1:
2: \documentstyle[aps,eqsecnum,epsf,epsfig]{revtex}
3:
4: \begin{document}
5:
6: \baselineskip=16pt
7: %\draft
8:
9: \title{Scenario for Ultrarelativistic Nuclear Collisions: \\
10: II.~ Geometry of quantum states at the earliest stage. }
11: \author{ A. Makhlin }
12: \address{Department of Physics and Astronomy, Wayne State University,
13: Detroit, MI 48202}
14: \date{July 26, 2000}
15: \maketitle
16: \begin{abstract}
17:
18: We suggest that the ultrarelativistic collisions of heavy ions provide the
19: simplest situation for the study of strong interactions which can be understood
20: from first principles and without any model assumptions about the microscopic
21: structure of the colliding nuclei. We argue that the boost-invariant geometry
22: of the collision, and the existence of hard partons in the final states, both
23: supported by the data, make a sufficient basis for the quantum theory of the
24: phenomenon. We conclude that the quantum nature of the entire process is
25: defined by its global geometry, which is enforced by a macroscopic finite size
26: of the colliding objects. In this paper, we study the qualitative aspects of
27: the theory and review its development in two subsequent papers. Our key
28: result is that the effective mass of the quark in the expanding system formed
29: in the collision of the two nuclei is gradually built up reaching its maximum by
30: the time the quark mode becomes sufficiently localized. The
31: chromo-magneto-static interaction of the color currents flowing in the
32: rapidity direction is the main mechanism which is responsible for the
33: generation of the effective mass of the soft quark mode and therefore, for the
34: physical scale at the earliest stage of the collision.
35:
36: \end{abstract}
37:
38: \section{Introduction}
39: \label{sec:S1}
40:
41:
42: In our previous papers \cite{QFK,QGD,tev}, we began a systematic theoretical
43: study of the scenario of ultrarelativistic collision of heavy ions. Our main
44: result obtained in Ref.~\cite{tev} (further quoted as paper [I]) was that the
45: dense system of quark and gluons which is commonly associated with the
46: quark-gluon plasma (QGP) can be formed only {\em in a single quantum
47: transition}. In this and two subsequent papers we continue to develop this
48: approach in greater detail. We come to a conclusion that ultrarelativistic
49: nuclear collisions is a unique physical phenomenon when the quantum dynamics of
50: the process is enforced by a macroscopic finite size of the colliding objects
51: rather than by a microscopic origin of their constituents.
52:
53: The entropy (the number of excited degrees of freedom) produced in collisions
54: of heavy ions is a natural measure of the strength of the colored fields
55: interaction. Indeed, before the collision, the quark and gluon fields are
56: assembled into two coherent wave packets (the nuclei) and therefore, the
57: initial entropy equals zero. The coherence is lost, and entropy is created due
58: to the interaction. A search for the QGP in heavy-ion collisions is, in the
59: first place, a search for evidence of entropy production. Though one may wish
60: to rely on the invariant formula $S={\rm Tr}\rho\ln\rho$, which expresses the
61: entropy $S$ via the density matrix $\rho$, at least one basis of states should
62: be found explicitly. It is imperative to design such a basis, and to
63: practically study the collective effects that take place at the earliest
64: ($<1fm$) stage of the collision.
65:
66: In any standard {\em exclusive} scattering process, no entropy can be produced
67: since the scattering process begins with a pure quantum state of two stable
68: colliding particles and the final state is also given as a pure state of
69: several particles in exactly known quantum states. The only way one can address
70: the quantum problem of entropy production is to consider {\em inclusive}
71: measurements. Since these measurements are not complete (i.e., are not
72: exclusive), they indeed form an ensemble with finite entropy.
73:
74: Quantum chromodynamics still cannot provide the theory of nuclear collisions
75: with detailed information about nuclear structure before the collision. We
76: face a formidable task to build a reliable theory of nuclear collision knowing
77: almost nothing about the initial state. We may rely safely only upon the fact
78: that the nuclei are stable bound states of the QCD and therefore, their
79: configuration is dominated by the stationary quark and gluon fields which are
80: genuine constituents of these quantum states. Fortunately, this, at first
81: glance very scarce, information appears to be sufficient for the understanding
82: of many intimate details of the collision process, if the problem is addressed
83: from first principles.
84:
85: We hope that the final state is defined more accurately and {\em believe} that
86: a single-particle distribution of quarks and gluons at some early moment after
87: the nuclei have intersected, describe it sufficiently. Thus, we may count upon
88: a reasonably well-defined quantum observable. The measurement of the
89: one-particle distribution is an inclusive measurement. The corresponding
90: operator should count the number of final-state particles defined as the
91: excitations above the perturbative vacuum. As long as we expect that this
92: counting makes sense on the event-by-event basis, the collision is indeed
93: producing the entropy. To develop the theory for this transition process we
94: have to cope with a binding feature that the ``final'' state has to be defined
95: at a finite time. This may look disturbing for readers well versed in
96: scattering theory, because the whole idea of {\em a scenario as a temporal
97: sequence of different stages} is alien to the standard S-matrix theory. The
98: general framework of an appropriate theory, named quantum field kinetics (QFK),
99: has been developed in our previous papers\cite{QFK,QGD,tev}. It is based on a
100: remarkable similarity; the measurement of one-particle distributions is as
101: inclusive as the measurement of the distribution of the final-state electron in
102: deeply inelastic {\em ep}-scattering (DIS). This conceptual similarity,
103: however, meets difficulties in its practical implementation.
104:
105: ~ (i) The inclusive DIS directly measures only the electromagnetic
106: fluctuations in the proton. The problem is posed according to the S-matrix
107: scattering theory improved by means of the renormalization group. The concept
108: of running coupling emerges precisely in this context. The operator product
109: expansion (OPE) allows one to hide all the unknown information about the
110: proton (commonly associated with large distances) into the local operators of
111: various dimensions. Introduced in this way, structure functions (given
112: explicitly in terms of their momenta) are applicable to DIS process, and only
113: to DIS.
114:
115: ~(ii) It is impossible to derive structure functions of {\em pp}-scattering
116: (not to say about {\em AA} ) using the OPE method, because in this case the
117: composite QCD operators become essentially non-local.\footnote{ Similar
118: situation takes place in the {\em ep}-process if a jet is chosen as the
119: inclusive observable. Then the dynamics of the process is sensitive to the QCD
120: content of the electron (see Sec.IV in paper [I]).}
121:
122: ~(iii) Historically, the escape was provided by the parton model ( the
123: factorization hypothesis ), which was successfully applied to various
124: processes that accompany {\em pp}-scattering (like Drell-Yan pairs
125: production), where the factorization scale can be kept under the data
126: control, since the number of particles in the final state is relatively
127: small. In {\em AA}-collisions, the control over the factorization scale is
128: practically impossible because of the enormously high multiplicity.
129: Furthermore, the phase space of the final state is densely populated and the
130: picture of an independent emission (unitary cut in the Feynman diagram)
131: employed for the derivation of DIS structure functions does not hold any more.
132:
133: As it has been already mentioned, in the {\em AA}-case we need a
134: developing in time scenario which cannot be accessed from the S-matrix
135: scattering theory, while the DIS structure functions are constructed within
136: S-matrix approach. The QFK method has been developed in order to resolve these
137: problems by addressing not only the nuclear collision as a transient process,
138: but the {\em ep}-DIS also. Our major hope was to derive QCD evolution
139: equations and to introduce the structure functions using the framework of an
140: independently initiated theory of nuclear collisions. The first step along
141: this guideline was an immediate success \cite{QFK,QGD}. It was demonstrated,
142: that the evolution equations indeed describe a transient process that ends as
143: an electromagnetic fluctuations inclusively probed by the electron.
144:
145:
146: To give a flavor of how the method works practically, let us start with a
147: qualitative description of the inclusive {\em e-p} DIS measurement (for now,
148: at the tree level without discussion of the effects of interference). In this
149: experiment, the only observable is the number of electrons with a given
150: momentum in the final state. Something {\em in the past} has to create the
151: electromagnetic field that deflects the electron. {\em Before} this field is
152: created, the electromagnetic current, which is the source of this field, has to
153: be formed. Since the momentum transfer in the process is very high, the current
154: has to be sufficiently localized. This localization requires, in its turn, that
155: the electric charges which carry this current must be dynamically decoupled
156: from the bulk of the proton {\em before} the scattering field is created (to
157: prevent a recoil to the other parts of the proton which could spread the
158: emission domain). Such a dynamical decoupling of a quark requires a proper
159: rearrangement of the gluonic component of the proton with the creation of
160: short-wave components of a gluon field. By causality, corresponding gluonic
161: fluctuation must happen {\em before} the current has decoupled, etc. Thus we
162: arrive at the picture of the sequential-dynamical fluctuations which create an
163: electromagnetic field probed by the electron. The lifetimes of these
164: fluctuations can be very short. Nevertheless, they all {\em coherently} add up
165: to form a stable proton, unless the interaction of measurement breaks the
166: proper balance of phases. This intervention freezes some instantaneous picture
167: of the fluctuations, but with wrong ``initial velocities'' which results in a
168: new wave function, and collapse of the old one. This qualitative picture has
169: been described many times and with many variations in the literature, starting
170: with the pioneering lecture by Gribov \cite{Gribov}, and including a recent
171: textbook \cite{BPQCD}; however, the sequential temporal ordering of the
172: fluctuations has never been a key issue. We derive this ordering as a
173: consequence of the Heisenberg equations of motion for the observables. The
174: practical scheme of calculation that emerges in this way appears to be {\em a
175: special form of quantum mechanics which describes an inclusive measurement as a
176: transient process}. Translated into mathematical language in momentum space,
177: this picture leads to the most general form of the evolution equations, which
178: may then be reduced (under different assumptions) to the known DGLAP, BFKL, or
179: GRV equations \cite{QGD}. The evolution equations were derived immediately in
180: the closed form of the integral equations avoiding a selective summation of
181: the perturbation series. The standard inclusive {\em e-p} DIS indeed delivers
182: information about quantum fluctuations which may {\em dynamically} develop in
183: the proton {\em before} it is destroyed by a hard electromagnetic probe. One
184: of the most amazing features that has been discovered in the framework of QFK
185: is that the QCD evolution equations are an {\em intrinsic property of the
186: inclusive measurement process}, and they are not limited by the factorization
187: condition.
188:
189: In paper [I], we studied the problem of loop corrections in the QFK evolution
190: equations. First, we found that they do not corrupt the causal picture of the
191: measurement described above, at the tree level. Second, they indeed provide a
192: {\em scale} to the entire process. This scale is connected with collective
193: interactions in the final state, which dynamically generate masses for the
194: final states of emission, thus regulating the abundant collinear divergences
195: of the null-plane dynamics. We required that the real parts of all radiative
196: corrections (phase shifts) must vanish along the direction of the
197: initial-state propagation of the colliding objects. Thus, we explicitly
198: accounted for the integrity of the nuclear wave function {\em before} the
199: collision. This special choice of the renormalization point, is {\em natural}
200: for ultrarelativistic nuclear collisions, since it allows one to treat
201: nuclei as finite-size quantum objects and incorporate their Lorentz
202: contraction as a classical boundary condition imposed on the space-time
203: evolution of quantum fields after the nuclear coherence is broken.
204:
205: The net yield of our previous study in paper [I] can be summarized as
206: follows: The interaction between the two ultrarelativistic nuclei switches
207: on almost instantaneously. This interaction explores all possible quantum
208: fluctuations which could have developed by the moment of the collision and
209: freezes (as the final states) only the fluctuations compatible with the
210: measured observable. These snapshots cannot have an arbitrary structure,
211: since the emerging configurations must be consistent with all the
212: interactions which are effective on the time-scale of the emission process.
213: In other words, the modes of the radiation field which are excited in the
214: course of the nuclear collision should be the collective excitations of the
215: dense quark-gluon system. This conclusion is the result of an intensive
216: search of the {\em scale} inherent in the process of a heavy-ion collision.
217: We proved that the scale is determined only by the physical properties of
218: the final state.
219:
220: Our previous study clearly indicates that a theory that describes both
221: phenomena (i.e. ep-DIS and AA-collisions) from a common point of view can be
222: built on two premises: causality, and the condition of emission. The latter
223: is also known as the principle of cluster decomposition, which must hold in
224: any reasonable field theory. What the ``resolved clusters'' are is a very
225: delicate question. These states should be defined with an explicit reference
226: as to how they are detected. Conventional detectors deal with hadrons and
227: allow one to hypothesize about jets. QGP turns out to be a kind of collective
228: detector for quarks and gluons. In DIS experiment, the new wave function is
229: measured {\em inclusively} which itself could be the source of the entropy
230: production if the final state had some properties of the collective system.
231: This collective system would then be a detector. It turns out that a dramatic
232: difference in population of the final states is the sole fact that makes DIS
233: and heavy ion collisions so different.
234:
235:
236: At this moment, the natural line of development of these ideas brought
237: us to the point when any further progress is impossible without
238: explicit knowledge of the {\em normal modes of the expanding dense
239: quark-gluon system}. There are several conceptual and technical
240: problems of different caliber where this knowledge is crucial.
241:
242: ~{\em 1.} Most of the entropy is expected to be produced during the initial
243: breakup of the nuclei coherence. Computing the entropy amounts to the digital
244: counting of the exited degrees of freedom. Therefore, the states themselves
245: must be precisely defined. From this point of view, the role of dynamical
246: masses of the normal modes is decisive. They provide an infrared boundary
247: for the space of final states thus making the possible number of the excited
248: states (the entropy) finite.
249:
250: ~{\em 2.} Only after the infrared boundary for the QCD states is found can
251: we hope to have a self-consistent perturbation theory. This was an original
252: idea which motivated the search of the QGP \cite{Shuryak}. A perturbative
253: description at the kinetic stage of the scenario cannot rely on massless QCD,
254: which has no intrinsic scale. It can be effective only if it is
255: based on the interaction of the partons-plasmons, i.e., quarks and gluons
256: with the effective masses. Built on these premises, the scenario for the
257: ultra- relativistic nuclear collision promises to be more perturbative than
258: the standard pQCD.
259:
260: ~{\em 2.} Standard perturbative calculations with massless gauge fields
261: always lead to collinear singularities that require a parameter of resolution
262: for their practical removal. When this singularity is due to the emission
263: into the final state, then this parameter is usually found as a property of
264: the detector. As long as we consider the QGP itself as a detector, no external
265: parameters of this kind can be in the theory. Collinear problems also appear
266: in loop corrections, even in their imaginary parts. Therefore, they are also
267: due to real processes. [ In physical gauges, the collinear singularities in the
268: loop corrections can also be connected with the spurious poles of the gluon
269: propagator, which is a consequence of an incomplete gauge fixing and imperfect
270: separation between the longitudinal fields and the fields of radiation.]
271:
272: As it has been discussed in paper [I], the collinear problems in perturbative
273: QCD show up only because the unphysical states are added to the list of the
274: possible final states of the radiation processes. These states can be
275: eliminated from the theory by accounting for the real interactions in the
276: final state which provide effective masses for all radiated fields. In the
277: null-plane dynamics, this appeared to be impossible, since any type of kinetics
278: that may lead to the formation of the effective mass is frozen on the light
279: cone. In order to have meaningful evolution equations for heavy ion collisions
280: we must account for the dynamical masses of the realistic final states in dense
281: expanding matter; the QCD evolution has to provide a kind of self-screening of
282: the collinear singularities. The way the effective mass was {\em estimated} in
283: paper [I] was crude, and it was our original goal to improve the calculations
284: using the full framework of {\em wedge dynamics}.
285:
286:
287:
288:
289:
290:
291: \section{Outline of ideas, calculations, main results, and conclusion}
292: \label{sec:SNo}
293:
294:
295: In this section, we review the work presented in this and two subsequent
296: papers \cite{geg,fse} (hereafter quoted as papers [III] and [IV]). Our approach
297: is strongly motivated by an idea, that collisions of ultrarelativistic heavy
298: ions is the cleanest laboratory where one can study the dynamics of strong
299: interactions. We consider an adequate choice of the interacting quantum states
300: at different stages of the scenario as the issue of first priority. The focus
301: of our previous study was on the QCD evolution equations in the environment of
302: the heavy ion collision. Now, we concentrate on the possible properties of the
303: state that emerges immediately {\em after} the coherence of the nuclei is
304: broken by the first hard interaction. As in paper [I], we view the dynamics of
305: the early stage as a single quantum process and concentrate on the study of
306: quantum fluctuations subjected to the condition of a simple inclusive
307: measurement (currently, on the inclusive one-particle distributions). We
308: endeavor to take full advantage of approaching the problem from first
309: principles.
310:
311: \subsection{Heuristic arguments.}
312: \label{subsec:So0}
313:
314: Collision of ultrarelativistic heavy ions is such a unique physical phenomenon,
315: that it is difficult to find its complete analog throughout everything that
316: has been studied in physics previously. However, we can point to several
317: examples which share some common distinctive patterns with the process under
318: investigation. We begin with these examples in order to help the reader
319: understand the ideas of our new synthesis.
320:
321: {\em 1}. Let an electron-positron pair be created by two photons. If the energy
322: of the collision is large, then the electron and positron are created in the
323: states of freely propagating particles and the cross-section of this process
324: accurately agrees with the tree-level perturbative calculation. However, if the
325: energy of the collision is near the threshold of the process, then the relative
326: velocity of the electron and positron is small, and they are likely to form
327: positronium. It would be incredibly difficult to compute this case using
328: scattering theory. Indeed, one has to account for the multiple emission of
329: soft photons which gradually builds up the Coulomb field between the electron
330: and positron and binds them into the positronium. However, the problem is
331: easily solved if we realize that the bound state {\em is} the final state for
332: the process. We can still use low-order perturbation theory to study the
333: transition between the two photons and the bound state of a pair
334: \cite{Sakharov}.
335:
336: {\em 2}. Let an excited atom be in a cavity with ideally conducting walls. The
337: system is characterized by three parameters: the size $L$ of the cavity, the
338: wave-length $\lambda\ll L$ of the emission, and the life-time $\Delta
339: t=1/\Gamma$ of the excited state. The questions are, in what case will the
340: emitted photon bounce between the cavity walls, and when will the emission
341: field be one of the normal cavity modes? The answer is very simple. If $c\Delta
342: t\ll L$, the photon will behave like a bouncing ball. When the line of
343: emission is very narrow, $c\Delta t\gg L$, the cavity mode will be excited. It
344: is perfectly clear that in the first case, the transition current that emits
345: the photon is localized in the atom. In the second case it is not. By the time
346: of emission, the currents in the conducting walls have to rearrange charges in
347: such a way that the emission field immediately satisfies the proper boundary
348: conditions. We thus have a collective transition in an extended system.
349:
350: {}From a practical point of view, these two different problems are united by
351: the method of obtaining their solutions. A part of the interaction (Coulomb
352: interaction in the first case, and the interaction of radiation with the cavity
353: walls in the second case) is attributed to the new ``bare'' Hamiltonian which
354: is diagonalized by the wave functions of the final state modes. The less
355: significant interactions can be accounted for by means of perturbation
356: theory. For us, the most important message is that it is possible to avoid a
357: difficult study of the transient process that physically creates these modes.
358:
359: {\em 3}. Let an experimental device consists of quantum detectors that
360: register photons emitted by a pulse source. Each pulse initiates an
361: ``event''. Let a sheet of glass is placed somewhere between the source and
362: detectors. If this glass were installed permanently in a fixed position, then
363: the method to account for its presence would be trivial. One must expand the
364: field of the initial light pulse over the system of modes (Fresnel triplets of
365: incident, reflected, and refracted waves) that satisfy the continuity
366: conditions on the glass boundaries. The quantum theory would then treat these
367: triplets as the photons, etc. When the position of the glass sheet is unknown,
368: e.g., it changes in the time periods between the pulses, then such a universal
369: decomposition becomes impossible. Nevertheless, in each particular event there
370: exists an important element of {\em classical boundary conditions}. Using
371: special tricks (e.g., by measurement of the times of arrival of the
372: precursors), one may determine the glass position and thus to learn how the
373: translational symmetry of free space was actually broken and what are the
374: photons of a particular event. Though the whole set up of this example is
375: artificial, it illustrates the major idea. The quantum theory of an individual
376: event can be fully recovered, even if macroscopic parameters of the theory
377: are not known until the event is completely recorded. Indeed, the prepared at
378: a large distance light pulse can be expanded over any of the systems of the
379: Fresnel triplets (corresponding to different positions of the glass sheet).
380: Only after analysis of the data can it be learned, which of these
381: decompositions is meaningful. One can fill the space between the detectors with
382: gas and account for the interaction between the light and gas (or even for a
383: non-linear interaction of photons in the gaseous medium) by perturbation
384: theory. Being the non-interacting waves in ``free space'', the Fresnel triplets
385: will serve as the zeroth-order approximation of a quantum theory. One may also
386: decide not to begin with the triple waves. Then the glass must be treated as
387: an active element. The same triplets will be recovered in the course of a real
388: transient process on the glass surface. The translational symmetry will be
389: broken dynamically.
390:
391: A very similar picture develops during the heavy ion collision. The normal
392: modes of the final state are formed in the course of real interactions. The
393: mechanism responsible for effective mass of the plasmons is illustrated by the
394: first two examples. The third example points us to an optimal choice of the
395: zeroth-order approximation. Exactly in the same way as the reflected and
396: refracted waves of the Fresnel triplet cannot physically appear before the
397: light front reaches the glass surface, nothing can happen with the nuclei
398: before they overlap geometrically. Only at this instance the interaction
399: determines the collision coordinates in $(tz)$-plane. The symmetry gets
400: uniquely broken, and the normal modes of the propagating colored fields after
401: the interaction exist only inside the future region of the interaction domain.
402: If the coupling is small then we may disregard later interactions. However,
403: the system of the final-state free fields will have a broken translational
404: symmetry, which will be remembered by the normal modes that obey certain
405: macroscopic boundary conditions.
406:
407: Referring to the above examples, one should keep in mind the source of the
408: major difference between the QED and QCD phenomena. The local gauge symmetry of
409: QED can be extended to a global gauge symmetry which generates the conserved
410: global quantum number (electrical charge) which can be sensed at a distance.
411: The proper field of an electric charge is the main obstacle for the definition
412: of its size. On the other hand, the radiation field of QED appears as a result
413: of the changes in the extended proper fields of accelerated charges, and one
414: can physically create such an object as a front of electromagnetic wave. In QCD,
415: the local gauge invariance of the color group does not correspond to any
416: conserved charge. Hence, we can easily determine the size of the colorless
417: nucleus, but we cannot create a front of color radiation in the gauge-invariant
418: vacuum. These two properties of QCD both work for us. They allow one to use
419: the Lorenz contraction to localize the initial moment of the collision and
420: thus, to impose the classical boundary conditions on the propagating color
421: fields at the later times. The existence of the collective propagating quark
422: and gluon modes at these times is the conjecture that has to be verified by the
423: study of heavy ion collisions.
424:
425: The finite size of the colliding nuclei, and a strong localization of the
426: initial interaction as its consequence, is a sufficient input for the theory
427: that describes the earliest stage of the collision. The formalism of quantum
428: field theory appears to be a powerful tool that allows one to derive many
429: properties of the quark-gluon system after the collision.
430:
431: \subsection{Phenomenological input.}
432: \label{subsec:So1}
433:
434: ~~(i)~ We consider the rapidity plateau seen event-by-event in nuclear
435: collisions at very large energy as a confirmed by the data indication that the
436: quantum transient process has no scale corresponding to the finite resolution
437: in the $t$- and $z$-directions. By a common wisdom, the absence of this scale
438: must cause the boost-invariant expansion.\footnote{This plateau in the
439: distribution of the final-state hadrons is clearly seen even in the inclusive
440: jet distribution in $ep$-DIS data, but only statistically.}
441:
442: ~~(ii)~ All existing data indicate that, regardless of the nature of the
443: colliding objects, a certain number of particles with large transverse momentum
444: are found in the final state. At high $p_t$, the cross section reasonably well
445: follows the Rutherford formula. We rely on the universality of Rutherford
446: scattering as an indication that there is no scale parameter of resolution in
447: the transverse $(xy)$-plane that characterize this process. We assume that in
448: nuclear collisions, these hard states created at the very early instance of
449: the collision can be described by the one-particle distribution measured on
450: event-by-event basis.
451:
452:
453:
454: \subsection{Ideas.}
455: \label{subsec:So2}
456:
457: ~~(i)~ The finite size of colliding nuclei plays a crucial role in our approach
458: since it allows for a realistic measurement of the Lorentz contraction thus
459: precisely fixing the time and the coordinate of the collision point. In the
460: laboratory frame, both nuclei are Lorentz contracted to a longitudinal size
461: $R_0/\gamma \sim 0.1 fm$, while the scale relevant for the hadron structure is
462: $ \sim 0.3 fm$. Therefore, in the center-of-mass frame, both nuclei are
463: passing through a ``pin-hole'', and the detailed information about the
464: microscopic nuclear structure is not essential. A precise measurement of the
465: {\em velocity}, i.e. the coordinates at two close time moments, is impossible
466: \cite{Landau}. Hence, a celebrated rapidity plateau in every single collision
467: of two ultrarelativistic ions is a direct consequence of this type of
468: measurement. We accept the fact of the rapidity plateau as a classical boundary
469: condition for the quantum sector of the theory.
470:
471: ~~(ii)~ There is no doubt that the entire collision process must develop
472: inside the future light cone of the collision domain. Only there can the
473: dynamics of the propagating color fields become a physical reality. In other
474: words, the resolution of colored degrees of freedom is a consequence of the
475: precise measurement of the coordinate by means of strong interactions.
476:
477: ~~(iii)~ The true scale of the entire quantum process coincides with its
478: infrared boundary, which is build dynamically in the course of this process.
479: Namely, the hard partons, which are produced as localized and countable
480: particles at the earliest time of the process, define masses for the soft field
481: states formed at the later times thus bringing the transient process to its
482: saturation.
483:
484:
485:
486: \subsection{Strategy and theoretical foundations.}
487: \label{subsec:So3}
488:
489: Addressing the problem of normal modes of the expanding quark-gluon system,
490: we
491: proceed in two major steps. First, we study the classical and quantum
492: properties of the normal modes subjected to the boundary condition of a
493: localized interaction that follows from the relativistic causality (being
494: the free fields in all other respects). Then, we use these modes as a basis
495: for the perturbation theory and compute the effective mass of the quark
496: propagating through the background distribution of hard partons.
497:
498: ~~(i)~ We begin in Sec.~\ref{subsec:SBe1} with the qualitative study of free
499: fields, fully incorporating the properties of the geometric background of the
500: expanding matter. Taking the simplest plane-wave of the scalar field as an
501: example, and studying the probability to detect this wave on the space-like
502: hypersurface of constant proper time $\tau$, we conclude that it is capable of
503: passing through the center $t=z=0$. The only price paid for this feature is the
504: full delocalization of the state along the hyperplanes $\tau^2=t^2-z^2=0$. The
505: state is completely delocalized at $\tau p_t\ll 1$, and it is sharply localized
506: in the rapidity direction at $\tau p_t\gg 1$. In this way, we approach an idea of
507: {\em wedge dynamics}, which employs the proper time $\tau$ as the natural
508: direction of the evolution. In Sec.~\ref{subsec:SBe2}, we consider a wave
509: packet and demonstrate that the process of localization at finite time $\tau$
510: is physical; it is accompanied by the gradual re-distribution of the charge
511: density and the current of this charge. From this observation, we may
512: anticipate a special role of the magneto-static interactions at the earliest
513: times, when the process of the charge density rearrangement is extremely
514: rapid. Further calculations of paper [IV] give even more evidence that the
515: quantum process of delocalization predicted by wedge dynamics is a material
516: process.
517:
518: ~~(ii)~ As a first step towards practical calculations, the fields are
519: described classically and quantized in the scope of wedge dynamics. In
520: Sec.~\ref{sec:SN4} of this paper, we accomplish this procedure for the fermion
521: fields. In Sec.~\ref{sec:SN5}, we derive the expressions for various quantum
522: correlators, which are used for the perturbative calculation of the fermion
523: self-energy in paper [IV]. An important observation made at this point is that
524: the material parts of the field correlators immediately have the form of Wigner
525: distributions. This is a unique property of wedge dynamics which relies on the
526: highly localized states as its one-particle basis.
527:
528: ~~(iii)~ The third one, technically the most complicated paper [III] of this
529: cycle, is dedicated to the vector gauge field in wedge dynamics. Several
530: conceptual and technical problems are addressed there. First of all, the
531: states of the free radiation field are studied classically. Also at the
532: classical level, we compute the retarded Green function of the vector gauge
533: field and explicitly separate the longitudinal (i.e., governed by Gauss law)
534: field and the field of radiation. It is found that if the physical charge
535: density $\rho=\tau j_\tau$ vanishes at the starting point $\tau=0$, then
536: Gauss law of the wedge dynamics, being in fact a constraint, becomes an
537: immediate consequence of the equations of motion. Therefore, Gauss law can
538: be explicitly used to eliminate the unphysical degrees of freedom of the gauge
539: field, and the gauge $A^\tau=0$ can be fixed completely. Using this result, we
540: were able to quantize the gluon field according to the standard procedure of
541: canonical quantization.
542:
543: The requirement $\rho(\tau=0)=0$ would not be physical in QED, where the
544: long-range proper fields of electric charges would limit the possible
545: localization of the first interaction, and the applicability of the wedge
546: dynamics. On the other hand, in the wedge dynamics of colorless objects built
547: from the colored fields, which are ``stretched'' at $\tau\to 0$ along a very
548: wide rapidity interval, this can be a true initial condition. The later
549: creation of the localized color charges can indeed be initiated by the color
550: currents in the color-neutral (at $\tau=0$) system.
551:
552: ~~(iv)~ A distinctive property of the longitudinal gauge fields in wedge
553: dynamics is that they do not look like usual static fields. The Hamiltonian
554: time $\tau$ does not coincide with a usual time of some particular inertial
555: Lorentz frame. This is a proper time for all observers that move with all
556: possible rapidities starting from the point $t=z=0$. The system, which is static
557: with respect to this time experiences a permanent expansion, and its Gauss
558: fields have magnetic components. As a consequence, the longitudinal part of the
559: gauge field propagator acquires a contact term,
560: $$D^{[contact]}_{\eta\eta}=-{\tau_1^2-\tau_2^2 \over 2} \delta (\eta) \delta
561: (\vec{r_t})~.$$ The component $D_{\eta\eta}$ establishes a connection between
562: the $A_\eta$ component of the potential and the $j_\eta$ component of the
563: current. In its turn, $A_\eta$ is responsible for the $\eta$-component
564: $E_\eta=\partial_\tau A_\eta$ of the electric field and the $x$- and
565: $y$-components, $B_x=\partial_y A_\eta$, $B_y=-\partial_x A_\eta$ of the
566: magnetic field. The electrical field in the longitudinal $\eta$-direction is
567: not capable of producing the scattering with transverse momentum transfer.
568: However, this transfer can be provided by the magnetic forces; the two currents
569: $j_\eta$ can interact via the magnetic field ${\vec B}_t=(B_x,B_y)$. The origin
570: of these currents is intrinsically connected with the geometry of states in the
571: wedge form of dynamics. The existence of these currents indicates that the
572: delocalization of the nuclear wave packet is more than a formal decomposition
573: in terms of fancy modes. This is a physical phenomenon which plays an important
574: role in the formation of the IR scale of the entire process.
575:
576: \subsection{Calculation of the effective mass}
577: \label{subsec:So4}
578:
579: The first calculation that incorporates both the ideas and technical part of
580: the wedge dynamics is attempted in paper [IV]. We compute the effective
581: ``transverse mass'' $\mu(\tau,p_t)$ of the soft (i.e., $\tau p_t<1$) quark
582: mode propagating through the expanding background of hard (i.e., $\tau k_t>1$)
583: partons.
584:
585: ~~(i)~In order to find the normal modes of the quark field in the expanding
586: quark-gluon system, we solve the Dirac equation with radiative
587: corrections, which can be derived as a projection of the Schwinger-Dyson
588: equation for the retarded quark propagator onto the one-particle initial
589: state. This equation can be converted into a dispersion equation that includes
590: the retarded self-energy and connects the effective transverse mass
591: $\mu(\tau,p_t)$ of the soft mode with its transverse momentum $p_t$. This
592: equation depends on the current proper time $\tau$ as a parameter,
593: $$\mu(\tau,p_t) = p_t + \int_{0}^{\tau}d\tau_2 \sqrt{\tau\tau_2}
594: e^{i\mu(\tau,p_t)(\tau-\tau_2)} \Sigma_{ret}(\tau,\tau_2;p_t)~,$$
595: and we assumed that $d \ln\mu/d\ln\tau\ll 1$, in deriving it. The solution
596: with this property is indeed found.
597:
598: ~~(ii)~ The material part of the self-energy can be divided into several
599: parts corresponding to different processes of the forward quark scattering on
600: the hard partons of the expanding surroundings. First, the quark may scatter on
601: a real (transverse) gluon. The second process is quark-quark scattering,
602: which can be conveniently divided into two subprocesses. In one of them, the
603: interaction is mediated by the radiation part of the gluon field, in the other,
604: the mediator is the longitudinal part. The latter can be split further into the
605: contact and non-local parts. Our strategy was to find the leading terms of the
606: self-energy which are singular at $\tau-\tau_2=0$, and thus can significantly
607: contribute to the effective quark mass within a short time. Indeed, since we are
608: looking for the time-dependent $\mu(\tau,p_t)$, this mass has to be formed
609: during a sufficiently short time interval. Accordingly, we have chosen the
610: dimensionless parameter $\xi=(\tau-\tau_2)/\sqrt{\tau\tau_2}$ as a small
611: parameter.
612:
613: ~~(iii)~ The distribution of hard quarks and gluons that may provide an
614: effective mass to a soft quark mode with transverse momentum $p_t$ at
615: the time $\tau\leq 1/p_t$ are taken in agreement with the qualitative
616: arguments of sections \ref{subsec:So1} and \ref{subsec:So2},
617: $$ n_f(q_t,\theta)\approx {{\cal N}_f\over \pi R_\bot^2}
618: {\theta (q_t-p_\ast) \over q_t^2},~~~
619: n_g(k_t,\alpha)\approx {{\cal N}_g\over \pi R_\bot^2}
620: {\theta (k_t-p_\ast) \over k_t^2}~.$$
621: They are not related to any dynamical scale and the normalization factors
622: ${\cal N}_g$ and ${\cal N}_f$ are the only (apart from the coupling
623: $\alpha_s$) parameters of the theory. The impact cross section $\pi
624: R_\bot^2$ and the full width $2Y$ of the rapidity plateau are defined by
625: the geometry of a particular collision and the c.m.s. energy, respectively.
626: These are irrelevant for the local screening parameters we are interested in.
627:
628:
629: ~~(iv)~ Analysis of the terms that include radiation fields clearly reveals
630: two trends. On the one hand, the integration over the transverse momenta of
631: hard quarks and gluons is capable of creating a singularity when the rapidities
632: corresponding to the two lines in the loop coincide. On the other hand, the
633: interval of rapidities where the collinear geometry is possible is extremely
634: narrow due to the light-cone boundaries (causality) of the forward scattering
635: process. The second factor always wins, and the contribution of the collinear
636: domain is always small. We also found that the observed intermediate collinear
637: enhancement of the forward scattering amplitude is, as a matter of fact,
638: fictitious. It is entirely formed by the integration over the infinitely large
639: transverse momenta which are physically absent in the distribution of the hard
640: partons. (Formally, the infinite transverse momentum is needed to provide a
641: precise tuning of two states with given rapidities to each other.) These
642: collinear singularities are integrable, and they do not lead to a disaster of
643: collinear divergence.
644:
645: Our way to pick out the leading contributions from the space-time domains,
646: where the phases of the interacting fields are stationary, is a generalization
647: of the known method of isolating the leading terms using the pinch-poles in the
648: plane of complex energy. The wedge dynamics does not allow for a standard
649: momentum representation, since its geometric background is not homogeneous in
650: $t$- and $z$-directions. Nevertheless, the patches of phase space, where the
651: phases of certain field fragments are stationary and effectively overlap, do
652: now the same job as the pinch-poles, and yield the same answers when the
653: homogeneity required for the momentum representation is restored. This way to
654: tackle the problem is genuinely more general, because it addresses the
655: space-time picture of the interacting fields. The role of pinch-poles is taken
656: over by the geometrical overlap of the field patterns with the same rapidity.
657: This observation can serve as a footing for the future development of an
658: effective technique for perturbative calculations in wedge dynamics.
659:
660: ~~(v)~ The effect of the non-local components of the longitudinal part of the
661: gluon propagator that mediates the quark-quark scattering, was shown to be
662: small also. This interaction cannot lead to the collinear enhancement. However,
663: its yield could be not very small, because the interaction has long range. It
664: occurs, that the non-local electro- and magneto-static interactions of charges
665: just almost compensate each other.
666:
667: ~~(vi)~ The only term in the quark self-energy which is singular at small time
668: differences is due to the above mentioned contact term in the $D^{\eta\eta}$
669: component of the gluon propagator. This is the leading contribution to the
670: dispersion equation provided by the magneto-static interaction of the
671: longitudinal currents. Studied in the first approximation, the solution of the
672: dispersion equation indicates that in compliance with the original idea, the
673: effective mass $\mu(\tau,p_t)$ gradually increases with time reaching its
674: maximum when $\tau p_t\approx 1$. This is the major practical result of this
675: study. Evolution of the fields at the later times must be approached with
676: another set of normal modes that, from the very beginning, account for the
677: screening effects developed at the previous stage.
678:
679:
680: \subsection{ Conclusion and perspectives. }
681: \label{subsec:So5}
682:
683: In a series of papers reviewed in this section, we demonstrated that the field
684: theory is indeed able to describe a scenario. By scenario we mean a continuous
685: smoothly developing temporal sequence of one stage into another. These stages
686: are different only in the respect that each of them is characterized by its
687: individual {\em optimal set} of normal modes. In contrast with paper [I], we
688: do not focus on the stage of QCD evolution, since we have no clear image of the
689: objects that initiate destruction of nuclear coherence. Instead, we try to
690: understand, what can be the immediate products of this destruction. This is an
691: example of the continuity that stands behind the idea of the scenario. The next
692: stage will be the kinetics of the partons-plasmons, and we anticipate that it
693: will impose new restrictions, which will improve our current results. [Quantum
694: mechanics works remarkably in both directions: any information about the
695: properties of the final state imposes limitations on the possible line of the
696: evolution (including the initial data) at the earlier times exactly in the same
697: way as the known initial data imposes restrictions on the possible final
698: states.] By the same token, we must look for a connection between the objects
699: resolved in the first interaction of two nuclei and the known properties of
700: hadrons and the QCD vacuum. Unfortunately, this appealing opportunity is still
701: distant.
702:
703: First principles appeared to be a powerful tool for achieving our goals.
704: With minimal theoretical input and with the reference to the simplest data,
705: they allow one to build a self-consistent picture of the initial stage of
706: the collision. Colliding the nuclei, we probably create the theoretically
707: simplest situation for understanding the nature of the process. In the
708: course of this study, we relied only on the fact of boost invariance of the
709: process and an assumption that the field states with large transverse
710: momentum, even at very early times, may be associated with the localized
711: particles and thus can be described by the distribution with respect to their
712: rapidity and transverse momentum. Our strategy of looking for the leading
713: contributions and all our approximations in calculating the material part
714: of the quark self-energy are based on this assumption. If it appears
715: incorrect, then it is most likely that the quark-gluon matter created in
716: the collision of two nuclei never, and in no approximation, can be
717: considered as a system of nearly free and weakly interacting field states.
718:
719: Our decision to begin the exploration of potentialities of the wedge dynamics
720: with the computation of quark self-energy is motivated only by technical
721: reasons. The gluon propagator of wedge dynamics is a very complicated function,
722: and we preferred to start with the computation of the fermion loop which has
723: only one gluon correlator in it. We hope that the discovery of, in the course
724: of our study, an enormous simplifications (with respect to what we had to start
725: with) will allow us to address the more important problem of the gluon
726: self-energy in a reasonably economic way.
727:
728:
729: \section{Field states in the proper-time dynamics}
730: \label{sec:S2}
731:
732: The dynamical masses of normal modes at finite density are found from the
733: dispersion equation that includes the corresponding self-energy, i.e., the
734: amplitude of the forward scattering of the mode on the particles that populate
735: the phase space. In paper [I], we found that it is impossible to adequately
736: describe this basic process of forward scattering in the null-plane
737: dynamics. The problem arises due to the singular behavior of the field
738: pattern which is defined as the static field with respect to the Hamiltonian
739: time $x^+$. This singular behavior alone shows that the choice of the
740: dynamics
741: and the proper definition of the field states is a highly nontrivial and
742: important issue. Besides, if we tried to describe quantum fluctuations in the
743: second nucleus in the same fashion, then it would require a second
744: Hamiltonian time $x^-$, which is not acceptable. Thus, if we wish to view
745: the collision of two nuclei as a unique quantum process, then it is
746: imperative to find a way to describe quarks and gluons of both nuclei, as
747: well as the products of their interaction, using {\it the same Hamiltonian
748: dynamics}. An appropriate choice for the gluons is always difficult because
749: the gauge is a global object (as are the Hamiltonian dynamics) and both
750: nuclei should be described using the same gauge condition.
751:
752: Quantum field theory has a strict definition of {\it dynamics}. This
753: notion was introduced by Dirac \cite{Dirac} at the end of the 1940's in
754: connection with his attempt to build a quantum theory of the gravitational
755: field. Every (Hamiltonian) dynamics includes its specific definition of
756: the quantum mechanical observables on the (arbitrary) space-like surfaces,
757: as well as the means to describe the evolution of the observables from
758: the ``earlier'' space-like surface to the ``later'' one.
759:
760: The primary choice of the degrees of freedom is effective if, even without
761: any interaction, the dynamics of the normal modes adequately reflects the
762: main physical features of the phenomenon. The intuitive physical arguments
763: clearly indicate that the normal modes of the fields participating in the
764: collision of two nuclei should be compatible with their Lorentz
765: contraction. Unlike the incoming plane waves of the standard scattering
766: theory, the nuclei have a well-defined shape and the space-time domain of
767: their intersection is also well-defined. Hence, the geometric properties of
768: the expected normal modes follows, in fact, from the uncertainty principle.
769: Indeed, we may view the first touch of the nuclei as the first of the two
770: measurements which are necessary to determine the velocity. Since a
771: precise measurement of the nuclei coordinate at an exactly determined
772: moment appears to be an inelastic process that completely destroys the
773: nuclei, the spectrum of the longitudinal velocities of the
774: final-state components must become extremely wide \cite{Landau}. These
775: components may also be different by their transverse momenta. With respect
776: to the measurement of the longitudinal velocity, the latter plays a role of
777: an ``adjoint mass''. The velocity of a heavier object can be measured with
778: a larger accuracy. Therefore, the separation of the ``heavy'' final state
779: fragments by their longitudinal velocities requires less time and can be
780: verified earlier than for the ``light'' ones.\footnote{The boost-invariance
781: with the fixed center means the absence of a corresponding scale and {\em
782: vice-versa}. Any relativistic equations, regardless of their physical
783: content, will yield a self-similar solution. For example, the relativistic
784: hydrodynamic equations lead to a known Bjorken solution with the rapidity
785: plateau. In its turn, the Bjorken solution can be obtained as a limit of
786: the Landau solution with an infinite Lorentz contraction of the colliding
787: objects. We favor the arguments that are closer to quantum mechanics and
788: allow for the further connection with the properties of the quantum states.}
789:
790: The same conclusion can be reached formally: Of the ten symmetries of the
791: Poincar\'{e} group, only rotation around the collision $z$-axis, boost along
792: it, and the translations in the transverse $x-$ and $y$-directions
793: survive. The idea of the collision of two plane sheets immediately leads us
794: to the {\em wedge form}; the states of quark and gluon fields before and
795: after the collision must be confined within the past and future light cones
796: (wedges) with the $xy$-collision plane as the edge. Therefore, it is
797: profitable to choose, in advance, the set of normal modes which have the
798: symmetry of the localized initial interaction and carry quantum numbers
799: adequate to this symmetry. These quantum numbers are the transverse components
800: of momentum and the rapidity of the particle (which replaces the component
801: $p^z$ of its momentum). In this {\it ad hoc} approach, all the spectral
802: components of the nuclear wave functions ought to collapse in the
803: two-dimensional plane of the interaction, even if all the confining
804: interactions of the quarks and gluons in the hadrons and the coherence of the
805: hadronic wave functions are neglected.
806:
807: In the wedge form of dynamics, the states of free quark and gluon fields are
808: defined (normalized) on the space-like hyper-surfaces of the constant proper
809: time $\tau$, $\tau^2=t^2-z^2$. The main idea of this approach is to study the
810: dynamical evolution of the interacting fields along the Hamiltonian time
811: $\tau$. The gauge of the gluon field is fixed by the condition $A^\tau =0$.
812: This simple idea solves several problems. On the one hand, it becomes
813: possible to treat the two different light-front dynamics which describe each
814: nucleus of the initial state separately, as two limits of this single
815: dynamics. On the other hand, after the collision, this gauge simulates a
816: local (in rapidity) temporal-axial gauge. This feature provides a smooth
817: transition to the boost-invariant regime of the created matter expansion (as a
818: first approximation). Particularly, addressing the problem of screening, we
819: will be able to compute the plasmon mass in a uniform fashion, considering each
820: rapidity interval separately.
821:
822: As it was explained in the first two sections, the feature of the states to
823: collapse at the interaction vertex is crucial for understanding the dynamics
824: of a high-energy nuclear collision. A simple optical prototype of the wedge
825: dynamics is the {\em camera obscura} (a dark chamber with the pin-hole in the
826: wall). Amongst the many possible {\em a priori} ways to decompose the incoming
827: light, the camera selects only one. Only the spherical harmonics centered at
828: the pin-hole can penetrate inside the camera. The spherical waves reveal
829: their angular dependence at some distance from the center and build up the
830: image on the opposite wall. Here, we suggest to view the collision of two
831: nuclei as a kind of diffraction of the initial wave functions through the
832: ``pin-hole'' $t=0,~z=0$ in $tz$-plane.
833:
834: Using the proper time $\tau$ as the natural direction of the evolution of the
835: nuclear matter after the collisions has far reaching consequences. The
836: surfaces of constant $\tau$ are curved, and the oriented objects like spinors
837: and vectors have to be defined with due respect to this curvature. We have to
838: incorporate the tetrad formalism in order to differentiate them covariantly.
839: The properties of local invariance are modified also, since the different
840: directions in the tangent plane become not equivalent. The physical content of
841: the theory also undergoes an important change. The system of observers that
842: are used to {\em normalize} the quantum states of wedge dynamics is different
843: from the observers of any particular inertial Lorentz frame.
844:
845: \subsection{One-particle wave functions in wedge dynamics}
846: \label{subsec:SBe1}
847:
848: In order to study the main kinematic properties of the states of the wedge
849: dynamics, it is enough to consider the one-particle wave functions of the
850: scalar field. Let us take the wave function $\psi_{\theta,p_\perp}(x)$ of the
851: simplest form,
852: \begin{eqnarray}
853: \psi_{\theta,p_\perp}(x)= {1 \over 4\pi^{3/2}}
854: e^{-ip^0t+ip^zz+i{\vec p}_{\perp}{\vec r}_{\perp}}\equiv
855: \left\{ \begin{array}{l} 4^{-1}\pi^{-3/2}
856: e^{-im_{\perp}\tau\cosh(\eta-\theta)}
857: e^{i{\vec p}_{\perp}{\vec r}_{\perp}},~~~~\tau^2>0, \\
858: 4^{-1}\pi^{-3/2} e^{-im_{\perp}\tau\sinh(\eta-\theta)}
859: e^{i{\vec p}_{\perp}{\vec r}_{\perp}},~~~~ \tau^2<0~. \end{array}\right.
860: \label{eq:S1.1}\end{eqnarray}
861: where $p^0=m_{\perp}\cosh\theta$, $p^z=m_{\perp}\sinh\theta$ ($\theta$ being
862: the rapidity of the particle), and, as usual,
863: $m_{\perp}^{2}=p_{\perp}^{2}+m^2$. The above form implies that $\tau$ is
864: positive in the future of the wedge vertex and negative in its past. Even
865: though this wave function is obviously a plane wave which occupies the whole
866: space, it carries the quantum number $\theta$ (rapidity of the particle)
867: instead of the momentum $p_z$. A peculiar property of this wave function is
868: that it may be normalized in two different ways, either on the hypersurface
869: where $t=const$,
870: \begin{eqnarray}
871: \int_{t=const}\psi^{\ast}_{\theta',p'_\perp}(x)
872: ~i{\stackrel{\leftrightarrow}{\partial \over \partial t}}~
873: \psi_{\theta,p_\perp}(x)~dz d^2{\vec r}_{\bot}=
874: \delta (\theta -\theta')\delta ({\vec p}_{\bot}-{\vec p'}_{\bot})~~,
875: \label{eq:S1.2}\end{eqnarray}
876: or, equivalently, on the hypersurfaces $\tau=const$ in the
877: future- and the past--light wedges of the collision plane, where $\tau^2>0$,
878: \begin{eqnarray}
879: \int_{\tau=const}\psi^{\ast}_{\theta',p'_\perp}(x)
880: ~i{\stackrel{\leftrightarrow}{\partial \over \partial \tau}}~
881: \psi_{\theta,p_\perp}(x)~\tau d\eta d^2{\vec r}_{\bot}=
882: \delta (\theta -\theta')\delta ({\vec p}_{\bot}-{\vec p'}_{\bot})~.
883: \label{eq:S1.3}\end{eqnarray}
884: Being almost identical mathematically, these two equations are very
885: different physically. Eq.~(\ref{eq:S1.2}) implies that the state is
886: detected by a particular Lorentz observer equipped by a grid of detectors
887: that cover the whole space, while Eq.~(\ref{eq:S1.3}) normalizes the
888: measurements performed by an array of the detectors moving with all
889: possible velocities. At any particular time of the Lorentz observer, this
890: array even does not cover the whole space.
891:
892: The norm of a particle's wave function always corresponds to the conservation
893: of its charge or the probability to find it. Since the norm given by
894: Eq.~(\ref{eq:S1.3}) does not depend on $\tau$, the particle with a given
895: rapidity $\theta$ (or velocity $v=\tanh\theta=p^z/p^0$), which is
896: ``prepared'' on the surface $\tau=const$ in the past light wedge, cannot
897: flow through the light-like wedge boundaries; the particle is
898: predetermined to penetrate in the future light wedge through its vertex. The
899: dynamics of the penetration process can be understood in the following way.
900:
901: At large $~m_{\perp}|\tau|$, the phase of the wave function
902: $\psi_{\theta,p_\perp}$ is stationary in a very narrow interval around
903: $~\eta=\theta~$ (outside this interval, the function oscillates with
904: exponentially increasing frequency); the wave function describes a particle
905: with rapidity $~\theta$ moving along the classical trajectory. However, for
906: $m_{\perp}|\tau|\ll 1$, the phase is almost constant along the surface
907: $\tau=const$. The smaller $\tau~$ is, the more uniformly the domain of
908: stationary phase is stretched out along the light cone. A single particle
909: with the wave function $\psi_{\theta,p_\perp}$ begins its life as the wave
910: with the given rapidity $\theta$ at large negative $\tau$. Later, it becomes
911: spread out over the boundary of the past light wedge as $\tau\to -0$. Still
912: being spread out, it appears on the boundary of the future light wedge.
913: Eventually, it again becomes a wave with rapidity $\theta$ at large positive
914: $\tau$. The size and location of the interval where the phase of the wave
915: function is stationary plays a central role in all subsequent discussions,
916: since it is equivalent to the localization of a particle. Indeed, the
917: overlapping of the domains of stationary phases in space and time provides the
918: most effective interaction of the fields.
919:
920: The size $\Delta\eta$ of the $\eta$-interval around the particle rapidity
921: $\theta$, where the wave function is stationary, is easily evaluated.
922: Extracting the trivial factor $e^{-im_{\perp}\tau}$ which defines the
923: evolution of the wave function in the $\tau$-direction, we obtain an
924: estimate from the exponential of Eq.~(\ref{eq:S1.1}),
925: \begin{eqnarray}
926: 2~m_{\perp}\tau\sinh^2(\Delta\eta/2)\sim 1~~.
927: \label{eq:S1.4}\end{eqnarray}
928: The two limiting cases are as follows,
929: \begin{eqnarray}
930: \Delta\eta \sim \sqrt{2\over m_{\perp}\tau},
931: ~~{\rm when}~~ m_{\perp}\tau \gg 1~,~~~{\rm and}~~~
932: \Delta\eta \sim 2 \ln {2\over m_{\perp}\tau},
933: ~~{\rm when}~~ m_{\perp}\tau\ll 1~~.
934: \label{eq:S1.5}\end{eqnarray}
935: In the first case, one may boost this interval into the laboratory reference
936: frame and see that the interval of the stationary phase is Lorentz contracted
937: (according to the rapidity $\theta$) in $z$-direction. This estimate
938: confirms what follows from physical intuition; for a heavier quantum object,
939: the velocity can be measured with the higher accuracy. The states of the wedge
940: dynamics appear to be almost ideally suited for the analysis of the processes
941: that are localized at different times $\tau$ and intervals of rapidity $\eta$,
942: and are characterized by a different transverse momentum transfer. With
943: respect to any particular process, these states are easily divided into
944: slowly varying fields and localized particles. In this way, one may introduce
945: the distribution of particles and study their effect on the dynamics of the
946: fields. As a result, we can calculate the plasmon mass as a local (at some
947: scale) effect which agrees with our understanding of its physical origin.
948:
949: \subsection{Dynamical properties of states in wedge dynamics}
950: \label{subsec:SBe2}
951:
952: The property of the wave function to concentrate with the time near the
953: classical world line of a particle with the given velocity has important
954: implications. This is a gradual process and it must be accompanied by the
955: re-distribution of the charge density and the current of this charge. To see
956: how this happens explicitly, let us consider a particle in a superposition
957: state $|\theta_0\rangle$ of a normalized wave packet,
958: \begin{eqnarray}
959: |\theta_0\rangle =\int_{-\infty}^{\infty}d\theta f(\theta-\theta_0)
960: \bbox{a^\dag}_\theta |0\rangle~,~~~~~
961: \langle \theta_0|\theta_0\rangle = \int_{-\infty}^{\infty}d\theta
962: f^\ast(\theta-\theta_0) f(\theta-\theta_0) = 1~,
963: \label{eq:S1.6}\end{eqnarray}
964: where $\bbox{a^\dag}_\theta$ is the Fock creation operator for the
965: one-particle state with the rapidity $\theta$.\footnote{We do not describe
966: here the procedure of the scalar field quantization in wedge dynamics. It is
967: exactly the same as quantization of the fermion field in the next section.}
968: The explicit form of the weight function in Eq.~(\ref{eq:S1.6}) may vary.
969: Solely for convenience, we take the weight function $f(\theta-\theta_0)$ of
970: the form,
971: \begin{eqnarray}
972: f(\theta-\theta_0) = [K_{0}(2\xi)]^{-1/2} e^{-\xi\cosh(\theta-\theta_0)}
973: \approx (4\xi /\pi)^{1/4} e^\xi~e^{-\xi\cosh(\theta-\theta_0)}~,
974: \label{eq:S1.7}\end{eqnarray}
975: which provides a sharp localization of the wave packet. In the second of
976: these equations, we used an asymptotic approximation of the Kelvin function
977: $K_0(2\xi)$, which is reasonably accurate starting from $\xi\geq 1/2$.
978:
979: The operator of the four-current density for the complex scalar field
980: $\Psi$ is well known to be
981: \begin{eqnarray}
982: J_\mu (x) = \Psi^\dag (x)~
983: i{\stackrel{\leftrightarrow}\partial_\mu}~\Psi (x)~,
984: \label{eq:S1.8}\end{eqnarray}
985: and to obey the covariant conservation law,
986: \begin{eqnarray}
987: \nabla_\mu J^\mu (x) =
988: (-g)^{-1/2}\partial_\mu [(-g)^{1/2} g^{\mu\nu}J_\nu (x)]
989: = \tau^{-1}[\partial_\tau(\tau J_\tau) + \partial_\eta(\tau^{-1}J_\eta)]=0~.
990: \label{eq:S1.9}\end{eqnarray}
991: (Here, for simplicity, we consider the two-dimensional case and employ the
992: metric $g^{\tau\tau}=1$, $g^{\eta\eta}=-\tau^{-2}$.) The physical
993: components of the current (which are defined in such a way that the integral
994: form of the conservation law is not altered by the curvilinear metric) are
995: ${\cal J}_\tau=\tau J_\tau $ and ${\cal J}_\eta=\tau^{-1}J_\eta$. Using
996: Eqs.~(\ref{eq:S1.6}) and (\ref{eq:S1.8}), we can compute their expectation
997: values of these components in the state $|\theta_0\rangle$.
998: \begin{eqnarray}
999: \langle \theta_0|{\cal J}_\tau|\theta_0\rangle = \tau k_t
1000: \int_{-\infty}^{\infty}{d\theta_1d\theta_2\over 4\pi}
1001: f^\ast(\theta_1-\theta_0)f(\theta_2-\theta_0)
1002: [\cosh(\eta -\theta_1)+\cosh(\eta -\theta_2)]
1003: e^{i\tau k_t[\cosh(\eta -\theta_1)-\cosh(\eta -\theta_2)]}~,
1004: \label{eq:S1.10}\end{eqnarray}
1005: \begin{eqnarray}
1006: \langle \theta_0|{\cal J}_\eta|\theta_0\rangle = {k_t\over\tau}
1007: \int_{-\infty}^{\infty}{d\theta_1d\theta_2\over 4\pi}
1008: f^\ast(\theta_1-\theta_0)f(\theta_2-\theta_0)
1009: [\sinh(\eta -\theta_1)+\sinh(\eta -\theta_2)]
1010: e^{i\tau k_t[\cosh(\eta -\theta_1)-\cosh(\eta -\theta_2)]}~.
1011: \label{eq:S1.11}\end{eqnarray}
1012: The integrals over $\theta_1$ and $\theta_2$ can be estimated by means of
1013: the saddle point approximation even for the relatively small values of $\xi$,
1014: e.g. $\xi\sim 1$, because the hyperbolic functions in the exponents vary
1015: sufficiently rapidly near the stationary points. These calculations yield the
1016: following result,
1017: \begin{eqnarray}
1018: \langle \theta_0|{\cal J}_\tau|\theta_0\rangle =
1019: {2\tau k_t\over (\pi\xi)^{1/2}}~{\cosh(\eta-\theta_0)\over
1020: 1+{\tau^2k_t^2\over\xi^2}\cosh[2(\eta-\theta_0)]}~
1021: {e^{-{\tau^2k_t^2\over\xi}\sinh^2(\eta-\theta_0)}\over
1022: \sqrt{1+{\tau^2k_t^2\over\xi^2}\sinh^2(\eta-\theta_0)}} ~,
1023: \label{eq:S1.12}\end{eqnarray}
1024: \begin{eqnarray}
1025: \langle \theta_0|{\cal J}_\eta|\theta_0\rangle =
1026: {2k_t\over (\pi\xi)^{1/2}}~{\sinh(\eta-\theta_0)\over
1027: 1+{\tau^2k_t^2\over\xi^2}\cosh[2(\eta-\theta_0)]}~
1028: {e^{-{\tau^2k_t^2\over\xi}\sinh^2(\eta-\theta_0)}\over
1029: \sqrt{1+{\tau^2k_t^2\over\xi^2}\sinh^2(\eta-\theta_0)}} ~.
1030: \label{eq:S1.13}\end{eqnarray}
1031: These dependences are plotted in Fig.~\ref{fig:fig1} up to a common
1032: scale factor.
1033: \bigskip
1034:
1035: \begin{figure}[htb]
1036: \begin{center}
1037: \mbox{
1038: \psfig{file=./fig1b.ps,height=2.2in,bb=100 490 390 710}
1039: \hspace{1.cm}
1040: \psfig{file=./fig1a.ps,height=2.2in,bb=100 490 390 710}
1041: }
1042: \end{center}
1043: \caption{Charge density in the wave packet (left) and current density
1044: (right) evolution.}
1045: \label{fig:fig1}
1046: \end{figure}
1047:
1048: {}From the left figure, it is easy to see that the evolution of the charge
1049: density ${\cal J}_\tau$ starts from the lowest magnitude and the widest spread
1050: at small $\tau$. Then it gradually becomes narrow and builds up a significant
1051: amplitude near the classical trajectory with the rapidity $\theta_0$. This
1052: process is accompanied by the charge flow ${\cal J}_\eta$ (right figure) which
1053: has its maximal values at small $\tau$, and then gradually vanishes at later
1054: times, when the process of building the classical particle comes to its
1055: saturation. The extra factor $\tau^{-1}~$ in $~\langle \theta_0|{\cal J}_\eta
1056: |\theta_0\rangle$, which tends to boost current at small $\tau$, is of
1057: geometric origin. Thus, the behavior of the local observables in the wave
1058: packet confirms the simple arguments of Sec.~\ref{subsec:SBe1} based on the
1059: analysis of the domain where the wave function is stationary. One can guess
1060: about the possible nature of interactions at the earliest times by making an
1061: observation that the $\eta$-component of the current must produce the $x$- and
1062: $y$-components of the magnetic field. These fields are the strongest at the
1063: earliest times when $\tau k_t\ll 1$. The magnetic fields of the transition
1064: currents provide scattering with the most effective transfer of the transverse
1065: momentum. Indeed, at time $\tau_2$ the quark with the transverse momentum
1066: $p_t$, $\tau_2 p_t\ll 1$ interacts with the gluon field and acquires a large
1067: transverse momentum $k_t$, $\tau_2 k_t\gg 1$. This transition is characterized
1068: by a drastic narrowing of the charge spread in the rapidity direction, and must
1069: be accompanied by a strong $\eta$-component of the transition current. A
1070: similar transition in the opposite direction happens at the time $\tau_1$, when
1071: the gluon field interacts with another quark that has large initial transverse
1072: momentum $k_t$, and recovers the soft state with $\tau_2 p_t\ll 1$ in the
1073: course of this interaction. This second transition current readily interacts
1074: with the magnetic component of the gluon field. These speculations will be
1075: justified in paper [IV] of this cycle by the explicit calculation of quark
1076: self-energy in the expanding system.
1077:
1078: Three remarks are in order: First, the field of a free on-mass-shell particle
1079: can be only static, and it is common to think that, in the rest frame of the
1080: particle, it is a purely electric field. In the wedge dynamics, the particle is
1081: formed during a finite time and this formation process unavoidably generates
1082: the magnetic component of the {\em longitudinal} (i.e., governed by the Gauss
1083: law) field. This will be obtained more rigorously in the next paper when the
1084: full propagator of the gauge field in wedge dynamics will be found.
1085: Furthermore, in wedge dynamics, the source must be called as static if it
1086: expands in such a way that ${\cal J}_\tau=\tau J_\tau = const(\tau)$ , and its
1087: field strength also has a magnetic component. Second, the local color current
1088: density may be large even when the system is color-neutral (begins its
1089: evolution from the colorless state), as it seems to be the case in heavy-ion
1090: collisions. It will be also shown in paper [III], that in order to fix the
1091: gauge $~A^\tau=0~$ completely, one must require that the physical charge
1092: density $~{\cal J}^\tau=0~$ at $~\tau=0~$. Finally, in wedge dynamics we meet a
1093: unique structure of phase space, where two variables, the velocity of a
1094: particle and its rapidity coordinate, just duplicate each other at sufficiently
1095: late proper time $\tau$. The quantum mechanical uncertainty principle does not
1096: prohibit one to address them on equal footing, because the one-particle wave
1097: packets of the wedge dynamics, evolving in time, become more and more narrow in
1098: rapidity direction.
1099:
1100:
1101:
1102: \section{States of Fermions}
1103: \label{sec:SN4}
1104:
1105: The hypersurfaces of constant Hamiltonian time $\tau$ of wedge dynamics are
1106: curved. Therefore, all oriented objects like vectors or spinors are
1107: essentially defined only in the tangent space and therefore, their
1108: covariant derivatives should be calculated in the framework of the
1109: so-called tetrad
1110: formalism \cite{Fock,Witten}.\footnote{In what follows, we use the Greek
1111: indices for four-dimensional vectors and tensors in the curvilinear
1112: coordinates, and the Latin indices from $a$ to $d$ for the vectors in flat
1113: Minkowsky coordinates. We use Latin indices from $r$ to $w$ for the
1114: transverse $x$- and $y$-components ($r,...,w=1,2$), and the arrows over the
1115: letters to denote the two-dimensional vectors, {\em e.g.}, ${\vec
1116: k}=(k_x,k_y)$, $|{\vec k}|=k_{t}$. The Latin indices from $i$ to $n$
1117: ($i,...,n=1,2,3$) will be used for the three-dimensional internal
1118: coordinates $u^i=(x,y,\eta)$ on the hyper-surface $\tau=const$.}
1119: The covariant derivative of the tetrad vector includes two connections
1120: (gauge fields). One of them, the Levi-Civita connection
1121: $$\Gamma^{\lambda}_{~\mu\nu}={1\over 2} {\rm g}^{\lambda\rho}
1122: \bigg[{\partial {\rm g}_{\rho\mu} \over \partial x^\nu }
1123: +{\partial {\rm g}_{\rho\nu} \over \partial x^\mu }
1124: -{\partial {\rm g}_{\mu\nu} \over \partial x^\rho } \bigg]~,$$
1125: is the gauge field, which provides covariance with respect to the general
1126: transformation of coordinates. The second gauge field, the spin
1127: connection $\omega_{\mu}^{~ab}(x)$, provides covariance with respect to
1128: the local Lorentz rotation. Let $x^\mu =(\tau,x,y,\eta)$ be the
1129: contravariant components of the curvilinear coordinates and
1130: $x^a=(t,x,y,z)\equiv (x^0,x^1,x^2,x^3)$ are those of the flat Minkowsky
1131: space. Then the tetrad vectors $e^{a}_{~\mu}$ can be taken as follows,
1132: \begin{eqnarray}
1133: e^{0}_{~\mu}=(1,0,0,0),~~e^{1}_{~\mu}=(0,1,0,0),~~
1134: e^{2}_{~\mu}=(0,0,1,0),~~e^{3}_{~\mu}=(0,0,0,\tau)~.
1135: \label{eq:E3.1}\end{eqnarray}
1136: They correctly reproduce the curvilinear metric ${\rm g}_{\mu\nu}$ and
1137: the flat Minkowsky metric $g_{ab}$, {\em i.e.},
1138: \begin{eqnarray}
1139: {\rm g}_{\mu\nu}=g_{ab}e^{a}_{~\mu}e^{b}_{~\nu}
1140: ={\rm diag}[1,-1,-1,-\tau^2]~,~~~
1141: g^{ab}= {\rm g}^{\mu\nu} e^{a}_{~\mu}e^{b}_{~\nu}={\rm diag}[1,-1,-1,-1] ~.
1142: \label{eq:E3.2}\end{eqnarray}
1143: The spin connection can be found from the condition that the covariant
1144: derivatives of the tetrad vectors are equal to zero,
1145: \begin{eqnarray}
1146: \nabla_\mu e^{a}_{~\nu}=\partial_\mu e^{a}_{~\nu}
1147: +\omega^{~a}_{\mu~b}e^{b}_{~\nu} -
1148: \Gamma^{\lambda}_{~\mu\nu}e^{a}_{~\lambda} =0~.
1149: \label{eq:E3.3}
1150: \end{eqnarray}
1151: The covariant derivative of the spinor field includes only the spin
1152: connection,
1153: \begin{eqnarray}
1154: \nabla_\mu\psi(x)= \big[\partial_\mu +
1155: {1\over 4}\omega_{\mu}^{~ab}(x)\Sigma_{ab}\big]\psi(x)~,
1156: \label{eq:E3.4}\end{eqnarray}
1157: where $\Sigma^{ab}={1\over 2}[\gamma^a\gamma^b-\gamma^b\gamma^a]$ is
1158: an obvious generator of the Lorentz rotations and $\gamma^a$ are the
1159: Dirac matrices of Minkowsky space. Introducing the Dirac matrices
1160: in curvilinear coordinates, $\gamma^{\mu}(x)=e^{\mu}_{~a}(x)\gamma^a$,
1161: one obtains the Dirac equation in curvilinear coordinates,
1162: \begin{eqnarray}
1163: [\gamma^\mu(x)(i\nabla_\mu +gA_\mu(x))-m]\psi(x)=0 ~,
1164: \label{eq:E3.5}\end{eqnarray}
1165: where $A^\mu(x)$ is the gauge field associated with the local group of
1166: the internal
1167: symmetry. The conjugated spinor is defined as
1168: ${~\overline \psi}=\psi^{\dag}\gamma^0$, and obeys the equation,
1169: \begin{eqnarray}
1170: (-i\nabla_\mu +gA_\mu(x)){\overline\psi}(x)\gamma^\mu(x)-
1171: m{\overline\psi}(x)=0 ~~.
1172: \label{eq:E3.6}\end{eqnarray}
1173:
1174: These two Dirac equations correspond to the action,
1175: \begin{eqnarray}
1176: {\cal A}=\int d^4 x \sqrt{-g}{\cal L}(x)=
1177: \int d^4 x \sqrt{-g} \{ {i\over 2}[{\overline\psi}\gamma^\mu(x)
1178: \nabla_\mu\psi- (\nabla_\mu {\overline\psi})\gamma^\mu(x)\psi]
1179: + g{\overline\psi}\gamma^\mu(x) A_\mu\psi -
1180: m{\overline\psi}\psi \}~~,
1181: \label{eq:E3.7}
1182: \end{eqnarray}
1183: from which one easily obtains the locally conserved $U(1)$-current,
1184: \begin{eqnarray}
1185: J^\mu(x)= {\overline\psi}(x)\gamma^\mu(x)\psi(x)~,
1186: ~~~~ (-{\rm g})^{-1/2}\partial_\mu [(-{\rm g})^{1/2}
1187: {\rm g}^{\mu\nu}(x)J_\nu (x)]=0~.
1188: \label{eq:E3.8}\end{eqnarray}
1189: The Dirac equations (\ref{eq:E3.5}) and (\ref{eq:E3.6}) can be
1190: alternatively obtained as the equations of the
1191: Hamiltonian dynamics along the proper time $\tau$.
1192: The canonical momenta conjugated to the fields $\psi$ and ${\overline\psi}$
1193: are
1194: \begin{eqnarray}
1195: \pi_{\psi}(x)={\delta(\sqrt{-{\rm g}}{\cal L})\over\delta \dot{\psi}(x)}
1196: = {i\tau\over 2} {\overline\psi}(x)\gamma^0
1197: \;\; {\rm and} \;\;\;
1198: \pi_{{\overline\psi}}(x)=
1199: {\delta(\sqrt{-{\rm g}}{\cal L})\over\delta \dot{{\overline\psi}(x)} }
1200: = -{i\tau\over 2} \gamma^0\psi(x)~,
1201: \label{eq:E3.9}\end{eqnarray}
1202: respectively.
1203: The Hamiltonian of the Dirac field in the wedge dynamics has the following
1204: form,
1205: \begin{eqnarray}
1206: H=\int d\eta d^2 {\vec r}\;\sqrt{-{\rm g}}
1207: \{ -{i\over 2}[{\overline\psi}\gamma^i(x)
1208: \nabla_i\psi- (\nabla_i {\overline\psi})\gamma^i (x)\psi]
1209: - g{\overline\psi}\gamma^\mu(x) A_\mu\psi+
1210: m{\overline\psi}\psi \}~,
1211: \label{eq:E3.10}\end{eqnarray}
1212: and the wave equations are just the Hamiltonian equations of
1213: motion for the momenta.
1214:
1215: The non-vanishing components of the connections are
1216: $\Gamma^{\bf\cdot}_{\eta\tau\eta} = \Gamma^{\bf\cdot}_{\eta\eta\tau}=
1217: -\Gamma^{\bf\cdot}_{\tau\eta\eta}=-\tau$~ and ~$\omega_{\eta}^{~30}
1218: =-\omega_{\eta}^{~03}=1$. Moreover, we have $\gamma^\tau(x)=\gamma^0$
1219: and $\gamma^\eta(x)=\tau^{-1}\gamma^3 $. The explicit form of the
1220: Dirac equation in our case is as follows,
1221: \begin{eqnarray}
1222: [i \not\!\nabla -m]\psi(x)= [i\gamma^0(\partial_\tau +{1\over 2\tau})
1223: +i\gamma^3{1\over \tau}\partial_\eta +i \gamma^r \partial_r -m]\psi(x)=0~~.
1224: \label{eq:E3.11}\end{eqnarray}
1225: The one-particle solutions of this equation must be normalized according to
1226: the
1227: charge conservation law (\ref{eq:E3.8}). We choose the scalar
1228: product of the following form,
1229: \begin{eqnarray}
1230: (\psi_1,\psi_2)=
1231: \int \tau d\eta ~d^2 {\vec r}~
1232: {\overline \psi}_1 (\tau,\eta,{\vec r})\gamma^\tau
1233: \psi_2 (\tau,\eta,{\vec r})~~.
1234: \label{eq:E3.12}\end{eqnarray}
1235: With this definition of the scalar product, the Dirac equation is
1236: self-adjoint.
1237: The solutions to this equation will be looked for in the form
1238: $\psi(x)= [i \not\!\nabla +m]\chi(x)$, with the function $\chi(x)$ that
1239: obeys the ``squared'' Dirac equation,
1240: \begin{eqnarray}
1241: [i \not\!\nabla -m][-i \not\!\nabla +m]\chi(x)=
1242: \bigg[\partial_{\tau}^{2} +{1\over \tau}\partial_\tau
1243: -{1\over \tau^2}\partial_{\eta}^{2} - \partial_{r}^{2} +m^2-
1244: \gamma^0\gamma^3 {1\over \tau^2}\partial_{\eta} \bigg]\chi(x)=0~~.
1245: \label{eq:E3.13}\end{eqnarray}
1246: The spinor part $\beta_\sigma$ of the function $\chi(x)$ can be chosen
1247: as an eigenfunction of the operator $\gamma^0\gamma^3 $, namely,
1248: $~\gamma^0\gamma^3\beta_{\sigma} = \beta_{\sigma}$, and $\sigma=1,2~$.
1249: Therefore, the solution of the original Dirac equation
1250: (\ref{eq:E3.5}) can be written down
1251: as $~\psi^{\pm}_{\sigma} = w_\sigma \chi^{\pm}(x)$,~
1252: with the bi-spinor operators $w_\sigma=[i \not\!\nabla +m]\beta_\sigma $
1253: that act on the positive- and negative-frequency solutions $\chi^{\pm}(x)$
1254: of the scalar equation
1255: \begin{eqnarray}
1256: \bigg[\partial_{\tau}^{2} +{1\over \tau}\partial_\tau
1257: -{1\over \tau^2}(\partial_{\eta}+{1\over 2})^{2} - \partial_{r}^{2}
1258: +m^2\bigg]\chi^{\pm}(x)=0~~.
1259: \label{eq:E3.14}\end{eqnarray}
1260: In what follows, we shall employ only the partial waves with quantum numbers
1261: of transverse momentum $\vec{p_t}$ and rapidity $\theta$ of massless quarks.
1262: In this case, the (already normalized) scalar functions $\chi^{\pm}(x)$ are
1263: \begin{eqnarray}
1264: \chi^{\pm}_{\theta,{\vec p_t}}(x)= (2\pi)^{-3/2} (2p_t)^{-1/2}
1265: ~e^{(\theta-\eta)/2}
1266: e^{\mp i p_t\tau\cosh(\theta-\eta)}~e^{\pm i{\vec p}{\vec r}}~.
1267: \label{eq:E3.15}\end{eqnarray}
1268: Consequently, the one-particle solutions are
1269: \begin{eqnarray}
1270: \Omega^{\pm}_{\sigma,\theta,{\vec p_t}}(x)= (2\pi)^{-3/2} (2p_t)^{-1/2}
1271: i \not\!\nabla~\beta_\sigma e^{(\theta-\eta)/2}
1272: e^{\mp i p_t\tau\cosh(\theta-\eta)}~e^{\pm i{\vec p}{\vec r}}~,
1273: \label{eq:E3.16}\end{eqnarray}
1274: where the spinors $\beta_\sigma$ can be chosen in different ways. However,
1275: regardless of a particular choice of the spinors $\beta_\sigma$, the
1276: polarization sum is always
1277: \begin{eqnarray}
1278: \sum_\sigma \beta_\sigma \otimes \beta_\sigma= {1+\gamma^0\gamma^3\over 2}~.
1279: \label{eq:E3.17}\end{eqnarray}
1280: The waves $\Omega^{\pm}_{\sigma,\theta,{\vec p_t}}(x)$ are orthonormalized
1281: according to
1282: \begin{eqnarray}
1283: (\Omega^{(\pm)}_{\sigma,\theta,{\vec p}},
1284: \Omega^{(\pm)}_{\sigma',\theta',{\vec p}'})= \delta_{\sigma \sigma'}
1285: \delta({\vec p}-{\vec p}')\delta(\theta-\theta')~,~~~
1286: (\Omega^{(\pm)}_{\sigma,\theta,{\vec p}},
1287: \Omega^{(\mp)}_{\sigma',\theta',{\vec p}'})= 0~~.
1288: \label{eq:E3.18}\end{eqnarray}
1289: These partial waves form a complete set (cf. Eq.~(\ref{eq:E3.32a})) and
1290: therefore, can be used to decompose the fermion field,
1291: \begin{eqnarray}
1292: \Psi (x)= \sum_\sigma \int d^2 {\vec p_t} d \theta
1293: [ \bbox{a}_{\sigma,\theta,{\vec p_t}}\Omega^{(+)}_{\sigma,\theta,{\vec
1294: p_t}}(x)
1295: +\bbox{b}^{\dag}_{\sigma,\theta,{\vec p_t}}
1296: \Omega^{(-)}_{\sigma,\theta,{\vec p_t}}(x)],\nonumber \\
1297: \Psi^{\dag} (x)=\sum_\sigma\int d^2 {\vec p_t} d \theta
1298: [\bbox{a}^{\dag}_{\sigma,\theta,{\vec p_t}}
1299: \overline{\Omega}^{(+)}_{\sigma,\theta, {\vec p_t}}(x)
1300: +\bbox{b}_{\sigma,\theta,{\vec p_t}}
1301: \overline{\Omega}^{(-)}_{\sigma,\theta,{\vec p_t}}(x)]~.
1302: \label{eq:E3.19}\end{eqnarray}
1303: The canonical quantization procedure, which identifies the
1304: coefficients $\bbox{a}$ and $\bbox{b}$ with the Fock operators, is standard,
1305: and it leads to the anti-commutation relations,
1306: \begin{eqnarray}
1307: [\bbox{a}_{\sigma,\theta,{\vec p_t}},
1308: \bbox{a}^{\dag}_{\sigma',\theta',{\vec p_t}'}]_+ =
1309: [\bbox{b}_{\sigma,\theta,{\vec p_t}},
1310: \bbox{b}^{\dag}_{\sigma',\theta',{\vec p_t}'}]_+ =
1311: \delta_{\sigma \sigma'}
1312: \delta({\vec p}-{\vec p}')\delta(\theta-\theta')~,
1313: \label{eq:E3.19a}\end{eqnarray}
1314: all other anticommutators being zero.
1315: Non-trivial issues of the canonical quantization in wedge dynamics show up
1316: only in the gluon sector. They will be discussed in paper [III] of this cycle.
1317:
1318:
1319: \section{Fermion correlators}
1320: \label{sec:SN5}
1321:
1322: The field-theory calculations are based on various field correlators. A full
1323: set of these correlators is employed by the Keldysh-Schwinger formalism
1324: \cite{Keld} which will be used below in the form given in
1325: Refs.~\cite{QFK,QGD,tev}\footnote{The indices of the contour ordering, as
1326: well as the labels of linear combinations of variously ordered correlators, are
1327: placed in
1328: square brackets, e.g., $G_{[AB]}$, $G_{[ret]}= G_{[00]}-G_{[01]}$, etc.}. This
1329: set consists of two Wightman functions, where the field operators are taken in
1330: fixed order,
1331: \begin{eqnarray}
1332: G_{[10]}(x_1,x_2)= -i\langle \Psi(x_1)\overline{\Psi}(x_2)\rangle ,~~~~
1333: G_{[01]}(x_1,x_2)= i\langle \overline{\Psi}(x_2)\Psi(x_1)\rangle~,
1334: \label{eq:E3.20}\end{eqnarray}
1335: and two differently ordered Green functions
1336: \begin{eqnarray}
1337: G_{[00]}(x_1,x_2)= -i\langle T[\Psi(x_1)\overline{\Psi}(x_2)]\rangle ,~~~~
1338: G_{[11]}(x_1,x_2)= -i\langle T^\dag [\Psi(x_1)\overline{\Psi}(x_2)]\rangle~.
1339: \label{eq:E3.21}\end{eqnarray}
1340: Here, $\langle\cdot\cdot\cdot\rangle$ denotes the average over an ensemble of
1341: the excited modes. The vacuum state (of each particular mode) is a part of
1342: this
1343: ensemble. In order to find the explicit form of these correlators, we shall
1344: employ the modes corresponding to the states with a given transverse
1345: momentum $\vec{p_t}$ and an unusual rapidity quantum number $\theta$.
1346: Further on, it will be profitable to use the field correlators in the mixed
1347: representation when they are Fourier-transformed only by their transverse
1348: coordinates $\vec{r_t}$, while the dependence on the proper time $\tau$
1349: and the
1350: rapidity coordinate $\eta$ is retained explicitly. Below, we derive the
1351: corresponding expressions. The details of the derivation are important, since
1352: they help to clarify physical issues related to the the localization of
1353: quanta in the wedge dynamics, and are beneficial for the future analysis of
1354: collinear singularities in paper [IV]. As for the ``vacuum part'' of the
1355: correlators,
1356: we obtain the more or less known expressions and put them into the form
1357: which is convenient for future calculations.
1358:
1359: For the practical calculations, we shall need not the functions $G_{[AB]}$
1360: of Eqs. (\ref{eq:E3.20}) and (\ref{eq:E3.21}), but their linear combinations,
1361: the fermion anti-commutator $G_{[0]}$ and the density of states $G_{[1]}$,
1362: \begin{eqnarray}
1363: G_{[0]}(x_1,x_2)= G_{[10]}(x_1,x_2)- G_{[01]}(x_1,x_2),~~~~
1364: G_{[1]}(x_1,x_2)= G_{[10]}(x_1,x_2)+ G_{[01]}(x_1,x_2)~,
1365: \label{eq:E3.20a}\end{eqnarray}
1366: and the retarded and advanced Green functions,
1367: \begin{eqnarray}
1368: G_{[ret]}(x_1,x_2)= G_{[00]}(x_1,x_2)- G_{[01]}(x_1,x_2) =
1369: \theta (\tau_1-\tau_2)~G_{[0]}(x_1,x_2)~, \nonumber\\
1370: G_{[adv]}(x_1,x_2)= G_{[00]}(x_1,x_2)- G_{[10]}(x_1,x_2)=
1371: -\theta (\tau_2-\tau_1)~G_{[0]}(x_1,x_2)~.
1372: \label{eq:E3.21a}\end{eqnarray}
1373: Nevertheless, we have to start with the computation of the simplest
1374: correlators, the Wightman functions. Using Eqs.~(\ref{eq:E3.20}) and
1375: (\ref{eq:E3.19}), we obtain
1376: \begin{eqnarray}
1377: G_{[10]}(x_1,x_2;p_t)=-i\int {d\theta \over 8\pi}
1378: \big[ \gamma^+p_te^{-\theta}~e^{(\eta_1+\eta_2)/2}+
1379: \gamma^-p_te^{+\theta}~e^{-(\eta_1+\eta_2)/2}
1380: +p_r\gamma^r\gamma^0(\gamma^+e^{-{\eta_1-\eta_2\over 2}}
1381: +\gamma^-e^{+{\eta_1-\eta_2\over 2}})\big]\nonumber\\
1382: \times \bigg[[1-n^{+}(\theta,p_t)]
1383: e^{-ip_t[\tau_1\cosh(\theta-\eta_1)- \tau_2\cosh(\theta-\eta_2)]}
1384: +n^{-}(\theta,p_t)
1385: e^{+ip_t[\tau_1\cosh(\theta-\eta_1)- \tau_2\cosh(\theta-\eta_2)]}\bigg]~.
1386: \label{eq:E3.22}
1387: \end{eqnarray}
1388: It is useful to keep in mind a simple connection between this
1389: expression and the standard one. Since, $\gamma^\pm=\gamma^0\pm\gamma^3$,
1390: and $p_t e^{\pm\theta}=p^\pm\equiv p^0\pm p^3$, the first line in this
1391: formula can be rewritten as $~\Lambda(-\eta_1)\not\! p ~\Lambda(\eta_2)$,
1392: where
1393: \begin{eqnarray}
1394: \Lambda (\eta)=\cosh(\eta/2) + \gamma^0\gamma^3\sinh(\eta/2)=
1395: {\rm diag}[e^{\eta/2},e^{-\eta/2},e^{-\eta/2},e^{\eta/2}]~,
1396: \label{eq:E3.23}\end{eqnarray}
1397: is the matrix of Lorentz rotation with the boost $\eta$. Furthermore, the
1398: quantum number $\theta$ can be formally changed into $p_z$. Incorporating the
1399: mass-shell delta-function $\delta(p^2)$ and returning to Cartesian
1400: coordinates, we obtain,
1401: \begin{eqnarray}
1402: G_{[10]}(x_1,x_2)
1403: =\Lambda(-\eta_1) \int {d^4 p \over (2\pi)^4} e^{-ip(x-x')}
1404: [ -2\pi i\delta (p^2)\not\! p]
1405: \{(\theta(p^0)[1-n^{+}(p)] - (\theta(-p^0)n^{-}(p)\}\Lambda(\eta_2)~.
1406: \label{eq:E3.24}\end{eqnarray}
1407: The expression between the two spin-rotating matrices $\Lambda$ is what is
1408: commonly known for this type correlator in flat Minkowsky space, and it
1409: explicitly depends on the difference, $~x-x'~$, of Cartesian coordinates.
1410: The matrices $\Lambda(\eta)$ corrupt this invariance, because the curvature
1411: of the hypersurface of constant $\tau$ causes the effect known as Thomas
1412: precession of the fermion spin that can be seen by an observer that changes
1413: his rapidity coordinate and thus is subjected to an acceleration in
1414: $z$-direction. From the representation (\ref{eq:E3.24}), it is still
1415: difficult to see that the correlators (\ref{eq:E3.20}) depend only on the
1416: difference $\eta=\eta_1-\eta_2$ (provided the distributions
1417: $n^{\pm}(\theta,p_t)$ are boost invariant). This fact becomes clear after
1418: we change the variable, $\theta = \theta'+(\eta_1+ \eta_2)/2$. Then
1419: \begin{eqnarray}
1420: G_{[10]}(x_1,x_2;p_t)= - i\int{d\theta' \over 8\pi}
1421: \bigg[ \gamma^+~p_te^{-\theta'}+
1422: \gamma^-~p_te^{+\theta'} +
1423: p_r\gamma^r\gamma^0(\gamma^+e^{-{\eta\over 2}}
1424: +\gamma^-e^{+{\eta\over 2}})\bigg]\nonumber\\
1425: \times \bigg[[1- n^{+}\big({\eta_1+\eta_2\over 2}+ \theta',p_t\big)]
1426: e^{-ip_t[\tau_1\cosh(\theta-\eta/2)- \tau_2\cosh(\theta+\eta/2)]}\nonumber\\
1427: + n^{-}\big({\eta_1+\eta_2\over 2}+ \theta',p_t\big)
1428: e^{+ip_t[\tau_1\cosh(\theta-\eta/2)- \tau_2\cosh(\theta+\eta/2)]}\bigg]~.
1429: \label{eq:E3.25} \end{eqnarray}
1430: An amazing property of this formula is that the rapidity argument of the
1431: distributions $n^{\pm}(\theta,p_t)$ is shifted by $(\eta_1+\eta_2)/2$ towards
1432: the geometrical center of the correlator. Now, the spin rotation in the
1433: $(tz)$-plane is virtually eliminated in such a way that both the spin
1434: direction and
1435: the occupation numbers acquired a reference point exactly in the middle
1436: between the endpoints $\eta_1$ and $\eta_2$. Now, things look exactly
1437: as if we had
1438: performed the Wigner transform of the correlator. In actual fact, we did
1439: not. If the distributions $n^{\pm}$ are boost-invariant along some finite
1440: rapidity
1441: interval, then the fermion correlator (\ref{eq:E3.25}) will have the same
1442: property.
1443:
1444: The Wightman function (\ref{eq:E3.25}) has two different parts. One part is
1445: connected with the vacuum density of states. The second ``material'' part is
1446: connected with the occupation numbers. The first one is always boost
1447: invariant. Furthermore, we may expect that it depends (apart from the
1448: spin-rotation effects) only on the invariant interval $\tau_{12}^2=
1449: (t_1-t_2)^2-(z_1-z_2)^2$. The invariance of the material part is limited,
1450: e.g., by the full width $2Y$ of the rapidity plateau and we have to be
1451: careful in the course of further its transformation. In order to extract
1452: the dependence on $\tau_{12}$, we must make a second change of variable,
1453: $\theta' = \theta''+\psi$, where $\psi(\tau_1,\tau_2,\eta)$ depends on the
1454: sign of the interval $\tau_{12}$ between the points $(\tau_1,\eta_1)$ and
1455: $(\tau_2,\eta_2)$. Let the interval $\tau_{12}$ be time-like. Then
1456: \begin{eqnarray}
1457: \tau_{12}^2=\tau_{1}^2+\tau_{2}^2 - 2\tau_1\tau_2\cosh\eta > 0,~~~~
1458: \tanh\psi(\eta)=
1459: {\tau_1+\tau_2\over\tau_1-\tau_2}\tanh{\eta\over 2},\nonumber\\
1460: |\eta|<\eta_0 =\ln{\tau_1\over\tau_2}, ~~~
1461: \tanh\psi(\pm\eta_0)=\pm1,~~~\psi(\pm\eta_0)=\pm \infty~.
1462: \label{eq:E3.26}
1463: \end{eqnarray}
1464: Then, Eq.~(\ref{eq:E3.25}) becomes
1465: \begin{eqnarray}
1466: G_{[10]}(x_1,x_2;p_t)= - i\int{d\theta'' \over 8\pi}
1467: \bigg[ \gamma^+~e^{-\psi}p_te^{-\theta''}+
1468: \gamma^-~e^{\psi}p_te^{+\theta''} +
1469: p_r\gamma^r\gamma^0\bigg(\gamma^+e^{-{\eta\over 2}}
1470: +\gamma^-e^{+{\eta\over 2}}\bigg)\bigg]\nonumber\\
1471: \times \bigg[[1- n^{+}\big({\eta_1+\eta_2\over 2}+\psi + \theta'',p_t\big)]
1472: e^{-ip_t\tau_{12}\cosh\theta''}
1473: + n^{-}\big({\eta_1+\eta_2\over 2}+\psi + \theta'',p_t\big)
1474: e^{+ip_t\tau_{12}\cosh\theta''}\bigg]~,
1475: \label{eq:E3.27} \end{eqnarray}
1476: and we see that the rapidity distributions of particles are shifted by
1477: $\psi(\eta)$ towards the direction between the points $(\tau_1,\eta_1)$ and
1478: $(\tau_2,\eta_2)$. According to (\ref{eq:E3.26}), the rapidity $\psi$ may be
1479: infinite when this direction is light-like ($\tau_{12}^2=0$). Then this
1480: shifted argument appears to be beyond the physical rapidity limits $\pm Y$
1481: of the background distribution $n^{\pm}(\theta,p_t)$. This is extremely
1482: important, since this light-like direction is dangerous; it is solely
1483: responsible for the collinear singularities in various amplitudes. One may
1484: think that the cut-off $\pm Y$ will now appear as a parameter in the final
1485: answer. This would be counter-intuitive, e.g., for many local quantities
1486: related to the central rapidity region, like dynamical masses we are
1487: intending to compute. It will be shown later, that the theory is totally
1488: protected from collinear singularities even in its vacuum part and no
1489: explicit cut-off is necessary.
1490:
1491: For the case of a space-like interval $\tau_{12}$, we introduce
1492: \begin{eqnarray}
1493: \tilde{\tau}_{12}^2=-\tau_{12}^2=-\tau_{1}^2-\tau_{2}^2 +
1494: 2\tau_1\tau_2\cosh\eta >0~, \nonumber\\
1495: \tanh\tilde{\psi}(\eta)=
1496: {\tau_1-\tau_2\over\tau_1+\tau_2}\coth{\eta\over 2},~~~|\eta|>\eta_0~,
1497: \label{eq:E3.28}
1498: \end{eqnarray}
1499: and rewrite Eq.~(\ref{eq:E3.22}) as follows,
1500: \begin{eqnarray}
1501: G_{[10]}(x_1,x_2;\vec{p_t})= - i\int{d\theta'' \over 4\pi}
1502: \bigg[ {1\over 2}\gamma^+~e^{-\tilde{\psi}}p_te^{-\theta''}+
1503: {1\over 2}\gamma^-~e^{\tilde{\psi}}p_te^{\theta''}
1504: +p_r\gamma^r(\cosh{\eta\over 2}
1505: -\gamma^0\gamma^3\sinh{\eta\over 2})\bigg]\nonumber\\
1506: \times \bigg[ [1-
1507: n^{+}\big({\eta_1+\eta_2\over 2}+\tilde{\psi} + \theta'',p_t\big)]
1508: e^{-ip_t\tilde{\tau}_{12}{\rm sign}\eta\sinh\theta''}+
1509: n^{-}\big({\eta_1+\eta_2\over 2}+\tilde{\psi} + \theta'',p_t\big)
1510: e^{+ip_t\tilde{\tau}_{12}{\rm sign}\eta\sinh\theta''}\bigg]~.
1511: \label{eq:E3.29} \end{eqnarray}
1512: Now it is easy to see that we are protected from the null-plane singularities
1513: in the material sector of the theory on both sides of the light-like plane.
1514: Similar calculations can be done for the second Wightman function $G_{[10]}$
1515: which differs from $G_{[10]}$ by the obvious replacements, $1-n^+\to -n^+$, and
1516: $-n^-\to 1-n^-$. The results can be summarized as follows,
1517: \begin{eqnarray}
1518: G_{[10]}(\tau_1,\tau_2,\eta;\theta'', p_t) =[1-n^+(\theta,p_t)]
1519: G^{(0)}_{[10]}(\tau_1,\tau_2,\eta;\theta'', p_t) -
1520: n^-(\theta,p_t)
1521: G^{(0)}_{[01]}(\tau_1,\tau_2,\eta;\theta'',p_t)~~,\nonumber\\
1522: G_{[01]}(\tau_1,\tau_2,\eta;\theta'', p_t) = -n^+(\theta,p_t)
1523: G^{(0)}_{[10]}(\tau_1,\tau_2,\eta;\theta'', p_t) + [1-n^-(\theta,p_t)]
1524: G^{(0)}_{[01]}(\tau_1,\tau_2,\eta;\theta'', p_t)~,
1525: \label{eq:E3.30}
1526: \end{eqnarray}
1527: where according to Eqs.~(\ref{eq:E3.25}), (\ref{eq:E3.27}) and
1528: (\ref{eq:E3.29}),
1529: $\theta=(\eta_1+\eta_2)/ 2+\psi + \theta''$.
1530: Here, $G^{(0)}_{[\alpha]}$ is the vacuum counterpart of $G_{[\alpha]}$,
1531: and $~G^{(0)}_{[01]}(x_1,x_2;\theta,\vec{p_t})=
1532: [G^{(0)}_{[10]}(x_1,x_2;\theta,-\vec{p_t})]^\ast$.
1533: Using Eqs.~(\ref{eq:E3.30}), we may easily obtain the field correlators
1534: defined by Eqs.~(\ref{eq:E3.20a}).
1535: One of them is the causal anti-commutator,
1536: \begin{eqnarray}
1537: G_{[0]}(x_1,x_2;\vec{p_t})\equiv
1538: G_{[10]}(x_1,x_2;\vec{p_t})- G_{[10]}(x_1,x_2;\vec{p_t})=
1539: G^{(0)}_{[10]}(x_1,x_2;\vec{p_t})- G^{(0)}_{[10]}(x_1,x_2;\vec{p_t})~,
1540: \label{eq:E3.31}\end{eqnarray}
1541: which does not include occupation numbers, while the density of states,
1542: \begin{eqnarray}
1543: G_{[1]}(x_1,x_2;\vec{p_t})\equiv
1544: G_{[10]}(x_1,x_2;\vec{p_t})+ G_{[10]}(x_1,x_2;\vec{p_t})
1545: = [1-2n_f(\theta,p_t)]G^{(0)}_{[1]}(x_1,x_2;\vec{p_t})~,
1546: \label{eq:E3.32}\end{eqnarray}
1547: carries all the information about the phase-space population. In the last
1548: equation, we have put $n^-(\theta,p_t)=n^+(\theta,p_t)=n_f(\theta,p_t)$.
1549: It is straightforward to check that
1550: \begin{eqnarray}
1551: G_{[0]}(\tau,\eta_1,\vec{r}_{t1};\tau,\eta_2,\vec{r}_{t2} )
1552: = -i\langle 0|~[\Psi(\tau,\eta_1,\vec{r}_{t1}),
1553: \overline{\Psi}(\tau,\eta_2,\vec{r}_{t2} )]_+ ~|0\rangle=
1554: -i{\gamma^0\over\tau}\delta(\eta_1-\eta_2)\delta(\vec{r}_{t1}-\vec{r}_{t2})~.
1555: \label{eq:E3.32a}\end{eqnarray}
1556: This property of the equal-proper-time commutator is the canonical commutation
1557: relation which is translated into commutation relations for the Fock
1558: operators.
1559: It also verifies that the system of wave functions we employ forms a complete
1560: set.
1561:
1562: In any calculations connected with the local quantities in heavy ion
1563: collisions, we would like to rely on the rapidity plateau in all
1564: distributions and to avoid its width as a parameter in the final answers.
1565: If this is possible (which appears to be the case), then we may consider
1566: the occupation numbers as the functions of $p_t$ only, and accomplish the
1567: integration over the rapidity $\theta''$. This integration gives the vacuum
1568: correlators $G_{[\alpha]}(\tau_1,\tau_2,\eta; \vec{p_t})$ in closed
1569: form.\footnote{Some of the integrals over $\theta''$ are defined as
1570: distributions by means of analytic continuation.} The integrations are
1571: straightforward and result in the following representation of the fermion
1572: correlators,
1573: \begin{eqnarray}
1574: G_{[\alpha]}(\tau_1,\tau_2;\eta,\vec{p_t})=
1575: \gamma^+ ~p_t g^{L(+)}_{[\alpha]} +\gamma^- ~p_t g^{L(-)}_{[\alpha]} +
1576: p_r\gamma^r\gamma^0\gamma^+ ~g^{T(+)}_{[\alpha]} +
1577: p_r\gamma^r\gamma^0\gamma^- ~g^{T(-)}_{[\alpha]}~.
1578: \label{eq:E3.33}
1579: \end{eqnarray}
1580: The products of three gamma-matrices in this expression indicates that the
1581: fermion correlators acquire an axial component ($\sim \gamma^r\gamma^5$),
1582: which is consistent with the absence of complete Lorentz and rotational
1583: symmetry in our problem. In order to obtain the compact expressions for the
1584: invariants $g_{[\alpha]}$, one must note that in all domains, we can replace
1585: \begin{eqnarray}
1586: \bigg( ~e^{\mp\psi},~ \mp{\rm sign}\eta ~ e^{\mp \tilde{\psi}} ~\bigg)
1587: ~\to ~ {\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over \sqrt{|\tau_{12}^2|}}~.
1588: \label{eq:E3.34}
1589: \end{eqnarray}
1590: These transformations lead to the final expressions for the invariants of
1591: the fermion correlators that we shall use in our calculations. For the
1592: invariants of the causal anti-commutator $G_{[0]}$, we have
1593: \begin{eqnarray}
1594: g^{L(\pm)}_{[0]}=
1595: i~{\tau_1 e^{\mp\eta/2}-\tau_2 e^{\pm\eta/2}\over 4\sqrt{|\tau_{12}^2|}}
1596: \theta(\tau_{12}^2)~J_1(p_t\tau_{12})~,~~~~
1597: g^{T(\pm)}_{[0]}= -~{ e^{\mp\eta/2}\over 4}
1598: \theta(\tau_{12}^2)~J_0(p_t\tau_{12})~.
1599: \label{eq:E3.35}
1600: \end{eqnarray}
1601: They are causal in the sense, that they are completely confined to the
1602: interior of the future light wedge. Depending on the context, the invariants of
1603: the density $G_{[1]}$ will be used in two different representations,
1604: \begin{eqnarray}
1605: g^{L(\pm)}_{[1]}= - \int{d\theta' \over 4\pi}
1606: \big[1- 2 n_f\big({\eta_1+\eta_2\over 2}+ \theta',p_t\big)\big]
1607: ~e^{\mp\theta'}
1608: ~\sin\big( p_t[\tau_1\cosh(\theta-\eta/2)-
1609: \tau_2\cosh(\theta+\eta/2)]\big)\nonumber\\
1610: ={\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over 4\sqrt{|\tau_{12}^2|}}
1611: \bigg[\theta(\tau_{12}^2)~Y_1(p_t\tau_{12})
1612: +{2\over\pi}\theta(-\tau_{12}^2)
1613: K_1(p_t\tilde{\tau_{12}})\bigg]
1614: ~\big[ 1-2n_f(p_t) \big]~,
1615: \label{eq:E3.36}
1616: \end{eqnarray}
1617: \begin{eqnarray}
1618: g^{T(\pm)}_{[1]}= -i e^{\mp\eta/2} \int{d\theta' \over 4\pi}
1619: \big[1- 2 n_f\big({\eta_1+\eta_2\over 2}+ \theta',p_t\big)\big]
1620: ~\cos\big( p_t[\tau_1\cosh(\theta-\eta/2)-
1621: \tau_2\cosh(\theta+\eta/2)]\big)\nonumber\\
1622: = i{ e^{\mp\eta/2}\over 4}
1623: \bigg[\theta(\tau_{12}^2)~Y_0(p_t\tau_{12})
1624: -{2\over\pi}\theta(-\tau_{12}^2)
1625: K_0(p_t\tilde{\tau_{12}})\bigg]~\big[ 1-2n_f(p_t) \big]~.
1626: \label{eq:E3.37}
1627: \end{eqnarray}
1628: The first of these representations will be expedient when the quark from the
1629: distribution $n_f(p_t)$ in the self-energy loop is interacting with the
1630: radiation component of the gluon field which imposes the physical limits on
1631: the rapidity $\psi(\eta)$ in the phase, $\Phi=\tau_{12}(\eta) p_t \cosh
1632: (\theta'-\psi(\eta))$, in the integrands of Eqs.~(\ref{eq:E3.36}) and
1633: (\ref{eq:E3.37}). Then, the integration $d\theta'$ will have finite limits
1634: defined by the light cone and the localization of states with the large
1635: $p_t$. When the quark interacts with the longitudinal (static) component of
1636: the gluon field, no limitations of this kind appear and we are able to use
1637: the second analytic representation.
1638:
1639: \vspace{1cm}
1640:
1641:
1642: \noindent {\bf ACKNOWLEDGMENTS}
1643:
1644: The author is grateful to Berndt Muller, Edward Shuryak and Eugene Surdutovich
1645: for helpful discussions at various stages in the development of this work,
1646: and appreciate the help of Scott Payson who critically read the
1647: manuscript.
1648:
1649: This work was partially supported by the U.S. Department of Energy under
1650: Contract No. DE--FG02--94ER40831.
1651:
1652: \bigskip
1653:
1654:
1655: \begin{references}
1656: \bibitem{QFK} A. Makhlin, Phys.Rev. {\bf C 51} (1995) 34
1657: \bibitem{QGD} A. Makhlin, Phys.Rev. {\bf C 52} (1995) 995.
1658: \bibitem{tev} A. Makhlin and E. Surdutovich, Phys.Rev. {\bf C 58} (1998) 389
1659: (quoted as paper [I]).
1660: \bibitem{Gribov} V.N. Gribov, {\em Space-time description of hadron
1661: interactions at high energies}, in Proceedings of the 8-th
1662: Leningrad Nuclear Physics Winter School, February 16-27, 1973.
1663: \bibitem{BPQCD} Yu.L. Dokshitzer et al., {\em Basics of perturbative QCD},
1664: Editions Frontieres, 1991.
1665: \bibitem{Shuryak} E.V. Shuryak, Sov. Phys. JETP {\bf 47} (1978) 212.
1666: \bibitem{geg} A. Makhlin, {\em Scenario for Ultrarelativistic Nuclear
1667: Collisions: III.~ Gluons in the expanding geometry.}
1668: hep-ph/0007301 (quoted as paper [III])
1669: \bibitem{fse} A. Makhlin and E. Surdutovich,
1670: {\em Scenario for Ultrarelativistic Nuclear Collisions:
1671: IV.~Effective quark mass at the early stage.}
1672: hep-ph/0007302(quoted as paper [IV])
1673: \bibitem{Sakharov} A.D. Sakharov, JETP {\bf 18}, 631 (1948).
1674: \bibitem{Landau} L.D. Landau, E.M. Lifshits, Quantum mechanics, Sec.1,
1675: Oxford ; New York : Pergamon Press, 1977
1676: \bibitem{Dirac} P.A.M. Dirac, Rev.Mod. Phys, {\bf 21}, 392 (1949).
1677: \bibitem{Fock} V.A.Fock, Z. f. Phys. {\bf 57}, 261 (1929).
1678: \bibitem{Witten} M.B. Green, J.H. Schwarz, E. Witten, Superstring
1679: theory, Cambridge University press, 1987.
1680: \bibitem{Keld} L.V. Keldysh, Sov. Phys. JETP {\bf 20} (1964) 1018;
1681: E.M. Lifshits, L.P. Pitaevsky, Physical kinetics,
1682: Pergamon Press, Oxford, 1981.
1683: \end{references}
1684:
1685:
1686:
1687: \end{document}
1688:
1689: