dimensional
90: Planck mass $M_D$ is not far above $M_{\rm weak}$. If this idea is correct,
91: the very large measured value of $M_{\rm P}$ is an artefact of using a
92: 3--dimensional description of a world that actually has $3+n$ spatial
93: dimensions.
94:
95: This scenario leads to dramatic predictions for collisions of pointlike
96: particles at high center--of--mass (cms) energy. Collisions at energies around
97: $M_D$ would likely be dominated by the exchange of gravitons
98: \cite{'tHooft:1987rb}, rather than by exchange of SM gauge bosons. Collisions
99: at energies exceeding $M_D$ could lead to the formation of black holes
100: \cite{banks, Giddings:2001bu, Dimopoulos:2001hw, Eardley:2002re,
101: Solodukhin:2002ui, Hsu:2002bd, Yoshino:2002tx, Yoshino:2005hi}, with cross
102: section being given by the square of the (generalized) Schwarzschild radius.
103: Since this radius actually increases with increasing cms energy, this would
104: lead to ``the end of short--distance physics'' \cite{Giddings:2001bu}.
105:
106: These black holes with masses of a few TeV should decay very quickly via the
107: higher--dimension analogue \cite{Myers:1986un} of Hawking radiation
108: \cite{Hawking:1974sw}, into final states on the brane consisting mostly of SM
109: degrees of freedom \cite{Emparan:2000rs}. Since the Hawking temperature of
110: black holes that can be produced at the LHC is ${\cal O}(100 \ {\rm GeV})$,
111: its decay should produce roughly ${\cal O}(20)$ particles with average energy
112: $\sim 3$ times the Hawking temperature. This decay would appear to be
113: instantaneous to LHC experiments, since the typical lifetime of such a black
114: hole is only ${\cal O}(10^{-27} \ {\rm s})$.
115:
116: Since these black holes decay into all SM degrees of freedom with
117: approximately equal probability, most final state particles would be strongly
118: interacting quarks or gluons. This leads to the expectation that a typical
119: black hole event at the LHC would contain 10 or so jets plus a few (charged or
120: neutral) leptons and/or photons, each with typical energy of several hundred
121: GeV. This is obviously a very dramatic signature, which should be easy to
122: detect.
123:
124: However, it has been argued by Anchordoqui and Goldberg
125: \cite{Anchordoqui:2002cp} that the density of strongly interacting partons
126: just after the decay of the black hole would be so high that they would
127: frequently interact with each other, leading to the formation of a (more or
128: less) thermalized ``chromosphere'', i.e. a (nearly) spherical shell of
129: thousands of particles with rather small energies. This final state would also
130: be very easy to detect. However, in this case the final state would be
131: characterized almost uniquely by the mass of the black hole. More detailed
132: investigations of the primary decay spectrum, which could give information
133: about the number of additional dimensions as well as the spin of the decaying
134: black hole, could then only be performed with the (rather few) primary charged
135: leptons and photons.
136:
137: It is therefore of some importance to decide whether the assertion of
138: ref.\cite{Anchordoqui:2002cp} is in fact correct. In this paper we report
139: results of a Monte Carlo simulation of the QCD effects relevant for the decay
140: of TeV black holes, including both parton showering and partonic collisions.
141: This simulation has to keep track of the space--time evolution of the black
142: hole decay products. In contrast, the usual shower codes only keep track of
143: the virtuality of the partons in the shower, but treat the shower itself as
144: instantaneous. This is quite adequate for most applications, since a parton
145: shower only lasts $10^{-23}$ seconds or so, corresponding to a spatial
146: extension of a few Fermi, many orders of magnitude below the resolution of any
147: conceivable detector. However, if a chromosphere forms at all, it should do so
148: during those $10^{-23}$ seconds. A careful treatment of the spatio--temporal
149: evolution of the shower is therefore mandatory for us.
150:
151: We find that the effects of parton--parton scatterings after black hole decay
152: are essentially negligible for a black hole of 5 TeV, and are quite small even
153: for 10 TeV black holes. Even in the latter case, most partons do not scatter.
154: Moreover, most of these (relatively rare) interactions are rather soft, i.e.
155: they do not change the energies, trajectories or virtualities of the
156: participating partons very much. We therefore conclude that the decays of
157: black holes that might be produced at the LHC will {\em not} lead to the
158: formation of a chromosphere. Black hole event generators that ignore
159: interactions between black hole decay products \cite{charybdis,catfish} only
160: make a small mistake.
161:
162: The remainder of this note is organized as follows. In the next Section we
163: describe the production of TeV black holes, and their subsequent decay through
164: Hawking radiation, in slightly more detail. This determines the initial
165: set--up of the partonic system, which later may or may not develop into a
166: chromosphere. In Sec.~3 we summarize the argument in favor of chromosphere
167: formation, closely following ref.\cite{Anchordoqui:2002cp}; we also point out
168: some weaknesses in this argument. Sec.~4 is devoted to a description of the
169: simulation program we wrote. In Sec.~5 we present numerical results for the
170: angular correlation between pairs of charged particles, for the overall energy
171: flow of the hadronic black hole decay products, and for the microscopic
172: structure of these events. Finally, Sec.~6 contains a brief summary and
173: outlook.
174:
175: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
176: \section{Black hole production and decay}
177: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
178:
179: Lower bounds on the classical gravitational production cross section of black
180: holes are obtained in~\cite{Eardley:2002re,Yoshino:2002tx,Yoshino:2005hi}.
181: Quantum corrections are estimated in~\cite{Solodukhin:2002ui,Hsu:2002bd}.
182: The implication of the correspondence principle for black holes and strings is
183: considered in~\cite{Dimopoulos:2001qe}. All these results suggest that the
184: black hole production cross section grows geometrically above the
185: (higher--dimensional) Planck scale:
186: %
187: \begin{align} \label{cs}
188: \sigma(\hat{s}) &\simeq \pi r_h(M)^2 \propto \hat{s}^{1/(1+n)} \, .
189: \end{align}
190: %
191: Here, $n$ is the number of additional spatial dimensions, $\hat{s}$ is the
192: partonic cms energy, and $r_h$ is the horizon radius of the
193: $D=4+n$~dimensional Schwarz\-schild black hole with mass
194: $M=\sqrt{\hat{s}}$:\footnote{We use the notation of \cite{Anchordoqui:2002cp},
195: where $M_D$ stands for the reduced higher--dimensional Planck mass.}
196: %
197: \begin{equation} \label{radius}
198: r_h(M) = {1 \over M_D} \left[ \frac {M} {M_D} \frac {2^n \pi^{(n-3)/2}
199: \Gamma\left({3+n\over2}\right) } {n+2} \right]^{1/(1+n)} \, .
200: \end{equation}
201: %
202: Eq.(\ref{cs}) should hold as long as $r_h$ is small compared to the size of
203: the additional dimensions, which is true for all cases of interest to LHC
204: experiments.
205:
206: Astrophysical processes lead to the very strong lower bound $M_D \gg 10$ TeV
207: for $n \leq 3$ \cite{raffelt}. For larger $n$ the lower bound on $M_D$ comes
208: from searches for the production of gravitons (including their Kaluza--Klein
209: towers) at colliders, yielding $M_D \geq 0.65$ TeV for $n=6$ \cite{pdg}. It
210: has been argued \cite{anchor3} that the non--observation of black holes
211: produced by very energetic cosmic ray neutrinos yields the slightly stronger
212: bound $M_D \gsim 1$ TeV for $n \geq 5$. However, this bound relies on
213: assumptions on the flux of very energetic cosmic neutrinos. We will see
214: shortly that chromosphere formation is most likely for the smallest allowed
215: value of $M_D$. To be conservative, we will therefore present numerical
216: results for $n=6$ and $M_D = 0.65$ TeV. In that case according to
217: Eqs.(\ref{cs}), (\ref{radius}) the LHC operating at a proton--proton cms
218: energy of 14 TeV yields cross sections in excess of 200 pb (10 fb) for $M > 5
219: \ (10)$ TeV~\cite{Giddings:2001bu}. In other words, the LHC will be a
220: veritable black hole factory if a TeV scale gravity scenario is employed by
221: nature~\cite{Giddings:2001bu,Dimopoulos:2001hw}.\footnote{QCD initial state
222: radiation is, as usual, taken into account by using scale--dependent parton
223: distribution functions, the relevant momentum scale being set by $1/r_h$
224: \cite{Giddings:2001bu}. Numerical simulations indicate \cite{Yoshino:2002tx,
225: Yoshino:2005hi} that several tens of percent of the cms energy of the
226: colliding partons may escape in form of gravitational radiation. The same
227: calculations show that Eqs.(\ref{cs}), (\ref{radius}) underestimate the
228: cross section for black hole production for fixed $\hat{s}$. Nevertheless
229: the energy ``lost'' in gravitational radiation implies $M < \sqrt{\hat{s}}$,
230: in which case the cross section for the production of black holes with a
231: given mass at the LHC might be more than three orders of magnitude smaller
232: than indicated by Eqs.(\ref{cs}), (\ref{radius}) with $\sqrt{\hat{s}} = M$
233: \cite{anchor2}. Finally, if a ``generalized uncertainty principle'' imposes
234: a lower bound on physical lengths of order $1/M_D$, the cross section for
235: producing black holes with mass $M \gg M_D$ at the LHC would also be reduced
236: by several orders of magnitude \cite{hoss1}.}
237:
238: We note that the black hole is generically produced with a sizable angular
239: momentum~\cite{Ida1}. However, here we only consider the case of vanishing
240: angular momentum. While the spin of the black hole would affect the details of
241: its decay spectrum during the early ``spin--down'' phase \cite{Ida1, kanti,
242: Ida:2006tf}, these details are not likely to significantly change the
243: importance of the QCD processes which are the main focus of our analysis.
244:
245: Once produced, the black hole radiates off its mass~$M$ via Hawking
246: radiation~\cite{Hawking:1974sw} mainly into the brane--localized standard model
247: particles~\cite{Emparan:2000rs}.\footnote{Possible enhancement effects of bulk
248: graviton emission, especially for highly rotating black holes, have been
249: discussed in \cite{enhance}.} The number spectrum for the emission of a
250: spin$-s$ particle with energy $\omega$ per unit time is
251: %
252: \begin{align} \label{spectrum}
253: {d \dot{N}_s\over d\omega} &= {1\over2\pi} {\Gamma_s\over e^{\omega/T} -
254: (-1)^{2s}}\, .
255: \end{align}
256: %
257: Here the Hawking temperature $T$ is given by
258: %
259: \begin{align} \label{temp}
260: T &= {1+n\over 4\pi r_h}\, .
261: \end{align}
262: %
263: The ``greybody factor'' $\Gamma_s(E)$ in Eq.(\ref{spectrum}) is defined to be
264: the absorption rate of the incoming flux with energy $E$ at spatial infinity:
265: %
266: \begin{align} \label{gb}
267: \Gamma_s &= {\dot{N}_\text{in} - \dot{N}_\text{out} \over \dot{N}_\text{in}},
268: \end{align}
269: %
270: when the purely in--going boundary condition is put at the black hole horizon.
271: Physically $\Gamma_s$ is, in the ``time-reversed'' sense, the proportion of
272: the radiation that passes through the gravitational potential well from the
273: horizon towards spatial infinity.
274:
275: \begin{figure}[H] \label{specfig}
276: \begin{center}
277: \includegraphics[scale=2]{number_spectra.eps}
278: \end{center}
279: \caption{The energy spectrum of gluons (upper, black) and quarks (lower, grey)
280: emitted in black hole decay for $n=6$ additional dimensions. $\omega$ is the
281: energy of the emitted particle, and the Schwarzschild radius $r_h$ is given
282: by Eq.(\ref{radius}).}
283: \end{figure}
284:
285: \setcounter{footnote}{0}
286: Numerical results \cite{Ida:2006tf} for the spectra of quarks and gluons
287: produced in the decay of a black hole are shown in Fig.~1. In principle
288: Eqs.(\ref{spectrum}) and (\ref{temp}) predict that the spectrum of black hole
289: decay products changes with time: as more energy is radiated off, the mass of
290: the black hole becomes smaller. Eq.(\ref{radius}) shows that this also reduces
291: its radius, which according to Eq.(\ref{temp}) increases its
292: temperature.\footnote{Since Fig.~1 effectively shows the spectrum in units of
293: the inverse black hole radius, it includes all the information required for
294: including this effect in the simulation. However, these semi--classical
295: results only hold for $M \gg \omega$; they may therefore not describe the
296: late stages of black hole decay adequately. In fact, it has been argued
297: \cite{remnant} that a black hole will not evaporate completely, but leave
298: behind a stable or at least long--lived, and possibly charged, remnant with
299: mass $\sim M_D$. We do not consider this possibility in our work.} For
300: simplicity we employ the ``sudden decay approximation''
301: \cite{Dimopoulos:2001hw} where the entire decay spectrum is calculated using a
302: fixed black hole mass.
303:
304: All degrees of freedom with a given spin are emitted with equal probability in
305: black hole decay; however, Fig.~1 shows that the greybody factors do depend on
306: the spin. In order to include this effect, we define the integrals over the
307: decay spectra
308: %
309: \begin{equation} \label{is}
310: I_s = \int_0^\infty \frac {d \dot{N}_s} {d \omega} d z \, ,
311: \end{equation}
312: %
313: where $z = \omega r_h$ is the dimensionless variable shown in Fig.~1.
314: Numerically, $I_{1/2} \simeq 0.115, \, I_1 = 0.155$. In our simulation we are
315: primarily interested in the question whether quarks and gluons produced in
316: black hole decay will thermalize. Since scattering cross sections for heavy
317: quarks are smaller than those for massless quarks, we conservatively assume
318: that only $u,d,s,c$ quarks and gluons are emitted in black hole decays. The
319: relative abundance of gluons is then on average given by
320: %
321: \begin{equation} \label{gluons}
322: \frac {\langle N_g \rangle} {\langle N_g \rangle + \langle N_q \rangle} =
323: \frac {8 I_1} {8 I_1 + 24 I_{1/2}} \, .
324: \end{equation}
325: %
326: An analogous argument shows that about 75\% of the black hole mass will be
327: radiated into quarks and gluons.\footnote{Massive gauge and Higgs bosons
328: produced in black hole decay frequently also decay into quarks. However,
329: these bosons have much longer lifetimes than the black hole itself. The
330: quarks produced in their decays therefore come too late to contribute to the
331: formation of a chromosphere, although they might get trapped in the
332: chromosphere \cite{Anchordoqui:2002cp} should it indeed form.}
333:
334: The total (average) number of particles produced in the decay of a single
335: black hole can be estimated from the integrals
336: %
337: \begin{equation} \label{its}
338: \tilde I_s = \int_0^\infty z \frac {d \dot{N}_s} {d \omega} d z \, ;
339: \end{equation}
340: %
341: numerically, $\tilde I_{1/2} = 0.181, \, \tilde I_1 = 0.231$. The average
342: energies (in units of $r_h^{-1}$) of the produced particles are given by the
343: ratios $\tilde I_s / I_s$. Numerically, the average energy of a quark or gluon
344: is $\langle E_q \rangle = 313 \ (283)$ GeV and $\langle E_g \rangle = 296 \
345: (268)$ GeV for $M = 5 \ (10)$ TeV. Using the fact that about 75\% of the total
346: energy goes into hadrons, this gives average parton multiplicities
347: %
348: \begin{eqnarray} \label{nav}
349: \langle N_q \rangle &=& 8.4 \ (18.6) \nonumber \\
350: \langle N_g \rangle &=& 3.8 \ (8.3) \, ,
351: \end{eqnarray}
352: %
353: for $M = 5 \ (10)$ TeV, where $N_q$ counts both quarks and antiquarks.
354: Choosing a larger higher--dimensional Planck scale $M_D$ would lead to higher
355: Hawking temperature $T$, and hence to fewer, more energetic, primary black
356: hole decay products; this would obviously reduce the probability of
357: parton--parton scattering.
358:
359: Finally, Eq.(\ref{spectrum}) also allows to compute the lifetime of the black
360: hole. To that end, one computes the time derivative of the mass of the black
361: hole,
362: %
363: \begin{equation} \label{dmdt}
364: \frac {dM} {dt} = \int_0^\infty \omega \frac {d \dot{N} } {d \omega} d
365: \omega\, .
366: \end{equation}
367: %
368: If we set $t=0$ for the time of black hole production, its lifetime $\tau_{\rm
369: bh}$ is defined by $M(\tau_{\rm bh})$ = 0. Eq.(\ref{spectrum}) shows that
370: the rhs of Eq.(\ref{dmdt}) is $\propto r_h^{-2}$. The use of Eq.(\ref{radius})
371: then leads to \cite{Giddings:2001bu}
372: %
373: \begin{equation} \label{life}
374: \tau_{\rm bh} = \Gamma_{\rm bh}^{-1} = \frac {C} {M_D} \left(\frac {M} {M_D}
375: \right)^{(n+3)/(n+1)}\, .
376: \end{equation}
377: %
378: The coefficient $C$ can be computed by numerically integrating the spectra
379: shown in Fig.~1, leading to $C \simeq 0.2$. Eq.(\ref{life}) then
380: gives $\Gamma_{\rm bh} = 240\ (100)$ GeV for $M= 5 \ (10)$ TeV and $M_D = 0.65$
381: TeV, corresponding to lifetimes of order $10^{-27}$ seconds.
382:
383: Having presented the relevant properties of the partonic final state that
384: emerges from black hole decay, we are now ready to discuss whether
385: interactions between these partons might lead to formation of a chromosphere.
386:
387: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
388: \section{Arguments for and against chromosphere formation}
389: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
390:
391: The argument by Anchordoqui and Goldberg \cite{Anchordoqui:2002cp} starts from
392: the observation that the parton number density $n$ after black hole decay is
393: quite high: roughly ${\cal O}(10)$ partons are distributed over a sphere with
394: radius $\sim c \tau_{bh}$. They estimate the rate $\Gamma$ of bremsstrahlung
395: reactions (which increase the total number of partons) as $\Gamma = n c
396: \sigma_b$, with
397: %
398: \begin{equation} \label{brems}
399: \sigma_b \simeq \frac {8 \alpha_s^3} {\Lambda^2} \ln \left( \frac {2Q}
400: {\Lambda} \right) \, ,
401: \end{equation}
402: %
403: where $\alpha_s$ is the strong coupling constant and $Q$ the initial energy of
404: the parton in the rest frame of the black hole; the energy scale $\Lambda$ is
405: estimated as the inverse of the radius of the expanding shell of partons. This
406: leads to a total interaction rate per parton \cite{Anchordoqui:2002cp}
407: %
408: \begin{equation} \label{intrate}
409: {\cal N}_{\rm int} \simeq 0.15 \frac {N_{q,{\rm init}}} {10} \left( \frac
410: {\alpha_s(Q_{\rm min}) } {0.2} \right)^3 \ln \left( \frac {2 Q} {Q_{\rm min}}
411: \right) \ln \left( \frac {\Gamma_{\rm bh}} {Q_{\rm min}} \right) \, .
412: \end{equation}
413: %
414: For $Q \simeq 400$ GeV and initial parton number $N_{q,{\rm init}}=10$
415: Eq.(\ref{intrate}) predicts about 3 interactions per parton even with minimal
416: momentum transfer $Q_{\rm min} = 9$ GeV; this increases to $\sim 30$
417: interactions for $Q_{\rm min} =1.8$ GeV (the mass of the $\tau$
418: lepton). Anchordoqui and Goldberg argue that partons that interact so
419: frequently must thermalize, leading to the formation of an expanding shell of
420: particles with approximately thermal energy distribution.\footnote{In
421: contrast, the authors of ref.\cite{Giddings:2001bu} state that the shell of
422: black hole decay products is too thin to thermalize; however, no
423: quantitative estimate of the effects of interactions between these decay
424: products is given there.}
425:
426: This argument has several weaknesses. To begin with, most QCD interactions are
427: rather soft, as seen by the factor $1/\Lambda^2$ in Eq.(\ref{brems}).
428: Similarly, most emitted gluons will be soft and/or collinear; this leads to
429: the logarithmic factor in the cross section (\ref{brems}). In other words, a
430: ``typical'' interaction may not change the energies and momenta of the
431: participating partons very much, and may only lead to the emission of a soft
432: and/or collinear gluon. Such interactions would {\em not} impede the formation
433: of well--defined hadronic jets. In fact, during a QCD parton cascade, many
434: gluons will in any case be emitted from the partons produced in a given
435: ``hard'' process; most of these gluons will also be soft and/or collinear.
436: This gives rise to finite widths and masses of hadronic jets, but does not
437: destroy them.
438:
439: Secondly, Anchordoqui and Goldberg seem to have overlooked the fact that a
440: scattering reaction takes a finite time: according to the uncertainty
441: principle, the time at which a reaction with energy exchange of ${\cal
442: O}(\Lambda)$ occurs can only be determined with an intrinsic uncertainty of
443: order ${\cal O}(1/\Lambda)$. Most reactions have $\Lambda \sim r^{-1}$
444: \cite{Anchordoqui:2002cp}, $r$ being the radius of the expanding shell of
445: particles. Loosely speaking, a particle ``has time'' for only one such
446: reaction while traveling a distance ${\cal O}(r)$.
447:
448: The formation of a chromosphere seems unlikely on purely phenomenological
449: grounds. Note that the number of interactions per parton is proportional to
450: the number of initial partons. If ten initial partons lead to a chromosphere,
451: five or six initial partons should at least show significant effects from
452: these interactions. The observation of events with six well--defined jets has
453: been reported by both the UA2 \cite{ua2} and CDF \cite{cdf} collaborations.
454: CDF finds fair agreement between observations and QCD parton shower
455: simulations (based on leading--order matrix elements). They demand that the
456: total invariant mass of the 6--jet system exceed 520 GeV, while each jet
457: should have a transverse momentum of at least 20 GeV. Using $N_{\rm init} =
458: 6$, $Q = 30$ GeV and $\Gamma_{\rm bh} \longrightarrow M_{6-{\rm jet}} = 500$
459: GeV setting the scale for the initial hard reaction, Eq.(\ref{intrate})
460: predicts ${\cal N} \simeq 0.7 \ (9.5)$ interactions per parton for $Q_{\rm
461: min} = 9 \ (2)$ GeV. We find it difficult to believe that such high
462: interaction rates would leave no imprint in the properties of the observed
463: events, given that 3 (30) interactions are supposed to lead to nearly complete
464: thermalization.
465:
466: Even if a nearly thermal chromosphere does not form, parton--parton scattering
467: might still have significant impact on the hadronic final state from black
468: hole decay. Given the complexity of QCD processes even in the absence of
469: parton--parton scattering, a quantitative investigation of their effect can
470: only be performed with the help of a QCD simulation program. This is the topic
471: of the next Section.
472:
473:
474: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
475: \section{Simulation}
476: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
477:
478: We saw in the previous Section that a quantitative analysis of the effects of
479: parton--parton scattering is only possible if we also treat the QCD parton
480: showers that occur whenever a large four--momentum is transmitted to strongly
481: interacting particles. This implies that we need to follow the
482: spatio--temporal development of these QCD showers. In this Section we first
483: outline the general philosophy of our approach; a more detailed description
484: of the various stages treated by our code will be given in the subsequent
485: Subsections.
486:
487: Numerical codes that simulate parton showers are probabilistic, i.e. they
488: operate with squared amplitudes. Quantum mechanical interference effects can
489: therefore only be treated approximately (e.g. through angular ordering
490: \cite{order}, whereby subsequent gluons are emitted at smaller and smaller
491: angles.) The basic idea is that partons emerging from a hard reaction
492: (scattering or decay) initially have time--like virtuality, which is reduced
493: by ``branching off'' additional partons. This can be justified by the
494: observation that in a complete calculation of the relevant Feynman diagrams,
495: final state partons emitting additional partons indeed have to have time--like
496: virtualities. The beauty of such showering algorithms is that they exploit QCD
497: factorization theorems to sum such higher order processes to all orders in
498: perturbation theory, albeit (usually) only with leading logarithmic accuracy.
499:
500: Most of these codes follow the ``evolution'' of this shower not in time, but
501: in an energy variable, which in the simplest case is given by the virtuality
502: of the partons in the shower; since we are dealing with final--state showers,
503: the partons in question have time--like momenta, i.e. the shower has the same
504: kinematics as a cascade of two--body decays. (In this analogy, the lifetime of
505: the decaying particles would be given by the inverse virtuality of the
506: particles in the shower.)
507:
508: According to quantum mechanics, one cannot simultaneously determine this
509: shower energy scale and the (proper) time of a branching. Unfortunately for
510: our application we need to do precisely that. A certain additional abuse of
511: the principles of quantum mechanics is therefore inevitable. We do this by
512: identifying the {\em uncertainty} in time, as determined by the uncertainty
513: principle, with the actual {\em duration} of a given process. We use this
514: identification both for the branching and for parton--parton scattering; in
515: the former case, the time is given by the inverse of the virtuality of the
516: branching parton, whereas the time needed for a scattering is given by the
517: (space--like) virtuality of the parton exchanged in this scattering reaction.
518: Note that neither two branching steps, nor two scatterings, involving the same
519: partons can occur at the same time. In the absence of scattering, the
520: evolution in time would therefore strictly match the evolution towards smaller
521: virtualities. Moreover, between branching or scattering events, the partons
522: are taken to propagate along their classical paths. Note that this evolution
523: is nonetheless probabilistic, since after each branching or scattering the
524: 4--momenta of the outgoing partons are chosen randomly, with distributions
525: determined by perturbative QCD (and subject to energy--momentum conservation).
526:
527: We only aim at leading logarithmic accuracy. This means that we only use
528: leading order cross sections, and only include $2 \rightarrow 2$ scattering
529: processes; later gluon emission, which is treated explicitly in the estimates
530: of interaction rates in ref.\cite{Anchordoqui:2002cp}, is taken into account
531: by the subsequent shower evolution. In fact, including $2 \rightarrow 3$
532: processes in the scattering reactions would lead to double counting.
533: Scattering reactions can nevertheless increase the particle multiplicities,
534: since they can {\em increase} the virtuality of the participating particles;
535: in contrast, each branching reduces this virtuality. Moreover, the scattering
536: can change the 3--momenta of the particles, thereby potentially destroying the
537: jet structure. As mentioned in the previous Section, one needs large
538: scattering angles in order to establish a chromosphere from an initially small
539: number of very energetic partons.
540:
541: As usual, we treat showering and hard scattering independently, i.e. we apply
542: QCD factorization. This requires that the scattering indeed be sufficiently
543: hard; that is, the absolute value of the four--momentum exchanged in the
544: scattering reaction must be (much) larger than the virtualities of the partons
545: in both the initial and final state. However, as already noted, the partons in
546: the final state may be more off--shell than those in the initial state. Note
547: also that we should not include initial state showering in our approach, since
548: each ``initial state'' of a scattering reaction is part of the extended final
549: state shower that follows the Hawking evaporation of the black hole; including
550: initial state radiation would therefore also lead to double counting.
551:
552: Our simulation starts with a rather small number of energetic, and far
553: off--shell, partons, using the results of Section~2. It then uses small time
554: steps to follow the parton shower and/or scattering of all partons. This phase
555: ends when all partons are (nearly) on--shell and far apart from each
556: other, so that neither further branching nor further scattering is possible.
557: When the particles virtualities reach the QCD scale $\Lambda_{QCD}$,
558: hadronization will take place and the data of the final particles will be
559: stored for statistical analysis.
560:
561: The simulation code has been written nearly from scratch in C++, although the
562: global structure is based on the VNI 4.12b Monte Carlo simulation
563: \cite{Geiger:1998fq, Bass:1999pv} using its particle record and a selection of
564: modified routines from it. The VNI particle record uses the ``Les Houches''
565: format, extended to hold information necessary for the full space--time
566: evolution of the partons. VNI 4.12b was originally written in order to
567: simulate ultra--relativistic heavy--ion collisions; in that case, multiple
568: partonic scattering reactions are certainly important, and had to be modeled
569: carefully, making this code a good starting point for our work. In contrast,
570: scattering in the final state is thought to be unimportant for hard reactions
571: involving $e^\pm$ and/or $p/\bar p$ in the initial state.
572:
573: We also use some routines from the PYTHIA Monte Carlo simulation
574: \cite{Sjostrand:2003wg}. For numerical integration and a simulation grade
575: random number generator the GNU Scientific Library \cite{gnu} for C/C++ was
576: used.
577:
578: \subsection{Initial setup}
579: \setcounter{footnote}{0}
580:
581: As in ref.\cite{Anchordoqui:2002cp} we randomly distribute the initial partons
582: inside a shell with thickness equal to the black hole lifetime $\tau_{bh}$
583: (\ref{life}) around the decayed black hole with radius \eqref{radius}. Quarks
584: and gluons are generated separately, with average multiplicities given by
585: Eq.(\ref{nav}) and energy distributions according to Fig.~1. The momenta of
586: most partons are chosen randomly inside that half of the solid angle which
587: points away from the black hole, as seen from the location at which the
588: particle is created. Choosing the partons to be always emitted radially by the
589: black hole, which would be proper for non--rotating black hole, would make
590: future collisions between them impossible in the absence of showering; our
591: choice therefore increases the possibility of parton--parton
592: scattering.\footnote{Non--radial emission should occur during the spin--down
593: phase, with quite complicated angular dependence
594: \cite{Ida1,kanti,Ida:2006tf}. Since most of the energy is released after the
595: black hole has shed its spin, our treatment most likely over--estimates the
596: probability of parton--parton scattering.} The 3--momentum of the last
597: parton is taken such that the total 3--momentum in the black hole rest frame
598: is zero.\footnote{This is not always possible, given that the energy of this
599: parton has already been fixed and its virtuality must be smaller than this
600: energy. If no solution is found, the event is abandoned, and the routine for
601: generating the initial set--up makes a new attempt.} This is not absolutely
602: necessary, because the 3--momenta of the strongly interacting particles alone
603: do not have to add up to zero, but it allows a good control of the simulation.
604:
605: Next, quark flavors are assigned randomly. We do not include the top and
606: bottom quark because we cannot assume them to be massless in all possible
607: collisions when we are using the massless QCD scattering amplitudes; indeed,
608: top production is likely to be somewhat suppressed since for our choice of
609: parameters, the Hawking temperature is somewhat below $m_t$. Charge
610: conservation is not enforced in our initial setup taking into account that the
611: black hole also radiates off other particles like leptons. However we take
612: care that the charges add up to some integer.
613:
614: We set the maximal initial virtualities of the partons equal to their
615: energies; the actual values of these virtualities will be chosen by the shower
616: algorithm described in a subsequent Subsection. This choice of maximal
617: virtuality reproduces features of hadronic $Z$ decays fairly accurately, when
618: starting from leading order, $Z \rightarrow q \bar q$, decays.\footnote{One
619: might argue that this choice is not appropriate for black hole decay, since
620: black holes are not pointlike, unlike $Z$ bosons. However, for choices of
621: $M_D$ and $M$ relevant for LHC phenomenology, the scales $1/r_h$ or
622: $\Gamma_{bh}$ that describe the spatio--temporal extension of the black hole
623: are quite close to the average energies of the decay partons; note also that
624: the final results will only depend logarithmically on the maximal initial
625: virtuality.}
626:
627:
628: After the initial setup the program enters the main loop, which simulates the
629: space--time evolution of the parton cascade. This evolution is determined by
630: two processes, parton scattering and branching, which are described in the
631: next two Subsections.
632:
633: %----------------------------------------------------------------------
634:
635: \subsection{Parton scattering}
636: \setcounter{footnote}{0}
637:
638: Every possible pair of partons is boosted from the black hole rest frame into
639: its cms frame, and it is checked if it has reached its closest possible
640: distance. In the next step it is checked whether partons which reached their
641: closest distance undergo a collision. We are using the cascade approach
642: \cite{Geiger:1991nj} for this purpose, according to which a collision takes
643: place if the closest distance of the parton pair is within the radius defined
644: by the total cross section of the specific process,
645: %
646: \begin{equation} \label{collision}
647: |\mathbf r_{a} - \mathbf r_{b}|_{\rm min} \leq \sqrt{ \frac
648: {\hat{\sigma}_{ab}} {\pi}} \, .
649: \end{equation}
650: %
651:
652: If there is more than one possible scattering channel for two partons the
653: total cross section will be the sum of the cross sections for all possible
654: final states,
655: %
656: \begin{equation} \label{sumcross}
657: \hat{\sigma}_{ab} = \sum_{c,d} \hat{\sigma}_{ab \rightarrow cd} \, .
658: \end{equation}
659: %
660: The individual cross sections are calculated numerically by integrating the
661: corresponding differential cross sections
662: %
663: \begin{equation} \label{diffcross}
664: \hat{\sigma}_{ab \rightarrow cd} = \int_{\hat{t}_{\rm min}}^{\hat{t}_{\rm max}}
665: \left(\frac{d\hat{\sigma}(\hat{s}, \hat{t}, \hat{u})} {d\hat{t}} \right)_{ab
666: \rightarrow cd} d\hat{t} \, ,
667: \end{equation}
668: %
669: with Mandelstam variables $\hat{s} = (p_a + p_b)^2$, $\hat{t} = (p_a -
670: p_c)^2$, $\hat{u} = (p_a - p_d)^2$. The choice of the upper and lower bounds
671: $\hat{t}_{\rm max}$, $\hat{t}_{\rm min}$ in equation \eqref{diffcross}
672: requires some care. When two on--shell partons are scattering, the matrix
673: elements for the ($2 \rightarrow 2$) QCD cross sections diverge for forward
674: ($\hat{t} \rightarrow 0$) and/or backward ($\hat{u} \rightarrow 0$)
675: scattering, thus a minimal momentum transfer is needed that determines
676: $\hat{t}_{\rm max}$ and $\hat{t}_{\rm min}$ for given $\hat{s}$. In this case
677: we take the commonly used value (in the parton--parton cms) of $p_{\perp {\rm
678: min}} = 1$ GeV; this requires $\sqrt{\hat{s}} > 2$ GeV. Collisions with
679: $p_\perp < 1$ GeV are considered to be soft and are not evaluated since only
680: collisions generating high transverse momentum can change the jet structure of
681: the event.
682:
683: Collisions with at least one virtual particle in the initial state need
684: special treatment, since off--shell initial or final partons lead to non gauge
685: invariant ($2 \rightarrow 2$) amplitudes \cite{Geiger:1991nj}. The authors of
686: \cite{Geiger:1991nj} solve this problem by combining the scattering with
687: space-- or time--like branching in a single (rather large) time step so that
688: at the end of a scattering only on--shell particles are left and only
689: on--shell particles will scatter again. This should be sufficient for the
690: simulation of heavy ion collisions, where (nearly) all virtualities and
691: exchanged 4--momenta are quite small. However, in our case this procedure
692: would allow a parton to instantaneously\footnote{Our program uses much shorter
693: time steps than the original VNI code, typically $\sim 2 \cdot 10^{-4}$
694: GeV$^{-1}$ initially. We checked that choosing even shorter steps does not
695: change the result.} shower off a virtuality of hundreds of
696: GeV and then undergo a collision with momentum exchange of only a few GeV.
697: This does not make sense, since such a relatively soft collision would take
698: much more time than the (initial part of) the showering; moreover, the (many)
699: particles produced in the shower might undergo scatterings of their own.
700:
701: We therefore take another approach: we allow virtual particles to take part in
702: a collision only if the scattering scale $Q^2_{\rm scatt}$ is at least as high
703: as half the sum of the virtualities $Q^2_a$, $Q^2_b$ of the particles $a$ and
704: $b$ in the initial state,
705: %
706: \begin{equation} \label{qscale}
707: Q^2_{\rm scatt} \geq \frac{Q^2_a + Q^2_b}{2} \, .
708: \end{equation}
709: %
710: In this case the scattering will at least not take longer than the parton
711: shower up to that point. Moreover, it can be hoped that the scattering is hard
712: enough that the virtuality of the particles in the initial state becomes
713: irrelevant, so that we can describe the scattering using massless ($2
714: \rightarrow 2$) QCD cross sections. In fact, basically the same condition is
715: chosen by the usual QCD simulators when setting the showering scale for
716: initial state radiation, although in these programs the scattering scale
717: $Q^2_{\rm scatt}$ is fixed first. In our simulation the scattering scale is
718: taken to be
719: %
720: \begin{equation} \label{qscatt}
721: Q^2_{\rm scatt} = \frac {\hat{t} \hat{u}} {\hat{s}}
722: \simeq \mathbf p_{\perp}^2 \, .
723: \end{equation}
724: %
725: This also fixes the scale in the running strong coupling constant,
726: %
727: \begin{equation} \label{alphas}
728: \alpha_s(Q^2_{\rm scatt}) = \frac{12\pi}{25\ \log(\frac{Q_{\rm scatt}^2}
729: {\Lambda_{QCD}^2})} \, ,
730: \end{equation}
731: %
732: where we took $N_F = 4$ active flavors and QCD scale $\Lambda_{QCD} = 0.2$ GeV.
733: The requirement \eqref{qscale} determines the boundaries of the
734: integration over $\hat{t}$ in \eqref{diffcross}:
735: %
736: \begin{equation} \label{tbounds}
737: \hat{t}_{\rm min, \, max} = \frac{1}{2} \left( -\hat{s} \mp \sqrt{\hat{s}^2 -
738: 4Q^2_{ab}\hat{s}} \right) \, ,
739: \end{equation}
740: %
741: where $Q^2_{ab} = \frac{Q^2_a + Q^2_b}{2}$.
742:
743: We did not include ($2 \rightarrow 1$) fusion processes like $g + g
744: \rightarrow g^*$ (with $g^*$ being off--shell), since the first branching of
745: the produced off--shell parton would again result in a $2 \rightarrow 2$
746: process, leading to double counting. It could be argued that we should also
747: require $Q_{\rm scatt} \geq 1 / |\mathbf r_{a} - \mathbf r_{b}|_{\rm min}$.
748: However, this would exclude interactions between partons that happen somewhat
749: before or after the time of their closest approach, which seems unphysical. We
750: therefore do not impose this additional requirement, which would reduce the
751: number of partonic scatterings significantly. In fact, the opposite
752: requirement, $Q_{\rm scatt} < 1 / |\mathbf r_{a} - \mathbf r_{b}|_{\rm min}$,
753: seems more reasonable, since according to the uncertainty principle a highly
754: virtual exchanged parton can only travel a very short distance. We do not
755: impose this constraint, i.e. we allow very large momentum exchange also
756: between relatively distant partons. This again increases the importance of
757: collisions, since a large $Q_{\rm scatt}$ is required for a collision to
758: modify the jet structure. However, we will see that not imposing this upper
759: bound on $Q_{\rm scatt}$ has very little effect in practice, since high values
760: of $Q_{\rm scatt}$ are in any case very unlikely.
761:
762: In some rare cases one parton has two possible collision partners in a single
763: time step. In that case for both partners cross section and distance will be
764: checked. If only one collision takes place this will be the one evaluated. If
765: both collisions would take place only the one with the larger cross section
766: will be evaluated.
767:
768: After a pair of particles has been identified as colliding, the scattering
769: kinematics can be generated. For initial states which have more than one
770: channel, the final state is chosen according to the probability given by the
771: relevant total cross sections. Next the value of $\hat{t} \in \left[
772: \hat{t}_{\rm min}, \, \hat{t}_{\rm max} \right]$ is generated, with
773: distribution given by the corresponding differential cross section $d \hat
774: \sigma/ d \hat t$ (normalized to unity). The initial particle pair is boosted
775: into its cms frame and put on--shell. The absolute value of the transverse
776: momentum of the final particles is given by $Q^2_{\rm scatt}$ as determined by
777: the chosen value of $\hat{t}$. The azimuthal scattering angle, which fixes the
778: direction of the $\mathbf{p_\perp}$ vectors, is chosen randomly, with flat
779: distribution between 0 and 2$\pi$.
780:
781: Finally the virtualities of the outgoing particles are created by the
782: branching routine (see the next Subsection) and color charges are assigned
783: according to the color flow of the specific process. It should be noted that
784: in case of the scattering of virtual particles, the virtuality after
785: scattering can become even lower than before, since $Q_{\rm scatt}$ is the
786: {\em maximal} virtuality of the particles in the final state. We see no
787: physical reason why such virtuality--reducing scattering reactions should be
788: suppressed.
789:
790: Due to the lifetime of a virtual (exchanged) particle the scattering will take
791: a finite amount of time determined by the scattering scale. In the rest frame
792: of the black hole this time is given by
793: %
794: \begin{equation} \label{tscatt}
795: \tau_{\rm scatt} = \frac{\gamma_{ab}}{Q_{\rm scatt}} \, ,
796: \end{equation}
797: %
798: where $\gamma_{ab} = (E_a + E_b) / \sqrt{\hat{s}_{ab}}$ is the boost factor
799: between the cms of colliding partons $a$ and $b$ and the black hole rest
800: frame. During this time the colliding particles are not allowed to collide
801: again, nor can they branch off additional partons; in fact, in quantum
802: mechanics one cannot say whether the parent partons $a,b$ or the children
803: $c,d$ exist during the period $\tau_{\rm scatt}$. After the scattering time is
804: over the partons are able to initiate final state radiation or collide
805: again.\footnote{In principle it might be more appropriate to place this ``dead
806: time'' symmetrically around the time of closest approach, rather than
807: letting it start at the time of closest approach. However, this would be
808: technically difficult, since the program would then have to go back in time,
809: and ``un--do'' any possible branchings that happened in the ``dead'' period
810: before the time of closest approach. Since by construction $Q_{\rm scatt}$
811: is larger than the virtuality of the scattering particles, this asymmetric
812: placement of the ``dead time'' should not matter much in practice.}
813:
814: %-----------------------------------------------------------------------
815:
816: \subsection{Parton branching}
817: \setcounter{footnote}{0}
818:
819: While the inclusion of partonic collisions is the main new ingredient of our
820: simulation, we will see that parton branching plays a far more important role
821: in determining the characteristics of the final state. We model this using a
822: modified version of the relevant routine of VNI 4.12b, which in turn is based
823: on the PYTHIA branching algorithm \cite{Sjostrand:2003wg, Bengtsson:1986et,
824: Webber:1986mc}.
825:
826: The modifications implemented by us address the need to do branching for a
827: single parton, instead of treating two partons at once as is done normally in
828: order to ensure 4--momentum conservation. To make sure we can conserve energy
829: and momentum for a single parton we always have to determine the virtualities
830: at which the parton branches one step ahead in the branching algorithm: the
831: kinematics of $a \rightarrow b + c$ can only be fixed if the ``masses'', i.e.
832: (time--like) virtualities, of all three participating partons are known. Our
833: version of the routine therefore determines the virtualities of $b$ and $c$
834: earlier than the original routine does. Note that these actual virtualities
835: are typically much smaller than the corresponding maximal values; this is the
836: reason why energetic partons, with initial {\em maximal} virtualities given by
837: their energies, typically produce quite narrow jets. Another modification is
838: the introduction of the scattering--induced ``dead time'' described at the end
839: of the previous Subsection, during which partons are not allowed to branch.
840:
841: Just like scattering reactions, the branching $a \rightarrow b + c$ also takes
842: a finite amount of time in our simulation, given by
843: %
844: \begin{equation} \label{tbranch}
845: \tau_{\rm branch} = \frac{\gamma_a}{Q_a} = \frac{E_a}{Q_a^2} \, ,
846: \end{equation}
847: %
848: with $\gamma_a = E_a/Q_a$ describing the boost from the rest frame of parton
849: $a$ to that of the black hole. In our simulation this is treated like the
850: lifetime of a decaying particle, i.e. the probability of an ``active'' parton
851: $a$ to branch during the time step $dt$ is given by \cite{Geiger:1998fq,
852: Bass:1999pv}
853: %
854: \begin{equation} \label{pbranch}
855: P_{\rm branch} = 1 - {\rm e}^{-dt/\tau_{\rm branch}} \simeq \frac {dt}
856: {\tau_{\rm branch}} \, ,
857: \end{equation}
858: %
859: where the second, approximate equality holds for the (realistic) situation $dt
860: \ll \tau_{\rm branch}$. Note that each parton is checked for possible
861: scattering before it is checked for branching. In our treatment partons $b,c$
862: are created instantaneously if a branching occurs, i.e. they are allowed to
863: undergo scattering immediately.\footnote{These partons are created at slightly
864: different locations, as determined by the uncertainty principle; moreover,
865: they are moving away from each other. These two partons can therefore not
866: scatter on each other, even if they are very close, but they can scatter on
867: other partons.} This can be justified from the requirement (\ref{qscale}),
868: which ensures that scattering reactions are much faster than branchings.
869: Imposing a branching ``dead time'' of (some fraction of) $\tau_{\rm branch}$
870: on $b,c$ before they are allowed to scatter would obviously reduce the number
871: of parton--parton scatterings, although numerically this effect is not very
872: large. The 4--momenta of partons $b,c$ are chosen as in the PYTHIA branching
873: algorithm \cite{Sjostrand:2003wg}.
874:
875: There are three situations in which the branching routine is used. First, all
876: initial partons which have some maximum virtuality from black hole decay are
877: sent through the branching routine once, in order to determine their actual
878: starting virtuality for the simulation. Secondly, the normal use when every
879: time step all partons which have any time--like virtuality left, and which did
880: not scatter recently, are sent through the routine to undergo branching with
881: probability given by Eq.(\ref{pbranch}). Third, similar to the first case, the
882: two partons coming out of a hard scattering which took place at the scale
883: $Q^2_{\rm scatt}$ are sent through the branching routine to determine the
884: actual virtualities they will start to propagate with.
885:
886: The next step after the parton branching algorithm is to propagate all partons
887: freely by one time step. Then the simulation calculates the average distance
888: between all pairs of partons to see if it exceeds the QCD scale
889: $1/\Lambda_{QCD}$, at which point the partonic simulation loop can be aborted.
890: If the abort condition is met hadronization will take over, or else time is
891: increased by one time step and the simulation loop starts again. For details of
892: the hadronization algorithm we again refer to the documentation of PYTHIA
893: \cite{Sjostrand:2003wg}.
894:
895: %------------------------------------------------------------------------
896:
897: \section{Results}
898:
899: We are now ready to present some numerical results. As stated in Sec.~2, we
900: set the higher--dimensional Planck scale to its lower bound, $M_D = 0.65$ TeV,
901: since this maximizes the number of initial partons, and hence the number of
902: parton--parton interactions within a given black hole decay. We also consider
903: relatively heavy black holes, with $M = 5$ and 10 TeV. Recall that decays of
904: heavier black holes are characterized by lower temperatures, and hence higher
905: multiplicities; however, the production cross section for black holes with
906: mass exceeding 10 TeV is certainly negligible at the LHC.
907:
908: Since the initial set--up is created totally randomly there are large
909: event--to--event fluctuations of the number and energies of the initial
910: partons. One therefore needs sufficient statistics to make reliable statements
911: about average quantities; the results presented below are based on 200 events
912: for each run. In order to illustrate the effects of parton--parton
913: interactions, we made separate runs for each black hole mass where these
914: interactions were turned on or turned off. In the latter case the
915: spatio--temporal evolution of the parton shower is irrelevant, i.e. our
916: program reproduces standard (PYTHIA) showering for the given set--up of
917: original partons.
918:
919: If black hole decays led to formation of a chromosphere, observables based on
920: jets would no longer be useful. In the following Subsections we therefore
921: analyze the distribution of two observables whose definition does not assume
922: the existence of jets: angular correlations between energetic charged hadrons,
923: and the overall energy flow. Of course, the observables are chosen such that
924: their distribution would be greatly affected if parton--parton scattering did
925: indeed lead to formation of a chromosphere. In the last Subsection we will
926: interpret these results with the help of the time structure of the parton
927: shower that develops after typical black hole decays.
928:
929: %---------------------------------------------------------------------
930:
931: \subsection{Angular correlation}
932: \setcounter{footnote}{0}
933:
934: After hadronization we are left with a large number of charged, long--lived
935: particles (mostly charged pions and kaons and some protons) as well as photons
936: and neutral hadrons. In this first analysis we focus on charged particles
937: because their momenta can be measured accurately using tracking information.
938: We only consider particles with an energy (in the black hole rest frame) above
939: 4 GeV; the number of charged particles passing this cut is denoted by $N_{\rm
940: ch}$. This cut should largely remove particles from the underlying event
941: which have nothing to do with black hole decay, and which have not been
942: included in our simulation.
943:
944: We compute the angles $\theta$ between any two of these charged particles;
945: altogether there are $N_{\rm pair} = N_{\rm ch}(N_{\rm ch} - 1) / 2$ such
946: pairs. The resulting distribution is binned in $\cos\theta$. Because $N_{\rm
947: ch}$ can vary a lot from event to event, we normalize the distribution for
948: each event to $N_{\rm pair}$, before averaging over the 200 generated events.
949:
950: Let us first consider some simple situations, in order to get a feeling for
951: what kind of distribution to expect. The simplest case, which certainly does
952: not describe black hole decay, would be events with two back--to--back jets,
953: each of which contains the same number $N_{\rm ch}/2$ of charged particles. In
954: this case $N^2_{\rm ch} / 4 - N_{\rm ch} / 2$ pairs would have an angle near
955: zero between the particles of the pair because they reside inside the same
956: jet; the remaining $N^2_{\rm ch} / 4$ pairs would have an angle near $\pi$
957: between the particles of the pair because the particles are in different jets.
958: In the limit of large $N_{\rm ch}$ these numbers will become almost equal,
959: leading to peaks at $1$ and $-1$ in our plot which have approximately the same
960: height.
961:
962: A case that is closer to what one may expect from the decay of a black hole
963: would be an event with $n_j$ jets. Let us keep the assumption that each of
964: these jets contains exactly the same number $N_{\rm ch}/n_j$ of energetic
965: charged particles. In this case there are
966: %
967: \begin{equation} \label{pospairs}
968: \frac{N_{\rm ch}} {2} \left(\frac{N_{\rm ch}}{n_j} - 1 \right)
969: \end{equation}
970: %
971: pairs residing in the same jet; the remaining $N^2_{\rm ch} (n_j -
972: 1) / (2n_j)$ pairs of charged particles are in different jets. We thus see
973: that the fraction of all pairs that reside inside the same jet will decrease
974: when the number of jets increases, i.e. the peak in our distribution at
975: $\cos\theta = +1$ will become smaller. Moreover, for $n_j \gg 2$ randomly
976: distributed jets the peak at $\cos\theta = -1$ will vanish because momentum
977: conservation no longer requires any two jets in the event to be
978: back--to--back. After averaging over many events all the angles between
979: different jets contribute more or less equally. In this case we therefore
980: expect a smooth distribution (although momentum conservation my still lead to
981: a slight increase of the correlation function at negative $\cos\theta$) with
982: only one peak at $\cos\theta = +1$. The distribution should remain
983: qualitatively the same in the more realistic scenario where we allow different
984: jets to contain different numbers of charged particles, although
985: eq.(\ref{pospairs}) will then no longer be valid.
986:
987: On the other hand, if we assume that a chromosphere is indeed spherical, the
988: correlation function should become almost perfectly flat; in particular, we
989: would not expect any visible peak at $\cos\theta = +1$.
990:
991: Figs.~2 show results for black hole mass $M = 5$ TeV (top) and 10 TeV
992: (bottom), with (red) and without (black) parton--parton scattering. Clearly
993: for both black hole masses there is still a strong peak at $\cos\theta = 1$,
994: leading to the conclusion that we should expect a jet structure after black
995: hole decay. The peak becomes smaller for a heavier black hole; this is
996: expected from the qualitative discussion presented above, given the increasing
997: initial parton multiplicity, see Eq.(\ref{nav}). The rise of the distribution
998: towards $\cos\theta=-1$ is also more pronounced for smaller black hole masses,
999: again as expected from our qualitative discussion. For $M=5$ TeV,
1000: parton--parton scattering has essentially no effect on this distribution. For
1001: $M=10$ TeV, it leads to a very slight broadening of the peak at
1002: $\cos\theta=+1$; however, given the uncertainties of our simulation, we do not
1003: claim that this effect is significant.
1004:
1005: \begin{figure}[H]
1006: \vspace*{-2cm}
1007: \begin{center}
1008: \includegraphics[scale=1.3]{5tev-1-new3.eps}\\
1009: \vspace*{-1.5cm}
1010: \includegraphics[scale=1.3]{10tev-1-new3.eps}
1011: \end{center}
1012: \vspace*{-5mm}
1013: \caption{The angular correlation function for charged particles from black
1014: hole decay with $E>4$ GeV, for black hole mass $M = 5$ (top) and 10 TeV
1015: (bottom), obtained by binning into 1,000 bins. The red and black curves have
1016: been obtained including and omitting parton--parton scattering,
1017: respectively.}
1018: \end{figure} \label{fig_angle}
1019:
1020: %-----------------------------------------------------------------------
1021:
1022: \subsection{Total energy flow}
1023: \setcounter{footnote}{0}
1024:
1025: The second quantity we investigate is the total energy flow from the hadronic
1026: black hole decay products. Here we envision a calorimetric measurement. We
1027: therefore divide phase space into azimuthal angle $\phi \in [0, 2\pi]$ and
1028: pseudo--rapidity $\eta \in [-4,4]$, with 15 bins in $\phi$ and 30 bins in
1029: $\eta$. For each of these 450 ``calorimeter cells'' the total visible energy
1030: is calculated, including hadrons and photons from hadronic decay, but no
1031: neutrinos or muons. We present the result by plotting the number of cells with
1032: deposited energy $E_{\rm cell} \leq E_{\rm max}$ as function of $E_{\rm max}$.
1033:
1034: Let us again first discuss the possible shapes of this distribution for
1035: different final parton configurations. If a chromosphere forms, we expect a
1036: very large number of hadrons in the final state, each of which has a
1037: relatively small energy. These would be distributed uniformly over phase
1038: space, i.e. had we defined our cells as having constant length in $\cos\theta$
1039: all cells would receive essentially the same energy. We prefer to use $\eta$
1040: to parameterize the phase space, since the energy flow pattern will then be
1041: invariant under motion of the black hole along the beam pipe. Since $d / d
1042: \eta = \sin^2(\theta) d / d \cos\theta$, cells at small $\cos\theta$, i.e.
1043: small $|\eta|$, will then receive significantly more energy than those at
1044: large $|\eta|$. There should nevertheless be almost no empty cells; the
1045: maximal energy deposit, in cells with $\eta \sim 0$, would be about $0.009
1046: E_{\rm tot}$, where $E_{\rm tot} \simeq 0.75 M$ is the total hadronic energy
1047: released in the decay of a black hole with mass M. In particular, there should
1048: not be any cells with energy comparable to the initial average partonic
1049: energy, which amounts to about 300 GeV for our choices of parameters [see the
1050: discussion of Eq.(\ref{nav}) in Sec.~2]. The distribution expected for
1051: ``ideal'' chromospheres, with energy flow being completely independent of
1052: $\phi$ and $\cos\theta$, is depicted by the blue step--like\footnote{These
1053: steps appear only for black hole decays at rest. In general one expects a
1054: smoothed-out version of these curves once a distribution of longitudinal
1055: momenta of the black holes is taken into account.} curves in Figs.~3.
1056:
1057: In the opposite extreme, where the final state consists of a relatively small
1058: number [given by Eq.(\ref{nav})] of very narrow jets, most cells would be
1059: empty, while in a few cells the deposited energy would be of order 300 GeV.
1060: However, even in the absence of parton--parton scattering, final state
1061: radiation implies that many jets will spread out over several cells. Together
1062: with the final hadronization step, this will lead to a significant number of
1063: cells in which a small, but nonzero, amount of energy is deposited, often in
1064: form of a single hadron. We therefore expect a non--trivial dependence on
1065: $E_{\rm max}$ in the entire range between $\sim 100$ MeV and 1 TeV.
1066:
1067: Our results for the energy flow of hadronic black hole decay products are
1068: shown in Figs.~3. We see that the number of cells with $E < E_{\rm max}$
1069: indeed shows nontrivial dependence on $E_{\rm max}$ over a wide range. The
1070: number of (almost) empty cells decreases with increasing black hole mass, as
1071: expected from the higher initial parton multiplicity (\ref{nav}). However,
1072: even for $M=10$ TeV, in about 50\% of the cells almost no energy is deposited;
1073: on the other hand, about twenty cells contain more than 100 GeV. We saw above
1074: that these results are consistent with the existence of well--defined jets.
1075: Parton--parton scattering has practically no effect on the number of cells
1076: containing at least 30 GeV, again indicating that it does not affect the jet
1077: structure at all.
1078:
1079:
1080: \begin{figure}[H]
1081: \begin{center}
1082: \vspace*{-2cm}
1083: \includegraphics[scale=1.3]{5tev-2-new3.eps}\\
1084: \vspace*{-1.5cm}
1085: \includegraphics[scale=1.3]{10tev-2-new3.eps}
1086: \end{center}
1087: \vspace*{-5mm}
1088: \caption{The number of phase space cells where the deposited energy satisfies
1089: $E < E_{\rm max}$ as function of $E_{\rm max}$, for black hole mass $M = 5$
1090: TeV (top) and 10 TeV (bottom). The notation is as in Fig.~2, except that the
1091: blue, step--like, curves show the prediction from an ideal chromosphere.}
1092: \end{figure} \label{fig_energy}
1093:
1094: Scattering does increase the total number of non--empty cells by about 5\%
1095: (10\%) for $M=5$ (10) TeV. However, the production of soft hadrons is a
1096: non--perturbative process, and can therefore not be treated from first
1097: principles; it is not clear whether the effect of parton--parton scattering is
1098: larger than the systematic error of our simulation.\footnote{The {\em
1099: relative} size of the effect of parton scattering should be rather
1100: insensitive to the details of the simulation; the statement that it
1101: increases the number of non--empty cells by 5 to 10\% should therefore be
1102: relatively robust. However, at the end one can only compare the {\em
1103: absolute} prediction of the simulation with actual events.} Moreover, it
1104: should be kept in mind that the ``underlying event'', which is created by the
1105: remnants of the colliding protons that do not participate in black hole
1106: formation, will also contribute a large number of (mostly soft) hadrons to the
1107: final state, which have not been included in our simulation. It is therefore
1108: not clear whether the increase of the cells containing some black hole decay
1109: products results in a measurable difference in the total energy flow in the
1110: event.
1111:
1112: %----------------------------------------------------------------------
1113:
1114: \subsection{Microscopic structure of the events}
1115: \setcounter{footnote}{0}
1116:
1117: The results of the two previous Subsections show that parton--parton
1118: interactions have little effect on the global characteristics of the hadronic
1119: final state that results from the decay of black holes with masses of a few
1120: TeV. In this Subsection we analyze the microscopic structure of the evolution
1121: of the hadronic final state that results from the decay of such black holes.
1122:
1123: In Table~1 we list the average parton multiplicities (just before
1124: hadronization), as well as the average number of parton--parton collisions,
1125: for $M=5$ and 10 TeV. We see that even in the absence of parton--parton
1126: scattering, QCD branching (final state radiation) increases the multiplicity
1127: by an order of magnitude, relative to the initial multiplicity given by
1128: Eq.(\ref{nav}). We also see that the average number of parton--parton
1129: scatterings per black hole decay is not so small; as expected from simple
1130: statistical arguments, it increases roughly quadratically with the (initial or
1131: final) partonic multiplicity. Scattering increases the final parton
1132: multiplicity by 16\% (23\%) for $M=5$ (10) TeV. This effect is even larger
1133: than that on the number of non--empty calorimeter cells. Note, however, that
1134: the number of scatterings still remains well below the number of partons even
1135: for $M=10$ TeV.
1136:
1137: \begin{table}[h]
1138: \begin{center}
1139: \begin{tabular}{|c|| c | c | c || c| c| c| c|}
1140: \hline
1141: $M$ [TeV] & \multicolumn{3}{c||}{w/o scattering } &
1142: \multicolumn{4}{c|}{with scattering } \\
1143: & $\langle n_q \rangle$ & $\langle n_g \rangle$ & $\langle n_{\rm parton}
1144: \rangle$ &$\langle n_q \rangle$ & $\langle n_g \rangle$ & $\langle n_{\rm
1145: parton} \rangle$ & $\langle n_{\rm scatt} \rangle$ \\
1146: \hline
1147: 5 & 19 & 116 & 135 & 21 & 134 & 156 & 16 \\
1148: 10 & 41 & 241 & 282 & 50 & 295 & 346 & 53 \\
1149: \hline
1150: \end{tabular}
1151: \end{center}
1152: \caption{Average final quark, gluon and total parton multiplicities from the
1153: decay of black holes with 5 and 10 TeV mass, with and without including the
1154: effect of parton--parton scattering. The last column gives the average
1155: number of partonic collisions. The statistical errors on the average
1156: multiplicities are about 2 percent.}
1157: \end{table}
1158:
1159: On the other hand, the probability $\langle n_{\rm parton} \rangle / \langle
1160: n_{\rm scatt} \rangle$ that a given parton resulted from scattering is not
1161: negligible. It rises roughly linearly with the parton multiplicity, in
1162: agreement with the estimate (\ref{intrate}). However, this expression, with
1163: $Q_{\rm min} \sim 1$ GeV as in our simulation, greatly over--estimates the
1164: number of scattering reactions even if we normalize it to the initial parton
1165: multiplicity.
1166:
1167: As mentioned earlier, parton--parton scattering can only destroy the jet
1168: structure if it involves large momentum exchange, i.e. if the scale $Q_{\rm
1169: scatt}$ defined in Eq.(\ref{qscatt}) is large. In Fig.~4 we show a scatter
1170: plot of $Q_{\rm scatt}$ values vs. time in the simulation; here entries from
1171: 100 decays of black holes with mass $M=5$ TeV have been collected. We see
1172: that the average value of $Q_{\rm scatt}$ decreases with time, once $t > 1/
1173: \langle E_{\rm parton} \rangle \sim 3 \cdot 10^{-2}$ GeV$^{-1}$. This can be
1174: understood from the observation that at early times, most partons are quite
1175: far off--shell; our condition (\ref{qscale}) then implies that early
1176: scatterings must involve rather large momentum exchange. The existence of a
1177: few early scatterings with low $Q_{\rm scatt}$ is due to the fact that the
1178: initial virtualities with which the partons are actually created by the
1179: simulation may be much smaller than their maximal values, which is given by
1180: the energies of these partons.
1181:
1182: \begin{figure}[H]
1183: \begin{center}
1184: \includegraphics[scale=1.2]{scale.eps}
1185: \end{center}
1186: \caption{Scatter plot showing the scale $Q_{\rm scale}$ defined in
1187: Eq.(\ref{qscatt}) vs. time after black hole decay, for $M=10$ TeV;
1188: scatterings from 100 decays have been added into this plot. The appearance
1189: of discrete lines at early times is an artefact of the finite time step size
1190: used.}
1191: \end{figure} \label{fig_qscale}
1192:
1193: Since the average virtualities of the partons diminish with increasing time,
1194: more scatterings with small $Q_{\rm scatt}$ become possible. Since the QCD
1195: cross sections satisfy $d \sigma / d Q^2_{\rm scatt} \propto \alpha_s(Q_{\rm
1196: scatt})^2 Q^{-4}_{\rm scatt}$ if $\hat t \hat u \ll \hat s^2$, reactions
1197: with large $Q_{\rm scatt}$ are then greatly disfavored. Moreover, the average
1198: parton energies also decrease with time. Later collisions therefore tend to
1199: have smaller $\hat{s}$, which implies reduced kinematical upper bounds on
1200: $Q_{\rm scatt}$.
1201:
1202: The time dependence of the number of scattering reactions is determined by
1203: several competing effects. While the number of partons increases due to
1204: multiple branchings as time goes on, the number {\em density} of partons
1205: decreases, making it increasingly less likely that two partons will come close
1206: to each other. As a result, the number of collisions per unit time decreases
1207: at first. On the other hand, the decreasing virtuality of the partons,
1208: and corresponding decreasing lower bound on $Q_{\rm scatt}$, means that their
1209: scattering cross section increases with time. This leads to an accumulation of
1210: relatively soft scatterings at $t \sim 0.1/$GeV.
1211:
1212: A typical decay of a 5 TeV black hole therefore develops as follows.
1213: Initially there are about 12 highly off--shell partons with average energies
1214: around 300 GeV. One or two of them may undergo scattering very soon after
1215: their creation, with typical scattering scale $Q_{\rm scatt}$ of a few tens of
1216: GeV. Note that these rather hard reactions still mostly have $Q_{\rm scatt}
1217: \ll E$, so that the direction of the participating partons is not changed very
1218: much. Moreover, these early scatterings may actually reduce the virtualities
1219: of the participating partons. As time goes on, the number of partons in the
1220: event increases, mostly due to branching processes. The number of scatterings
1221: also increases, but most of these reactions are relatively soft. We saw in
1222: Table~1 that they do increase the partonic multiplicity at the end of the QCD
1223: cascade. However, most of these scattering reactions involve partons inside
1224: the same jet, and therefore have little effect on the global quantities we
1225: analyzed in the previous Subsections. In fact, the creation of an additional
1226: very soft parton nearby in phase--space may not be visible in the (hadronic)
1227: final state at all. This explains why the results in Table~1 seem to indicate
1228: larger effects from parton--parton scattering than what we saw in Figs.~2 and
1229: 3.
1230:
1231: \section{Summary and conclusions}
1232:
1233: In this paper we investigated the question if the decays of black holes that
1234: might be produced at the LHC would be characterized by a relatively large
1235: number of discrete jets, as is usually assumed, or if they would lead to the
1236: formation of quasi--thermal ``chromospheres'' of rather soft hadrons, as
1237: suggested in ref.\cite{Anchordoqui:2002cp}. To this end we first described the
1238: partonic state produced by the decay of a black hole. We found that, if the
1239: higher--dimensional Planck mass $M_D$ is at its current lower bound, the
1240: average initial number of partons will be less than about 25 for black holes
1241: with mass $M \leq 10$ TeV, which might have a chance to be produced with
1242: appreciable rates at the LHC; higher values of $M_D$ would lead to initial
1243: states with fewer partons, and hence to less scattering.
1244:
1245: In Sec.~3 we summarized the argument in favor of formation of a
1246: chromosphere. We pointed out that this argument ignores the fact that
1247: scattering reactions take a finite amount of time. Moreover, a very similar
1248: argument should apply to ordinary QCD multi--jet events. It would predict
1249: sizable effects in six jet events, which have been studied experimentally by
1250: the UA2 and CDF collaborations, who found good agreement with standard QCD
1251: predictions which ignore interactions between the partons in the final state.
1252:
1253: Clearly a dedicated Monte Carlo study is required in order to determine the
1254: quantitative effects of such final state interactions on black hole events at
1255: the LHC. As described in Sec.~4, we wrote such a simulation code, based on the
1256: VNI program \cite{Geiger:1998fq, Bass:1999pv}. We modified it by forcing
1257: scattering reactions to take a finite amount of time, which we estimated using
1258: the uncertainty principle. A very similar argument implies that, as time goes
1259: on, the parton system evolves towards smaller virtualities through branching
1260: processes. Scattering reactions may increase the virtualities of the
1261: participating partons again, but can also be a shortcut towards smaller
1262: virtualities.
1263:
1264: Results of our simulation are described in Sec.~5. We found the effects of
1265: parton--parton scattering in the final state to be essentially negligible both
1266: for the angular correlation between energetic charged hadrons, and for the
1267: number of phase space cells containing a large amount of energy. We found some
1268: effect on the number of cells containing only one or a few soft hadrons. We
1269: interpreted these results in terms of the microscopic structure of the event.
1270: In particular, we saw that ``hard'' reactions, with large momentum exchange,
1271: only occur early on; in this case they may well reduce the virtualities of the
1272: participating partons. Later scatterings are all quite soft. At both early and
1273: late times, most scattering reactions have momentum exchange well below the
1274: energies of the participating partons; such reactions cannot significantly
1275: change the directions into which these partons are traveling. We therefore
1276: conclude that a chromosphere will {\em not} form; programs that ignore
1277: interactions between the partons from the decay of black holes created at the
1278: LHC only make a negligible mistake.
1279:
1280: Our simulation has some shortcomings. Like all QCD simulation codes, it
1281: essentially works on the level of squared matrix elements, i.e. quantum
1282: mechanical interference effects can only be included in an approximate
1283: manner. In addition, we had to identify the quantum mechanical uncertainty in
1284: time with actual duration, both for the branching and scattering processes. It
1285: is not clear to us how this limitation can be overcome, even in principle.
1286:
1287: Our results starkly contradict the claims of ref.\cite{Anchordoqui:2002cp},
1288: even though we chose our parameters and initial set--up such as to maximize
1289: the likelihood of scattering reactions between partons in the final state. It
1290: might therefore be worthwhile to summarize the three effects that have reduced
1291: the number of such reactions in our simulation.
1292:
1293: Perhaps most important is that we only allow scattering of partons that
1294: approach each other. Interactions between partons that ``always'' (after their
1295: creation) move away from each other can certainly not be treated using the
1296: factorization into initial black hole decay and subsequent parton--parton
1297: scattering that underlies our simulation; for one thing, no $S-$matrix could
1298: be defined for such a state.\footnote{In some sense such partons can
1299: nevertheless still interact with each other. For example, the well--known
1300: Coulomb singularity \cite{schwinger} for near--threshold decays into charged
1301: particles can be interpreted in terms of the electromagnetic interaction
1302: between these particles. However, in the simplest case of two--body decays,
1303: this singularity only affects the partial width for the decay; it does not
1304: increase the final state multiplicity. In case of decays into more than two
1305: partons, it can also change kinematical distributions, e.g. the invariant
1306: mass distributions of subsets of partons. However, a proper treatment of
1307: these effects would require the inclusion of corrections to black hole
1308: decays due to gauge loops. It is currently not clear how this could be
1309: done.}
1310:
1311: Strictly speaking, in our case one cannot define an $S-$matrix for partons
1312: approaching each other, either, since they were created at a finite time, i.e.
1313: they did not exist in an ``in-''state defined at time $t=-\infty$. We
1314: circumvent this problem by requiring the 4--momentum exchanged in these
1315: reactions to be larger than the initial virtualities. In this case one should
1316: be able to approximately treat the scattering like that of on--shell
1317: particles. This roughly means that we require scattering reactions to be fast
1318: on the time scale of the shower evolution up to the time of the scattering.
1319: This constraint greatly reduces the scattering cross section at early times,
1320: when the parton density is highest, and therefore reduces the number of
1321: scattering reactions. The fact that we do not allow partons participating in a
1322: scattering to start another scattering while the first one is still ``in
1323: progress'' has a much smaller effect on the number of these reactions.
1324:
1325: Although the number of parton--parton collisions remains small for all black
1326: holes that might be produced at the LHC, we saw that it increases roughly
1327: quadratically with increasing black hole mass. Extrapolating from the results
1328: of Table~1, we estimate that there will be ${\cal O}(1)$ parton--parton
1329: scatterings per produced parton once $M \gsim 25$ TeV, if the
1330: higher--dimensional Planck scale $M_D$ is kept at its lower bound of 0.65
1331: TeV. This need not lead to formation of a chromosphere, however; we saw that
1332: most scattering reactions have little effect on the jet structure of the event.
1333: Note that in this case the average initial partonic multiplicity already
1334: exceeds 100, making the reconstruction of distinct jets quite unlikely, even
1335: if such heavy black holes could ever be produced at human--made colliders.
1336:
1337: \subsubsection*{Acknowledgments}
1338:
1339: The work of K.O. is partly supported by the Special Postdoctoral Researchers
1340: Program at RIKEN.
1341:
1342: \begin{thebibliography}{99}
1343:
1344: \bibitem{Arkani-Hamed:1998rs}
1345: N.~Arkani-Hamed, S.~Dimopoulos and G.~R.~Dvali,
1346: % ``The hierarchy problem and new dimensions at a millimeter,''
1347: Phys.\ Lett.\ B {\bf 429} (1998) 263, hep--ph/9803315;
1348: I. Antonaidis, N. Arkani-Hamed, S. Dimopoulos and G.R. Dvali, Phys. Lett. B
1349: {\bf 436} (1998) 257, hep--ph/9804398.
1350:
1351: \bibitem{Randall:1999ee}
1352: L.~Randall and R.~Sundrum,
1353: % ``A large mass hierarchy from a small extra dimension,''
1354: Phys.\ Rev.\ Lett.\ {\bf 83} (1999) 3370, hep--ph/9905221.
1355:
1356: \bibitem{'tHooft:1987rb}
1357: G.~'t Hooft,
1358: % ``Graviton Dominance In Ultrahigh-Energy Scattering,''
1359: Phys.\ Lett.\ B {\bf 198} (1987) 61.
1360:
1361: \bibitem{banks}
1362: T. Banks and W. Fischler, hep--th/9906038.
1363:
1364: \bibitem{Giddings:2001bu}
1365: S.~B.~Giddings and S.~D.~Thomas,
1366: % ``High energy colliders as black hole factories: The end of short distance
1367: % physics,''
1368: Phys.\ Rev.\ D {\bf 65} (2002) 056010, hep--ph/0106219.
1369:
1370: \bibitem{Dimopoulos:2001hw}
1371: S.~Dimopoulos and G.~Landsberg,
1372: % ``Black holes at the LHC,''
1373: Phys.\ Rev.\ Lett.\ {\bf 87} (2001) 161602, hep--ph/0106295.
1374:
1375: \bibitem{Eardley:2002re}
1376: D.~M.~Eardley and S.~B.~Giddings,
1377: % ``Classical black hole production in high-energy collisions,''
1378: Phys.\ Rev.\ D {\bf 66} (2002) 044011, gr--qc/0201034.
1379:
1380: \bibitem{Solodukhin:2002ui}
1381: S.~N.~Solodukhin,
1382: % ``Classical and quantum cross-section for black hole production in particle
1383: % collisions,''
1384: Phys.\ Lett.\ B {\bf 533} (2002) 153, hep--ph/0201248.
1385:
1386: \bibitem{Hsu:2002bd}
1387: S.~D.~H.~Hsu,
1388: % ``Quantum production of black holes,''
1389: Phys.\ Lett.\ B {\bf 555} (2003) 92, hep--ph/0203154.
1390:
1391: \bibitem{Yoshino:2002tx}
1392: H.~Yoshino and Y.~Nambu,
1393: % ``Black hole formation in the grazing collision of high-energy particles,''
1394: Phys.\ Rev.\ D {\bf 67} (2003) 024009, gr--qc/0209003.
1395:
1396: \bibitem{Yoshino:2005hi}
1397: H.~Yoshino and V.~S.~Rychkov,
1398: % ``Improved analysis of black hole formation in high-energy particle
1399: % collisions,''
1400: Phys.\ Rev.\ D {\bf 71} (2005) 104028, hep--th/0503171.
1401:
1402: \bibitem{Myers:1986un}
1403: R.~C.~Myers and M.~J.~Perry,
1404: % ``Black Holes In Higher Dimensional Space-Times,''
1405: Annals Phys.\ {\bf 172} (1986) 304.
1406:
1407: \bibitem{Hawking:1974sw}
1408: S.~W.~Hawking,
1409: % ``Particle Creation By Black Holes,''
1410: Commun.\ Math.\ Phys.\ {\bf 43} (1975) 199,
1411: [Erratum-ibid.\ {\bf 46} (1976) 206].
1412:
1413: \bibitem{Emparan:2000rs}
1414: R.~Emparan, G.~T.~Horowitz and R.~C.~Myers,
1415: % ``Black holes radiate mainly on the brane,''
1416: Phys.\ Rev.\ Lett.\ {\bf 85} (2000) 499, hep--th/0003118.
1417:
1418: \bibitem{Anchordoqui:2002cp}
1419: L.~Anchordoqui and H.~Goldberg,
1420: %``Black hole chromosphere at the LHC,''
1421: Phys.\ Rev.\ D {\bf 67}, 064010 (2003), hep--ph/0209337.
1422:
1423: \bibitem{charybdis}
1424: C.M. Harris, P. Richardson and B.R. Webber, JHEP {\bf 0308} (2003) 033,
1425: hep--ph/0307305.
1426:
1427: \bibitem{catfish}
1428: M. Cavaglia, R. Godang, L. Cremaldi and D. Summers, hep--ph/0609001.
1429:
1430: \bibitem{Dimopoulos:2001qe}
1431: S.~Dimopoulos and R.~Emparan,
1432: % ``String balls at the LHC and beyond,''
1433: Phys.\ Lett.\ B {\bf 526} (2002) 393, hep-ph/0108060;
1434: K. Oda and N. Okada, Phys. Rev. {\bf D66} (2002) 095005, hep--ph/0111298.
1435:
1436: \bibitem{raffelt}
1437: S. Hannestad and G.G. Raffelt, Phys. Rev. D {\bf 67} (2003) 125008
1438: [Erratum-ibid. D {\bf 69} (2004) 029901], hep--ph/0304029.
1439: %``Supernova and Neutron Star Limits on Large Extra Dimensions Reexamined.''
1440:
1441: \bibitem{pdg}
1442: Particle Data Group, W.-M. Yao et al., Journal of Physics G {\bf 33} (2006) 1.
1443:
1444: \bibitem{anchor3}
1445: L.A. Anchordoqui, J.L. Feng, H. Goldberg and A.D. Shapere, Phys. Rev. D {\bf
1446: 68} (2003) 104025, hep--ph/0307228.
1447: %``Updated Limits on TeV-Scale Gravity from Absence of Neutrino Cosmic Ray
1448: % Showers Mediated by Black Holes''
1449:
1450: \bibitem{anchor2}
1451: L.A. Anchordoqui, J.L. Feng, H. Goldberg and A.D. Shapere, Phys. Lett. B {\bf
1452: 594} (2004) 363, hep--ph/0311365.
1453: %``Inelastic Black Hole Production and Large Extra Dimensions''
1454:
1455: \bibitem{hoss1}
1456: S. Hossenfelder, Phys. Lett. B {\bf 598} (2004) 92, hep--th/0404232.
1457: %``Suppressed Black Hole Production from Minimal Length''
1458:
1459: \bibitem{Ida1}
1460: D.~Ida, K.~Oda and S.~C.~Park,
1461: %``Rotating black holes at future colliders: Greybody factors for brane
1462: % fields,''
1463: Phys.\ Rev.\ D {\bf 67} (2003) 064025, [Erratum-ibid.\ D {\bf 69} (2004)
1464: 049901], hep--th/0212108; and
1465: % ``Rotating black holes at future colliders. II: Anisotropic scalar field
1466: % emission,''
1467: Phys.\ Rev.\ D {\bf 71} (2005) 124039, hep--th/0503052.
1468:
1469: \bibitem{kanti}
1470: P. Kanti and J. March-Russell, Phys. Rev. D {\bf 66} (2002) 024023,
1471: hep--ph/0203223, and Phys. Rev. D {\bf 67} (2003) 104019, hep--ph/0212199;
1472: C.M. Harris and P. Kanti, JHEP {\bf 0310} (2003) 014, hep--ph/0309054, and
1473: Phys. Lett. B {\bf 633} (2006) 106, hep--th/0503010;
1474: G. Duffy, C. Harris, P. Kanti and E. Winstanley, JHEP {\bf 0509} (2005) 049,
1475: hep--th/0507274;
1476: M. Casals, P. Kanti and E. Winstanley, JHEP {\bf 0602} (2006) 051,
1477: hep--th/0511163;
1478: M. Casals, S.R. Dolan, P. Kanti and E. Winstanley, hep--th/0608193.
1479:
1480: \bibitem{Ida:2006tf}
1481: D.~Ida, K.~Oda and S.~C.~Park, Phys. Rev. D {\bf 73} (2006) 124022,
1482: hep--th/0602188.
1483:
1484: \bibitem{enhance}
1485: V. Frolov and D. Stojkovic, Phys. Rev. D. {\bf 66} (2002) 084002,
1486: hep--th/0206046, and Phys. Rev. Lett. {\bf 89} (2002) 151302, hep--th/0208102;
1487: M. Cavaglia, Phys. Lett. B {\bf 569} (2003) 7, hep--ph/0305256;
1488: D. Stojkovic, Phys. Rev. Lett. {\bf 94} (2005) 011603, hep--ph/0409124.
1489:
1490: \bibitem{remnant}
1491: A. Bonanno and M. Reuter, Phys. Rev. D {\bf 62} (2000) 043008, hep--th/0002196,
1492: and Phys. Rev. D {\bf 73} (2006) 083005, hep--th/0602159;
1493: S. Hossenfelder et al., Phys. Lett. B {\bf 575} (2003) 85, hep--th/0305262.
1494:
1495: \bibitem{ua2}
1496: UA2 collab., J. Alitti et al., Phys. Lett. B {\bf 268} (1991) 145.
1497:
1498: \bibitem{cdf}
1499: CDF collab., F. Abe et al., Phys. Rev. D {\bf 56} (1997) 2532.
1500:
1501: \bibitem{order}
1502: G. Marchesini and B.R. Webber, Nucl. Phys. B {\bf 238} (1984) 1.
1503:
1504: \bibitem{Geiger:1998fq}
1505: K.~Geiger, R.~Longacre and D.~K.~Srivastava, nucl--th/9806102.
1506: %VNI version 4.1: Simulation of high-energy particle collisions in QCD
1507: %Space-time evolution of e+ e- $\to$ A + B collisions with parton cascades,
1508: %parton hadron conversion, final-state hadron cascades,''
1509:
1510: \bibitem{Bass:1999pv}
1511: S.~A.~Bass, M.~Hofmann, M.~Bleicher, L.~Bravina, E.~Zabrodin, H.~Stoecker and
1512: W.~Greiner,
1513: %``Analysis of reaction dynamics at ultrarelativistic energies in a combined
1514: %parton hadron transport approach'',
1515: Phys.\ Rev.\ C {\bf 60} (1999) 021901, nucl--th/9902055. The code is available
1516: from http://www.phy.duke.edu/~bass/vni.html .
1517:
1518: \bibitem{Sjostrand:2003wg}
1519: T.~Sj\"ostrand, L.~Lonnblad, S.~Mrenna and P.~Skands, hep--ph/0308153.
1520: %``PYTHIA 6.3: Physics and manual,''
1521:
1522: \bibitem{gnu}
1523: See http://www.gnu.org/software/gsl.
1524:
1525: \bibitem{Geiger:1991nj}
1526: K.~Geiger and B.~M\"uller,
1527: %``Dynamics of parton cascades in highly relativistic nuclear collisions,''
1528: Nucl.\ Phys.\ B {\bf 369} (1992) 600.
1529:
1530: \bibitem{Bengtsson:1986et}
1531: M.~Bengtsson and T.~Sj\"ostrand,
1532: %``A Comparative Study Of Coherent And Noncoherent Parton Shower Evolution,''
1533: Nucl.\ Phys.\ B {\bf 289} (187) 810.
1534:
1535: \bibitem{Webber:1986mc}
1536: B.~R.~Webber,
1537: %``Monte Carlo Simulation Of Hard Hadronic Processes,''
1538: Ann.\ Rev.\ Nucl.\ Part.\ Sci.\ {\bf 36} (1986) 253.
1539:
1540: \bibitem{schwinger}
1541: See e.g. J. Schwinger, ``Particles, Sources and Fields'', Vol. II
1542: (Addison--Wesley, Reading, MA, 1970/73).
1543:
1544: \end{thebibliography}
1545: \end{document}
1546: