1:
2:
3: \documentstyle[aps,eqsecnum,epsf,epsfig]{revtex}
4:
5: \begin{document}
6:
7: \baselineskip=16pt
8: %\baselineskip=20pt
9: %\draft
10:
11: \title{Scenario for Ultrarelativistic Nuclear Collisions: \\
12: IV.~ Effective quark mass at the early stage. }
13: \author{ A. Makhlin and E. Surdutovich}
14: \address{Department of Physics and Astronomy, Wayne State University,
15: Detroit, MI 48202}
16: \date{July 26, 2000}
17: \maketitle
18:
19: \begin{abstract}
20:
21: Using the framework of wedge dynamics, we compute the effective transverse
22: mass of a soft quark mode propagating in the expanding background of hard
23: quarks and gluons created at the earliest time of the collision. We
24: discover that the wedge dynamics does not require any external infrared
25: or collinear cut-off. The effective mass is produced mainly due to the
26: forward quark-quark scattering mediated by the longitudinal (in sense of
27: Gauss' law) magnetic fields. Contribution of the radiation field is
28: parametrically suppressed.
29:
30: \end{abstract}
31:
32: \section{Introduction }
33: \label{sec:S1}
34:
35: In the first paper of this cycle \cite{tev} ( further quoted as paper
36: [I]), we formulated a program that might result in a theory of
37: ultrarelativistic nuclear collisions which is free from collinear problems
38: and naturally establishes the infrared boundary for the space of ``final''
39: states at the very early stage of the collision ($\leq 1fm$). We have
40: demonstrated that even at very early times (much less than is required for
41: any kinetic process to develop), the collective interactions in a dense
42: system provide the final states of the QCD evolution with finite
43: dynamically generated masses that shield mass singularities in the
44: evolution equations.\footnote{ The idea that screening effects should be
45: taken into account at the early kinetic stage of a collision has been
46: articulated earlier and with different motivations by Shuryak and Xiong
47: \cite{Xiong} and by Eskola, Muller and X.-N. Wang \cite{Eskola}}. It was
48: shown also that the null-plane dynamics are incapable of describing local
49: screening effects, because any type of kinetics is frozen on the light
50: cone. It was suggested, that a more adequate approach requires the change
51: of the global Hamiltonian dynamics which is used for the field-theory
52: description of nuclear collisions. We proposed the so-called {\em wedge
53: dynamics} which employs the proper time $\tau$ measured from the first
54: touch of the Lorenz-contracted nuclei as the Hamiltonian time. Our initial
55: estimates in paper [I] were very qualitative. In two consecutive papers
56: \cite{gqm,wdg} (further quoted as papers [II] and [III]), we have
57: studied, in detail, the space of states of wedge dynamics. In paper [II],
58: we extended the qualitative analysis of scalar fields initiated in [I],
59: and found that for the charged fields, the early-time evolution of the
60: wave function is accompanied by a gradual rearrangement of the charge
61: distribution, starting from its almost uniform spread along the light cone
62: at $\tau\to 0$, and up to a narrow wave packet with a well defined
63: rapidity at later times. We have shown that this re-distribution of the
64: charge leads to currents in the rapidity direction and that these currents
65: are the largest at the earliest $\tau$. The magnetic fields generated by
66: these currents can be responsible for the interactions between the
67: currents at the earliest moments of the QCD evolution. In paper [II], we
68: studied the states of fermions in wedge dynamics and found the fermion
69: field correlators that are used below for the calculation of the quark
70: self-energy in the expanding system. In paper [III], we addressed the
71: issue of gauge fields in wedge dynamics. Several important problems were
72: solved there. The natural gauge condition of wedge dynamics, $A^\tau =0$,
73: was proved to be completely fixed (at the level of perturbation theory).
74: The second (technically nearly most difficult) problem solved in paper
75: [III], was the separation of the longitudinal (i.e., governed by Gauss'
76: law) field and the field of radiation. In that paper, we also quantized
77: the gauge field in the scope of wedge dynamics and explicitly found the
78: Wightman functions and retarded propagator of the gluon field which are
79: used in this paper for the practical calculation of the fermion
80: self-energy.
81:
82: Our decision to begin the exploration of potentialities of the wedge
83: dynamics with the computation of quark self-energy is motivated only by
84: technical reasons. The gluon propagator of wedge dynamics is a very
85: complicated function, and we preferred to start with the computation of
86: the fermion loop which has only one gluon correlator in it. We hope that
87: the possibility of a technical simplification (compared to what we had to
88: start with) discovered in this paper, will allow us to address a more
89: important problem of the gluon self-energy in a reasonably economic way.
90:
91: In the course of this study, we employ a single heuristic assumption
92: (supported by the analysis of paper [II]) that the field states with large
93: transverse momentum, even at very early times, may be associated with the
94: localized particles and thus can be described by the distribution with
95: respect to the rapidity and transverse momentum. Our strategy of looking
96: for the leading contributions, as well as all our approximations, in the
97: calculation of the material part of the quark self-energy are based on
98: this assumption. If it appears incorrect, then it is most likely that the
99: quark-gluon matter created in the collision of two nuclei never and in no
100: approximation can be considered as a system of nearly free and weakly
101: interacting field states.
102:
103:
104: \section{Fermion retarded self-energy }
105: \label{sec:S3}
106:
107: In order to find the normal modes of the quark field in the expanding
108: quark-gluon system , we are going to solve the Dirac equation with the
109: radiative corrections, which can be derived as a projection of the
110: Schwinger-Dyson equation for the retarded quark propagator onto the
111: one-particle initial state. For the quark field without Lagrangian mass,
112: this equation reads
113: \begin{eqnarray}
114: i \gamma^\mu (x_1) \nabla_\mu (x_1) \psi(x_1)=
115: \int d^4 x_2 \Sigma_{[ret]}(x_1,x_2) \psi(x_2)~.
116: \label{eq:E3.0a}\end{eqnarray}
117: The covariant derivative $\nabla_\mu (x)$ of the spinor field in the
118: curvilinear coordinates of the wedge dynamics includes the spin connection
119: and it was found explicitly in paper [II]. For all calculations below, we
120: employ the mixed representation which is the most profitable in heavy-ion
121: problems. We are looking for the radiative corrections to the wave
122: function with a given transverse momentum $\vec{p_t}$ and rapidity
123: $\theta$ with the expectation that within the rapidity plateau nothing
124: will depend on $\theta$. However, we cannot totally eliminate the
125: coordinate $\eta$ from the theory. We have to keep it explicitly, since
126: the problem of the expanding field system cannot be reduced to (2+1)
127: dimensions. In its expanded form, Eq.~(\ref{eq:E3.0a}) reads,
128: \begin{eqnarray}
129: \bigg[i\gamma^0\big({\partial\over\partial\tau_1}+{1\over 2\tau_1}\big)
130: +{i\gamma^3\over\tau_1}{\partial\over\partial\eta_1} -p_r\gamma^r\bigg]
131: \psi (p_t;\tau_1,\eta_1)=
132: \int_{0}^{\tau_1}d\tau_2\int_{-\infty}^{\infty}\tau_2 d\eta_2
133: \Sigma_{[ret]}(p_t;\tau_1,\tau_2;\eta_1-\eta_2)\psi (p_t;\tau_2,\eta_2)~.
134: \label{eq:E3.0b}
135: \end{eqnarray}
136: The retarded self-energy is an object that naturally emerges in the
137: Schwinger-Dyson equation for the retarded propagator in Keldysh-Schwinger
138: formalism \cite{Keld}. Below, we employ its modified form developed
139: earlier with the view of application to the inclusive and transient
140: processes. We employ the notation used in
141: Refs.~\cite{QFK,QGD,tev}.\footnote{The indices of the field correlators
142: with the Keldysh contour ordering of the field operators (like $G_{[AB]}$)
143: as well as the labels of their linear combinations (like $G_{[ret]}$) are
144: placed in square brackets.} In this notation, the one-loop retarded
145: fermion self-energy in coordinate form is
146: \begin{eqnarray}
147: \Sigma_{[ret]}(x_1,x_2)={ig^{2}\over 2}
148: [t^{a} \gamma^{\mu} G_{[ret]}(x_1,x_2)t^{b}\gamma^{\lambda}
149: D^{ba}_{[1]\lambda\mu}(x_2,x_1)+t^{a} \gamma^{\mu} G_{[1]}(x_1,x_2)
150: t^{b}\gamma^{\lambda} D^{ba}_{[adv]\lambda\mu}(x_2,x_1)]~.
151: \label{eq:E3.1}
152: \end{eqnarray}
153: The two subprocesses that contribute this self-energy are depicted below.
154: \begin{figure}[htb]
155: \begin{center}
156: \mbox{
157: \psfig{file=./gret.ps,height=2.in,bb=130 460 440 675}
158: \hspace{2.cm}
159: \psfig{file=./gadv.ps,height=2.in,bb=140 460 430 670}
160: }
161: \end{center}
162: \caption{The retarded forward scattering amplitude is contributed by two
163: subprocesses, $qg\to qg$ and $qq\to qq$.}
164: \label{fig:fig4}
165: \end{figure}
166:
167: The retarded and advanced quark and gluon propagators $G_{[ret]}$ and
168: $D^{lm}_{[adv]}$ were found in papers [II] and [III] of this cycle
169: and are connected with the commutators $G_{[0]}$ and $D_{[0]}$,
170: \begin{eqnarray}
171: G_{[ret]}(x_1,x_2)=\theta(\tau_1-\tau_2)G_{[0]}(x_1,x_2),~~
172: D^{lm}_{[adv]}(x_2,x_1)=-\theta(\tau_1-\tau_2)D^{lm}_{[0]}(x_2,x_1)
173: +D^{lm}_{L}(x_2,x_1)~,
174: \label{eq:E3.2}
175: \end{eqnarray}
176: where $D^{lm}_{L}(x_2,x_1)$ is the longitudinal part of the gluon
177: propagator (governed by Gauss' law), and it enters in Eq.~(\ref{eq:E3.2})
178: in such a way that the condition $D_{[ret]}-D_{[adv]}=D_{[0]}$ is
179: satisfied and the non-causal longitudinal part of the propagator does not
180: violate the causal properties of the commutator $D_{[0]}$. The
181: correlators $G_{[1]}$ and $D_{[1]}$ include densities of vacuum states as
182: well as the information about the occupation numbers (phase-space
183: population). Eventually, we shall prove that an approximation of the
184: boost-invariance (infinite rapidity plateau) is not corrupted by any kind
185: of cut-offs (the vacuum part never is). Therefore, all correlators ($G$,
186: $D$, and $\Sigma$) will depend on two times $\tau_1$ and $\tau_2$
187: separately, the difference of rapidities $\eta=\eta_1-\eta_2$, and the
188: difference ${\vec r} ={\vec r_1}-{\vec r_2}$ of distances in $xy$ plane.
189: The latter is Fourier transformed to the transverse momentum dependence.
190: In this mixed representation,
191: \begin{eqnarray}
192: \Sigma_{[ret]}(\tau_1,\tau_2;\eta,\vec{p_t})={ig^{2}\over 2 (2\pi)^2}
193: \int d^2 \vec{k_t} [t^{a} \gamma^m(\tau_1)
194: G_{[ret]}(\tau_1,\tau_2;\eta,\vec{p_t}+\vec{k_t})
195: t^{b}\gamma^l(\tau_2)
196: D^{ba}_{[1]lm}(\tau_2,\tau_1;-\eta,\vec{k_t}) \nonumber \\
197: +t^{a} \gamma^m(\tau_1)
198: G_{[1]}(\tau_1,\tau_2;\eta,\vec{p_t}+\vec{k_t})
199: t^{b}\gamma^l(\tau_2)
200: D^{ba}_{[adv]lm}(\tau_2,\tau_1;-\eta,\vec{k_t})]~,
201: \label{eq:E3.4}
202: \end{eqnarray}
203: where $\gamma^{\eta}(\tau)=\gamma^3/\tau$.
204: As it has been shown in [I], all fermion correlators $G_{[\alpha]}$ can be
205: decomposed as
206: \begin{eqnarray}
207: G_{[\alpha]}(\tau_1,\tau_2;\eta,\vec{q_t})=
208: q_t \big[g^{0}_{[\alpha]}\gamma^0 +g^{3}_{[\alpha]}\gamma^3\big] +
209: g^{T}_{[\alpha]}q_r\gamma^r +
210: i g^{A}_{[\alpha]}q_r\epsilon^{ru}\gamma^u\gamma^5 \nonumber\\
211: =q_t \big[ g^{L(+)}_{[\alpha]}\gamma^+
212: +g^{L(-)}_{[\alpha]}\gamma^-\big] +
213: q_r\gamma^r\gamma^0 ~\big[ g^{T(+)}_{[\alpha]}\gamma^+
214: +g^{T(-)}_{[\alpha]}\gamma^-\big]~,
215: \label{eq:E3.5}
216: \end{eqnarray}
217: where, for the sake of brevity, we denote $\vec{q_t}=\vec{p_t}+\vec{k_t}$.
218: A similar decomposition takes place for the self-energy,
219: \begin{eqnarray}
220: \Sigma_{[ret]}(\tau_1,\tau_2;\eta,\vec{p_t})=
221: \Sigma^{0}\gamma^0 +\Sigma^{3}\gamma^3 +
222: \Sigma^{T}q_r\gamma^r +
223: i \Sigma^{A}p_r\epsilon^{ru}\gamma^u\gamma^5 \nonumber\\
224: = \Sigma^{L(+)}\gamma^+ +\Sigma^{L(-)}\gamma^- +
225: p_r\gamma^r\gamma^0~\big[ \Sigma^{T(+)}\gamma^+
226: +\Sigma^{T(-)}\gamma^-\big]~,
227: \label{eq:E3.6}
228: \end{eqnarray}
229: and we obviously have,
230: \begin{eqnarray}
231: g^{L(\pm)}_{[\alpha]}={1\over 2}(g^{0}_{[\alpha]}\pm g^{3}_{[\alpha]}),~~~~
232: g^{T(\pm)}_{[\alpha]}={1\over 2}
233: (g^{T}_{[\alpha]}\pm g^{A}_{[\alpha]})~,~~~~
234: \Sigma^{L(\pm)}={1\over 2}(\Sigma^{0}\pm \Sigma^{3}),~~~~
235: \Sigma^{T(\pm)}={1\over 2}(\Sigma^{T}\pm \Sigma^{A})~.
236: \label{eq:E3.7}
237: \end{eqnarray}
238:
239: It becomes easier to analyze the various pieces of the quark self-energy
240: if the gluon correlators $D_{[\alpha]lm}$ are taken in the form of the
241: following decomposition,\footnote{In what follows, we use the Greek
242: indices for the four-dimensional vectors and tensors in the curvilinear
243: coordinates (index $\eta$ is an exception, it always denotes the rapidity
244: direction), and the Latin indices from $a$ to $d$ for the vectors in flat
245: Minkowski coordinates. We use Latin indices from $r$ to $w$ for the
246: transverse $x$- and $y$-components ($r,...,w=1,2$), and the arrows over
247: the letters to denote the two-dimensional vectors, {\em e.g.}, ${\vec
248: k}=(k_x,k_y)$, $|{\vec k}|=k_{t}$. The Latin indices from $i$ to $n$
249: ($i,...,n=1,2,3$) will be used for the three-dimensional internal
250: coordinates $u^i=(x,y,\eta)$ on the hyper-surface $\tau=const$.}
251: \begin{eqnarray}
252: D_{[\alpha]rs}=\bigg(\delta_{rs}- {k_r k_s\over k_t^2}\bigg)
253: {\cal D}^{(TE)}_{[\alpha]}+
254: {k_r k_s\over k_t^2}{\cal D}^{(2)}_{[\alpha]},~~~
255: D_{[\alpha]\eta\eta}={\cal D}^{(\eta\eta)}_{[\alpha]},~~~
256: D_{[\alpha]r\eta}={k_r\over k_t^2}{\cal D}^{(r\eta)}_{[\alpha]},~~~
257: D_{[\alpha]\eta s}={k_s\over k_t^2}{\cal D}^{(\eta s)}_{[\alpha]}~,
258: \label{eq:E3.8}
259: \end{eqnarray}
260: where the first term in $D_{[\alpha]rs}$ is due to the transverse
261: electric mode, and all invariants of ${\cal D}_{[adv]}$ (except ${\cal
262: D}^{(TE)}_{[adv]}$ ) have two terms, ${\cal D}_{[0]}^{(\cdot\cdot\cdot)}$
263: from the transverse magnetic mode of the radiation field, and ${\cal
264: D}_{[long]}^{(\cdot\cdot\cdot)}$ from the longitudinal field. All these
265: components were found in paper [III] and are given in the Appendix in the
266: form which is used in the calculation below. After some algebra, we can
267: present the retarded quark self-energy in the form,
268: \begin{eqnarray}
269: \Sigma_{[ret]}(\tau_1,\tau_2;\eta,\vec{p_t})={i\alpha_s C_F\over 2\pi}
270: \int d^2 \vec{k_t} \big[\gamma^+ S^{L(+)}+\gamma^- S^{L(-)}+
271: p_r \gamma^r\gamma^0~\big(\gamma^+ S^{T(+)}
272: +\gamma^- S^{T(-)}\big)\big]~,
273: \label{eq:E3.9} \end{eqnarray}
274: where the scalar invariants of $\Sigma_{[ret]}$ are the bilinears of the
275: fermion and gluon scalars,
276: \begin{eqnarray}
277: S^{L(\pm)}=\sum_{[\alpha,\beta]} \bigg\{q_t
278: g^{L(\pm)}_{[\alpha]}({\cal D}^{(TE)}_{[\beta]}+{\cal D}^{(2)}_{[\beta]})
279: +{q_t\over\tau_1\tau_2}g_{[\alpha]}^{L(\mp)}
280: {\cal D}^{(\eta\eta)}_{[\beta]}
281: %\nonumber\\
282: \mp{(\vec{k_t} \vec{q_t})\over k_t^2}
283: \bigg({g^{T(\pm)}_{[\alpha]}{\cal D}^{(r\eta)}_{[\beta]}\over\tau_1}+
284: {g^{T(\mp)}_{[\alpha]}{\cal D}^{(\eta r)}_{[\beta]}\over\tau_2}
285: \bigg)\bigg\},
286: \label{eq:E3.10}\end{eqnarray}
287: \begin{eqnarray}
288: S^{T(\pm)}=\sum_{[\alpha,\beta]}
289: \bigg\{\bigg[{(\vec{p_t} \vec{q_t})\over p_t^2}
290: -2{(\vec{k_t} \vec{p_t})(\vec{k_t} \vec{q_t})\over k_t^2 p_t^2}\bigg]
291: g^{T(\pm)}_{[\alpha]}({\cal D}^{(TE)}_{[\beta]}+{\cal D}^{(2)}_{[\beta]})
292: -{(\vec{q_t} \vec{p_t})\over p_t^2\tau_1\tau_2}
293: g^{T(\mp)}_{[\alpha]}{\cal D}^{(\eta\eta)}_{[\beta]}\nonumber\\
294: \mp {(\vec{k_t} \vec{p_t}) \over k_t^2 p_t^2}
295: \bigg({g^{L(\pm)}_{[\alpha]} {\cal D}^{(r\eta)}_{[\beta]}\over\tau_1}-
296: {g^{L(\mp)}{\cal D}^{(\eta r)}_{[\beta]}\over\tau_2}\bigg)\bigg\} ~.
297: \label{eq:E3.11}
298: \end{eqnarray}
299: In these equations, the sum $\sum_{[\alpha,\beta]}$ runs over
300: $[\alpha,\beta]= \{[ret,1], [1, adv]\}$.
301:
302:
303: \section{Fermion modes in the expanding system }
304: \label{sec:S2}
305:
306: We shall look for the dispersion law of the fermions in the proper-time
307: dynamics studying the Dirac equation (\ref{eq:E3.0b}) with radiative
308: corrections. Since the fermions are massless, it is convenient to use the
309: spinor basis where the Dirac matrices are
310:
311: \begin{eqnarray}
312: \gamma^0= \left( \begin{array}{cc} 0 & I \\ I & 0 \end{array} \right),
313: ~~~~ \gamma^l= \left( \begin{array}{cc} 0 & -\sigma^l
314: \\ \sigma^l & 0 \end{array} \right)~, \nonumber
315: \end{eqnarray}
316: and the Dirac equation can be split into two
317: separate equations for the left- and right-handed two-component spinors.
318: The latter reads as
319: \begin{eqnarray}
320: G_{R}^{-1}(p_t;\tau_1,\eta_1)~\psi_{R}(p_t;\tau_1,\eta_1)=
321: \int_{0}^{\tau_1}d\tau_2\int_{-\infty}^{\infty}\tau_2 d\eta_2
322: \Sigma_R(p_t;\tau_1,\tau_2;\eta_1-\eta_2)\psi_{R}(p_t;\tau_2,\eta_2)~,
323: \label{eq:E2.1}
324: \end{eqnarray}
325: where the matrices of the right-handed differential operator $G_{R}^{-1}$
326: and of the right-handed self-energy $\Sigma_{R}$ are
327: \begin{eqnarray}
328: G_{R}^{-1}(p_t;\tau,\eta)\left[ \begin{array}{cc}
329: i(\partial_\tau+{1\over 2\tau}-
330: {1\over \tau}\partial_\eta) & p_x-ip_y \\
331: p_x+ip_y & i(\partial_\tau+{1\over 2\tau}+
332: {1\over \tau}\partial_\eta)
333: \end{array} \right],\nonumber \\
334: \Sigma_R(p_t;\tau_1,\tau_2;\eta_1-\eta_2)=
335: \left[ \begin{array}{cc}
336: \Sigma^{L(-)} & -(p_x-ip_y)\Sigma^{T(+)} \\
337: -(p_x+ip_y)\Sigma^{T(-)}& \Sigma^{L(+)} \end{array} \right] .
338: \label{eq:E2.2}
339: \end{eqnarray}
340: The equation for the left-handed spinors differs from (\ref{eq:E2.1}) only
341: by a change of some signs in matrices (\ref{eq:E2.2}) and leads to the
342: same dispersion law. A solution with positive energy is looked for in the
343: form,
344: \begin{eqnarray}
345: \psi_{R}(p_t,\theta;\tau,\eta)= \left( \begin{array}{c}
346: e^{(\eta-\theta)/2)}~p_t \\
347: -e^{-(\eta-\theta)/2)}~(p_x+ip_y) \end{array} \right)
348: ~e^{-i\mu\tau\cosh(\eta-\theta)} ~,
349: \label{eq:E2.3}\end{eqnarray}
350: where $\mu$ is the effective ``transverse mass'' of the mode.
351: For the free on-mass-shell solution we have $\mu=p_t$.
352: To solve Eq.~(\ref{eq:E2.1}), we introduce an auxiliary (left-handed)
353: spinor
354: \begin{eqnarray}
355: \tilde{\psi}(p_t,\theta';\tau,\eta)= \left( \begin{array}{c}
356: e^{-(\eta-\theta')/2)}~p_t \\
357: -e^{(\eta-\theta')/2)}~(p_x-ip_y) \end{array} \right)
358: ~e^{i\mu\tau\cosh(\eta-\theta')} ~.
359: \label{eq:E2.4}\end{eqnarray}
360: We insert (\ref{eq:E2.3}) into (\ref{eq:E2.1}), multiply it from the left
361: by (\ref{eq:E2.4}) and integrate this along the hypersurface
362: $\tau_1=const$.
363: Then the left side of the equation becomes
364: \begin{eqnarray}
365: \int_{-\infty}^{\infty}\tau_1 d\eta_1
366: \tilde{\psi}(p_t,\theta`;\tau_1,\eta_1)G_{R}^{-1}(p_t;\tau_1,\eta_1)
367: \psi_{R}(p_t,\theta;\tau_1,\eta_1)=
368: 4\pi{\mu-p_t\over\mu}~p_t^2~\delta(\theta-\theta')~.
369: \label{eq:E2.5}\end{eqnarray}
370: In deriving this equation, we assumed that $\mu$ is independent of $\tau_1$.
371: The weak dependence is admissible, provided $d\mu/d\tau_1\ll\mu/\tau_1$.
372: A solution that has this property does indeed exist. The right hand side
373: of the equation is, in fact, independent of $\theta'$ and is of the
374: following form,
375: \begin{eqnarray}
376: p_t^2~\int_{0}^{\tau_1}d\tau_2\int_{-\infty}^{\infty}\tau_1 \tau_2 d\eta_2
377: d\theta d(\eta_1-\eta_2)e^{i\mu\tau_1\cosh(\eta_1+\theta)}
378: e^{-i\mu\tau_2\cosh\eta_2} \nonumber \\
379: \times [e^{-{\eta_1-\eta_2+\theta\over 2}}\Sigma^{(L)-}+
380: e^{{\eta_1-\eta_2+\theta\over 2}}\Sigma^{(L)+}+
381: e^{-{\eta_1+\eta_2+\theta\over 2}}\Sigma^{(T)+}+
382: e^{{\eta_1+\eta_2+\theta\over 2}}\Sigma^{(T)-}]~,
383: \label{eq:E2.6}\end{eqnarray}
384: where the exponentials are due to the Thomas precession of the spinor field.
385: Next, we integrate both sides with respect to $\theta$. Two rapidity
386: integrals, $d\theta d\eta_2$, on the right absorb the precession factors
387: yielding the product of Hankel functions,
388: \begin{eqnarray}
389: \pi^2 H^{(1)}_{1/2}(\mu\tau_1)~H^{(2)}_{1/2}(\mu\tau_2)=
390: {2\pi\over\mu\sqrt{\tau_1\tau_2}}e^{i\mu(\tau_1-\tau_2)}~.
391: \label{eq:E2.7}\end{eqnarray}
392: Finally, we arrive at the dispersion equation that defines the fermion
393: ``transverse mass'' $\mu$ as a function of the transverse momentum and the
394: latest time $\tau_1$,
395: \begin{eqnarray}
396: \mu(p_t,\tau_1)-p_t ={1\over 2}
397: \int_{0}^{\tau_1}d\tau_2 \sqrt{\tau_1\tau_2}
398: e^{i\mu(p_t,\tau_1)(\tau_1-\tau_2)} \int_{-\infty}^{\infty}d\eta~
399: [\Sigma^{L(+)}+\Sigma^{L(-)}+p_t \Sigma^{T(+)} +p_t \Sigma^{T(-)}]~.
400: \label{eq:E2.8}\end{eqnarray}
401:
402: As it has been discussed in paper [II] for fermions (similar arguments
403: are true for gluons), only the independence of the quark and gluon
404: occupation numbers $n_f$ and $n_g$ on rapidity can provide that the
405: invariants $S^{L(\pm)}$ and $S^{T(\pm)}$ naturally depend only on the
406: difference $\eta=\eta_1-\eta_2$. We shall consider only this case of the
407: local homogeneity; we can do it safely only because no collinear
408: singularities which may require a rapidity cut-off (e.g., $\pm Y$) in
409: the phase space will appear in the theory. Since we are computing an
410: essentially local quantity, such a cut-off would be unphysical. With
411: this reservation, we may rewrite Eq.~(\ref{eq:E2.8}) as
412: \begin{eqnarray}
413: \mu(p_t) = p_t +
414: \int_{0}^{\tau_1}d\tau_2 \sqrt{\tau_1\tau_2} e^{i\mu(p_t)(\tau_1-\tau_2)}
415: [\Sigma^{0}(\tau_1,\tau_2)+ p_t \Sigma^{T}(\tau_1,\tau_2)]~,
416: \label{eq:E3.12}\end{eqnarray}
417: where we introduced the notation,
418: \begin{eqnarray}
419: \Sigma^{0}(\tau_1,\tau_2)=
420: {i\alpha_s C_F\over 4\pi}\sum_{[\alpha,\beta]} \int d^2 \vec{k_t}
421: \int_{-\infty}^{\infty}d\eta~ q_t g^{0}_{[\alpha]}
422: [{\cal D}^{(TE)}_{[\beta]}+{\cal D}^{(2)}_{[\beta]}
423: +{1\over \tau_1\tau_2}{\cal D}^{(\eta\eta)}_{[\beta]}],
424: \label{eq:E3.13}\end{eqnarray}
425: \begin{eqnarray}
426: \Sigma^{T}(\tau_1,\tau_2)={i\alpha_s C_F\over 4\pi}\sum_{[\alpha,\beta]}
427: \int d^2\vec{k_t}\int_{-\infty}^{\infty}d\eta~
428: g^{T}_{[\alpha]} \bigg\{
429: \bigg[{(\vec{p_t} \vec{q_t})\over p_t^2}
430: -2{(\vec{k_t} \vec{p_t})(\vec{k_t} \vec{q_t})\over k_t^2 p_t^2}\bigg]
431: ({\cal D}^{(TE)}_{[\beta]} +{\cal D}^{(2)}_{[\beta]})+
432: {(\vec{q_t} \vec{p_t})\over p_t^2\tau_1\tau_2}
433: {\cal D}^{(\eta\eta)}_{[\beta]}\bigg\}.
434: \label{eq:E3.13a}\end{eqnarray}
435: Comparing these equations with Eqs.~(\ref{eq:E3.10}) and (\ref{eq:E3.11}), we
436: may observe a significant simplification. The terms with the off-diagonal
437: components ${\cal D}^{(\eta r)}$ and ${\cal D}^{(r\eta)}$ have dropped out.
438: These terms, as it can be seen from Eqs.~(\ref{eq:A1.6})-- (\ref{eq:A1.7}),
439: (\ref{eq:A1.10})--(\ref{eq:A1.11}), and
440: (\ref{eq:A1.16})--(\ref{eq:A1.17}), are odd with respect to $\eta$, while
441: the invariants $g^{0}=g^{L(+)}+g^{L(-)}$ and $g^{T}=g^{T(+)}+g^{T(-)}$ are
442: even. Therefore, integration over $\eta$ eliminates the terms with the
443: off-diagonal components.
444:
445:
446:
447: \section{Propagators, densities of states, and occupation numbers
448: in the expanding system }
449: \label{sec:S4}
450:
451: In this section, we collect condensed information about various correlators
452: of quark and gluon fields derived in papers [II] and [III] which are
453: necessary for the calculation of the quark self-energy. We also discuss our
454: specific choice of occupation numbers $n_g(k_t,\alpha)$ and
455: $n_f(q_t,\theta)$. All field correlators are defined as the expectation
456: values over the distribution of the background particles. The latter are
457: the excitations of the modes allowed by the constraints and the boundary
458: conditions of wedge dynamics. The Fock space of these excitations was
459: constructed in papers [II] and [III]. We have analyzed two sets of quantum
460: numbers that may label the states. Both sets include the transverse
461: momentum $\vec{p_t}$ and polarization index. In one set, the remaining
462: variable was the boost $\nu$ (the variable conjugated to the coordinate
463: $\eta$); this set proved to be very useful in the practical calculation of
464: the gluon propagators. In the second set, the particles are labeled by
465: their velocity $v_z=\tanh\theta$ in the direction of the collision axis.
466: This representation is used below. The fermion spectral functions are
467: \begin{eqnarray}
468: G_{[10]}(q_t,\theta;\tau_1,\tau_2) =[1-n_f^+(q_t,\theta)]
469: G^{(0)}_{[10]}(q_t,\theta;\tau_1,\tau_2) -
470: n_f^-(q_t,\theta) G^{(0)}_{[01]}(q_t,\theta;\tau_1,\tau_2)~~,\nonumber\\
471: G_{[01]}(q_t,\theta;\tau_1,\tau_2) = -n_f^+(q_t,\theta)
472: G^{(0)}_{[10]}(q_t,\theta;\tau_1,\tau_2) + [1-n_f^-(q_t,\theta)]
473: G^{(0)}_{[01]}(q_t,\theta;\tau_1,\tau_2)~.
474: \label{eq:E4.1}\end{eqnarray}
475: Their gluon counterparts are of a similar form,
476: \begin{eqnarray}
477: D_{[10]}(k_t,\alpha;\tau_1,\tau_2) =[1+n_g(k_t,\alpha)]
478: D^{(0)}_{[10]}(k_t,\alpha;\tau_1,\tau_2) +
479: n_g(q_t,\alpha) D^{(0)}_{[01]}(q_t,\alpha;\tau_1,\tau_2)~,\nonumber\\
480: D_{[01]}(k_t,\alpha;\tau_1,\tau_2) =n_g(k_t,\alpha)
481: D^{(0)}_{[10]}(k_t,\alpha;\tau_1,\tau_2) +
482: [1+ n_g(q_t,\alpha)] D^{(0)}_{[01]}(q_t,\alpha;\tau_1,\tau_2)~,
483: \label{eq:E4.2}
484: \end{eqnarray}
485: where $D^{(0)}_{[\alpha]}$ and $G^{(0)}_{[\alpha]}$ are the vacuum
486: correlators of a given type $[\alpha]$. They are defined as vacuum
487: expectation values of the binary products of field operators,
488: \begin{eqnarray}
489: G^{(0)}_{[10]}(x_1,x_2) =
490: -i\langle 0|\Psi (x_1)\overline{\Psi}(x_2)|0\rangle~,~~~~~
491: G^{(0)}_{[01]}(x_1,x_2)=
492: i\langle 0|\overline{\Psi}(x_2)\Psi (x_1)|0\rangle~, \nonumber\\
493: D^{(0)}_{[10]lm}(x_1,x_2) =
494: -i\langle 0|A_ll(x_1) A(x_2)_m|0\rangle~,~~~~
495: D^{(0)}_{[01]lm}(x_1,x_2)=-i\langle 0|A_m(x_2)A_l(x_1)|0\rangle~.
496: \label{eq:E4.2a}
497: \end{eqnarray}
498: In this approximation, the field (anti-)commutators,
499: \begin{eqnarray}
500: G_{[0]}= G_{[10]}-G_{[01]}=G^{(0)}_{[10]}-G^{(0)}_{[01]}=
501: G^{(0)}_{[0]}~,\nonumber\\
502: D_{[0]}= D_{[10]}-D_{[01]}=D^{(0)}_{[10]}-D^{(0)}_{[01]}=
503: D^{(0)}_{[0]}~
504: \label{eq:E4.3}
505: \end{eqnarray}
506: appear to be insensitive to the presence of the particle distribution,
507: while their counterparts,
508: \begin{eqnarray}
509: G_{[1]}= G_{[10]}+G_{[01]}= [1-2n_f]G^{(0)}_{[1]}~,\nonumber\\
510: D_{[1]}= D_{[10]}+D_{[01]}= [1+2n_g]D^{(0)}_{[1]}~,
511: \label{eq:E4.4}
512: \end{eqnarray}
513: include the occupation numbers which modify the original vacuum density
514: of states. For the sake of simplicity, we take $n_f^+=n_f^-=n_f$, which
515: corresponds to a neutral system.
516:
517: The Wightman functions (\ref{eq:E4.1}) and (\ref{eq:E4.2}) (or their
518: various linear combinations $G_{[\beta]}$ and $D_{[\beta]}$) eventually
519: appear under the integrals $d\theta$ and $d\alpha$. One must keep in mind
520: that in order to reduce $G^{(0)}_{[\beta]}$ and $D^{(0)}_{[\beta]}$ to
521: the standard form of the vacuum correlators, at least two shifts of the
522: integration variables is necessary. Only after that will
523: $G^{(0)}_{[\beta]}$ and $D^{(0)}_{[\beta]}$ explicitly depend on the
524: boost-invariant variables $\eta$ and $\tau_{12}$. The functions
525: $n_g(k_t,\alpha)$ and $n_f(q_t,\theta)$ are not indifferent to this shift.
526: It may well happen that a formal shift in $\theta$ or $\alpha$ will drive
527: the stationary points of the wave functions or the singularities of the
528: field correlators outside the physical boundaries of the distributions
529: $n_g(k_t,\alpha)$ and $n_f(q_t,\theta)$. Therefore, different
530: representations of $G_{[1]}$ and $D_{[1]}$ must be used for the study of
531: different subprocesses. One has to account for the reservations stemming
532: from the derivation procedure described in Sec.~4 of paper [II]. These
533: different representations of the quark and gluon correlators are quoted in
534: the Appendix.
535:
536: In our picture, first outlined in paper [I], the fermion vacuum mode with
537: small transverse momentum $p_t$ and zero rapidity is modified by its
538: forward scattering either on gluons with high momentum $k_t$ and rapidity
539: $\alpha$, $k_t\gg p_t$ or on quarks with high momentum $q_t$ and rapidity
540: $\theta$, $q_t\gg p_t$. These hard modes are created at the earliest
541: moment of the collision and can be treated as well formed particles by the
542: time $\tau\sim 1/p_t$, since at that time $\tau k_t\gg 1$, and $\tau q_t\gg
543: 1$. Therefore, they may be consistently described by the distributions,
544: \begin{eqnarray}
545: n_f(q_t,\theta)\approx {{\cal N}_f\over \pi R_\bot^2}
546: {\theta (q_t-p_\ast) \over q_t^2},~~~
547: n_g(k_t,\alpha)\approx {{\cal N}_g\over \pi R_\bot^2}
548: {\theta (k_t-p_\ast) \over k_t^2}~.
549: \label{eq:E4.5}
550: \end{eqnarray}
551: where $p_\ast$ is the lower bound of the ``hard'' partons distribution.
552: Both distributions (per unit area, per unit rapidity) are chosen on
553: purely dimensional grounds, since we believe that the creation of a parton
554: with large transverse momentum is described by perturbative QCD which has
555: no intrinsic scale.
556:
557: Currently, the normalization factors ${\cal N}_g$ and ${\cal N}_f$ are the
558: only (apart from the coupling $\alpha_s$) parameters of the theory. The
559: cross section $\pi R_\bot^2$ and the full width $2Y$ of the rapidity
560: plateau are defined by the geometry of a particular collision and the
561: c.m.s. energy, respectively. These are irrelevant for the local screening
562: parameters we are interested in. In the first approximation, one may try to
563: extract them from the event-by-event measurement of the high-$p_t$ tail of
564: the collision products and incorporating the standard phenomenology of the
565: fragmentation functions for the analysis.
566:
567: As it was pointed out in paper [I], even in dense systems, the QCD
568: evolution at large $Q^2$ is not likely to be affected by finite-density
569: effects. Thus, one may also try to employ the known structure functions
570: (without shadowing corrections) and the factorization scheme in order to
571: estimate ${\cal N}_g$ and ${\cal N}_f$. A most appealing opportunity to
572: find $n_g(k_t,\alpha)$ and $n_f(q_t,\theta)$ from first principles,
573: associating them with the known properties of hadrons and the QCD vacuum,
574: is still very distant.
575:
576: The distributions (\ref{eq:E4.5}) are used below with the following
577: informal reservations. First, the total energy of any collision is finite
578: and $k_t$ and $q_t$ have (though very high, but finite) upper boundary.
579: Eventually, this leads to the self-energy which is free from collinear
580: singularities in the interaction of charges with the vector gauge field.
581: Second, though the distributions (\ref{eq:E4.5}) are boost-invariant, only
582: the particles which physically affect the forward scattering must be
583: accounted for. There is a strong correlation between the position $\eta$
584: where the particle with large transverse momentum $q_t$ is measured (or is
585: interacting) and its rapidity $\theta$. Hence, the limits of integrals
586: $d\alpha$ and $d\theta$ over the rapidities of real particles (which either
587: mediate the scattering or are in the final states) cannot exceed the actual
588: rapidity boundaries of the scattering process. In its turn, this puts an
589: additional requirement on the notion of the distribution itself. It must be
590: normalizeable in the physical volume of the reaction. This volume is
591: defined, in fact, by the light cone (i.e. causality of the forward
592: scattering amplitude). [We remind the reader that the notion of a
593: distribution itself makes sense only after it is prepared (measured) at
594: least in a {\em gedanken} experiment. Hence, the distributions $n_g$ and
595: $n_f$ must exist, in this sense, both at final time $\tau_1$ and at the
596: initial time $\tau_2$ in the expression for the self-energy. In its turn,
597: this limits the time $\tau_2$ from below.]
598:
599: \section{Leading part of the dispersion equation}
600: \label{sec:S3a}
601:
602:
603: \subsection{Derivation of the dispersion equation}
604: \label{subsec:Sb3a}
605:
606:
607: The most important outcome of this work is that the major contribution to
608: the effective quark mass comes from the $\eta\eta$-component of the
609: propagator of the longitudinal field. This contribution is computed in all
610: details below. All other terms are associated with the propagation of the
611: transverse fields and they appear to be parametrically small in the domain
612: $\tau_1p_t <1$, $(\tau_1-\tau_2) p_t \ll 1$, $\tau_1-\tau_2 \ll\tau_1$,
613: where the dynamical mass of the fermion is effectively formed. The
614: component $D_{\eta\eta}$ of the propagator establishes the connection
615: between the $A_\eta$ component of the potential and the $j_\eta$ component
616: of the current. In its turn, $A_\eta$ is responsible for the
617: $\eta$-component $E_\eta=\partial_\tau A_\eta$ of the electric field and
618: the $x$- and $y$-components, $B_x=\partial_y A_\eta$, $B_y=-\partial_x
619: A_\eta$ of the magnetic field. The electrical field in the longitudinal
620: $\eta$-direction is not capable of producing scattering with transverse
621: momentum transfer. However, this transfer can be provided by the magnetic
622: forces; the two currents $j_\eta$ can interact via the magnetic field
623: ${\vec B}_t=(B_x,B_y)$. The origin of these currents is intrinsically
624: connected with the geometry of states in the wedge form of dynamics. Any
625: state with a given $p_t$ begins its life being widely spread along the
626: light cone. If the state is charged, then local charge density is small.
627: With time going on, the spread of the wave function diminishes and the
628: charge become localized in a narrower rapidity interval (see
629: Fig.~\ref{fig:fig0}). Therefore, any charged state carries a current in
630: the longitudinal (rapidity) direction.
631:
632: \begin{figure}[htb]
633: \begin{center}
634: \mbox{ \psfig{file=./fig1b.ps,height=2.2in,bb=100 480 400 715} }
635: \end{center}
636: \caption{Evolution of the charge density in the typical state of wedge
637: dynamics}
638: \label{fig:fig0}\end{figure}
639: The magnetic fields of the {\em transition currents} provide scattering
640: with the most effective transfer of the transverse momentum. Indeed, at
641: time $\tau_2$ a quark with the transverse momentum $p_t$, $\tau_2 p_t\ll
642: 1$, interacts with the gluon field and acquires a large transverse momentum
643: $k_t$, $\tau_2 k_t\gg 1$. This transition is characterized by a drastic
644: narrowing of the charge spread in the rapidity direction, and must be
645: accompanied by a strong $\eta$-component of the transition current. A
646: similar transition in the opposite direction happens at time $\tau_1$, when
647: the gluon field interacts with another quark that has large initial
648: transverse momentum $k_t$, and recovers the soft state with $\tau_2 p_t\ll
649: 1$ in the course of this interaction. This second transition current
650: readily interacts with the magnetic component of the gluon field. Our
651: estimates indicate that the leading contribution comes from the term of
652: $D_{\eta\eta}(\tau_2,\tau_1; \eta_2-\eta_1;\vec{k_t})$ which is
653: proportional to $\delta (\eta_1-\eta_2)$ and does not depend on ${\vec
654: k_t}$ [in coordinate representation, this term is just proportional to
655: $\delta (\eta_1-\eta_2) \delta (\vec{r_1}-\vec{r_2})$ ]. This is a
656: long-range contact interaction of the two currents, and is not limited by
657: the light-cone boundaries (which suppress the interaction via the strongly
658: localized states of the radiation field). Furthermore, the contact part of
659: the longitudinal propagator is the only one that brings into the integrand
660: of the dispersion equation (\ref{eq:E3.12}) the term which is {\em
661: singular} at $\tau_1-\tau_2\to 0$. Therefore, it is capable of providing
662: an appreciable contribution into the effective quark mass, which is {\em
663: defined locally}. This part of the self-energy allows for an exact
664: calculation with a simple analytic answer which is presented below.
665: Estimates of all small terms are explained in Sec.~\ref{sec:S6} and
666: \ref{sec:SA2}\footnote{ The authors appreciate discussions with Edward
667: Shuryak, who pointed out that the small effect of the radiation fields is
668: much less surprising than the finite contribution from the longitudinal
669: fields.}. This contact part of the longitudinal propagator is
670: \begin{eqnarray}
671: D^{[contact]}_{\eta\eta}(\tau_2,\tau_1;\eta_2-\eta_1;\vec{k_t})
672: =-{\tau_1^2-\tau_2^2 \over 2}\delta (\eta).
673: \label{eq:E3.14}\end{eqnarray}
674: Because of the extreme locality of $D^{[contact]}_{\eta\eta}$ provided
675: by $\delta(\eta)$, the invariants $g_{[1]}$ of the fermion density function
676: in (\ref{eq:E3.13}) lose their kinematic coefficients,
677: \begin{eqnarray}
678: g^{0}_{[1]}= {-2{\cal N}_f\over \pi R_t^2}~
679: {Y_1(\tau_{12}q_t)\over q_t^2}~,~~~~~~
680: g^{T}_{[1]}= {-2{\cal N}_f\over \pi R_t^2}~{Y_0(\tau_{12}q_t)\over q_t^2}~.
681: \label{eq:E3.16}\end{eqnarray}
682: First, we integrate over $\eta$, which leads to $\tau_{12}^{2}=
683: (\tau_1-\tau_2)^2$. Next, we change $d^2\vec{k_t}$ for $d^2\vec{q_t}$ and
684: integrate over the orientation of $\vec{q_t}$ gaining the factor $2\pi$ in
685: $\Sigma^{0}$. In $\Sigma^{T}$, $g^{T}_{[1]}$ is integrated with the weight
686: factor $(\vec{q}_t\cdot\vec{p}_t)/p_t^2$. Therefore, this term identically
687: vanishes after integration over the azimuthal angle. The only remaining
688: integral over the transverse momenta of hard partons is
689: \begin{eqnarray}
690: \int_{p_\ast}^{\infty} Y_1[(\tau_1 -\tau_2)q_t]~d q_t =
691: {Y_0[(\tau_1 -\tau_2)p_\ast]\over\tau_1 -\tau_2}~.
692: \label{eq:E3.17}\end{eqnarray}
693: Eventually, we may write the dispersion equation (\ref{eq:E3.12}) as
694: follows,
695: \begin{eqnarray}
696: \mu = p_t + {i\alpha_s C_F {\cal N}_f\over 2\pi R_t^2}
697: \int_{0}^{\tau_1}d\tau_2 {\tau_1+\tau_2\over 2\sqrt{\tau_1\tau_2}}
698: e^{i\mu(\tau_1-\tau_2)} Y_0[(\tau_1 -\tau_2)p_\ast] ~.
699: \label{eq:E3.18}\end{eqnarray}
700:
701: \subsection{Study of the dispersion equation.}
702: \label{subsec:Sb3b}
703:
704: According to the qualitative analysis of paper [II], the dynamics of states
705: is different in the two limiting cases, $\tau p_t <1$ and $\tau p_t >1$.
706: With respect to $p_t\sim 1/\tau$, the states are divided into ``hard'' and
707: ``soft'' states. Therefore, it is natural to take $p_\ast\sim p_t$ in
708: Eq.~(\ref{eq:E3.18}). Further, it is convenient to trade variable $\tau_2$
709: for $y=(\tau_1 -\tau_2)/\tau_1$,
710: \begin{eqnarray}
711: {\mu(p_t,\tau_1)\over p_t} = 1+
712: {i\alpha_s C_F {\cal N}_f \tau_1 p_t\over 2\pi R_t^2 p_t^2}
713: \int_{0}^{1}dy {1-y/2\over \sqrt{1-y}}
714: e^{i\mu(p_t,\tau_1)\tau_1y} Y_0(\tau_1 p_t y) ~.
715: \label{eq:E3.19}\end{eqnarray}
716: In this form, the dispersion equation clearly reveals two distinctive
717: regimes. When $\tau p_t <1$, then the function $Y_0(x)$ behaves as a
718: logarithm, and the right hand side of Eq.~(\ref{eq:E3.19}) becomes
719: proportional to $\ln(2/\tau_1 p_t)$, the effective width of the rapidity
720: interval occupied by the state at the early time of the evolution. When
721: $\tau p_t >1$, then $Y_0(x)\sim 1/\sqrt{x}$, and the integral becomes
722: proportional to $1/\sqrt{\tau p_t}$, the effective rapidity width at
723: later times. Thus, the dispersion equation (\ref{eq:E3.19}) clearly
724: reveals two distinctive regimes which were qualitatively analyzed in paper
725: [II]. The solution of Eq.~(\ref{eq:E3.19}) is generally complex. Taking
726: $\mu=\mu'+i\mu''$ we can separate real and imaginary parts of this
727: equation,
728: \begin{eqnarray}
729: \tau_1\mu'- \tau_1 p_t =
730: - {\alpha_s C_F {\cal N}_f \over 2\pi (R_t^2 /\tau_1^2)}
731: \int_{0}^{1}{dy\over 2}~\bigg[ ~{1\over \sqrt{1-y}}+\sqrt{1-y}\bigg] ~
732: e^{-\mu''\tau_1 y}~
733: \sin (\mu'\tau_1 y) Y_0(\tau_1 p_t y) ~,
734: \label{eq:E3.20}\end{eqnarray}
735: \begin{eqnarray}
736: \tau_1\mu'' =
737: {\alpha_s C_F {\cal N}_f \over 2\pi (R_t^2/\tau_1^2)}
738: \int_{0}^{1}dy ~\bigg[ ~{1\over \sqrt{1-y}}+\sqrt{1-y}\bigg]~
739: e^{-\mu'' \tau_1 y}~
740: \cos (\mu'\tau_1 y) Y_0(\tau_1 p_t y) ~,
741: \label{eq:E3.21}
742: \end{eqnarray}
743: (The unit upper limit in these integrals corresponds to $\tau_2=0$, and is,
744: as a matter of fact, fictitious. Practically, we are interested only in
745: the domain where $\tau_2 p_t\sim 1$.) We have rearranged the factor in
746: front of the integral in such a way, that at early times, this factor is
747: small. It has been shown in paper [III], that the longitudinal part of the
748: gluon propagator vanishes when the distance $r_t$ exceeds $\tau_1$.
749: Therefore, this factor is proportional to the (small) number of hard
750: partons per transverse area occupied by the soft quark mode. Hence, we can
751: analyze Eqs,~(\ref{eq:E3.20}) and (\ref{eq:E3.21}) by successive
752: approximations. It is clear, that in the lowest approximation, we can take
753: $\mu'=p_t$ in the RHS of these equations, and that the imaginary part
754: $\mu''$ can be neglected. Using $$\sin x\approx x~,~~~ \cos x\approx 1~,
755: ~~~ Y_0(x) \approx 2\pi^{-1} [\gamma_E +\ln x]~, $$ as an approximation,
756: and computing the remaining integrals, we arrive at
757: \begin{eqnarray}
758: {\mu'\over p_t} -1 =
759: {\alpha_s C_F {\cal N}_f \over \pi^2 R_t^2 p_t^2}~(\tau_1 p_t)^2~
760: \bigg[ {4\over 5}\bigg( -\gamma_E +\ln{2\over \tau_1 p_t}\bigg)
761: +{104-120\ln 2\over 75}\bigg]~,
762: \label{eq:E3.20a}\end{eqnarray}
763: \begin{eqnarray}
764: {\mu'' \over p_t} =
765: {\alpha_s C_F {\cal N}_f \over \pi^2 R_t^2 p_t^2}~\tau_1 p_t~
766: \bigg[{4\over 3}(\gamma_E -\ln{2\over \tau_1 p_t}\bigg)
767: + {26 - 24\ln 2\over 9}\bigg]~.
768: \label{eq:E3.21a}
769: \end{eqnarray}
770: These dependences are plotted below as functions
771: of the argument $\tau_1 p_t$ up to the pre-factor
772: $\alpha_s C_F {\cal N}_f /\pi^2 R_t^2 p_t^2 $.
773:
774: \begin{figure}[htb]
775: \begin{center}
776: \mbox{
777: \psfig{file=./dis2.ps,height=2.2in,bb=95 535 395 725}
778: }
779: \end{center}
780: \caption{Self-energy corrections to the real part ($\mu'/ p_t -1$,
781: upper curve) and to the imaginary part ($\mu''/ p_t $,
782: lower curve) of the effective transverse mass as functions
783: of $\tau_1p_t$.}
784: \label{fig:fig1}
785: \end{figure}
786: Eq.~(\ref{eq:E3.19}) describes the evolution of the effective transverse
787: mass $\mu$ of the state with a given transverse momentum $p_t$ as a
788: function of the proper time $\tau_1$. We see, that the real part $\mu'$
789: gradually grows with time reaching its maximum at $\tau_1p_t\approx 1$. The
790: mode acquires an ``adjoint mass'' due to the interaction with hard
791: partons, as was anticipated. The curves cannot be trusted above the
792: boundary $\tau_1\approx p_t^{-1}$, since at later times, the mode becomes
793: ``hard''. It cannot be viewed as a soft cloud swept with uniformly
794: distributed hard particles. The condition $$ {d\mu \over d\tau_1}\ll{\mu
795: \over \tau_1},$$ which was used in the course of the dispersion equation
796: derivation, is clearly fulfilled near the maximum of the dispersion curve.
797:
798: One more important dependence is hidden in the pre-factor
799: $\alpha_s C_F {\cal N}_f /\pi^2 R_t^2 p_t^2 $, and is not visible from the
800: figure above. This factor scales as $p_t^{-2} $, clearly indicating that at
801: large $p_t$, the effect of screening is small. There is almost no hard
802: particles with $k_t,q_t>p_t$.
803:
804:
805: \section{Cancelation of collinear terms in the vacuum part of
806: the self-energy}
807: \label{sec:S5}
808:
809: Usually, the self-energy $\Sigma$ is studied in the momentum
810: representation, and the first subject of concern is the ultraviolet
811: divergence of this function. It is well known that this divergence can be
812: at most logarithmic. Thus, when we compute the four-dimensional integral
813: over the momentum in the loop, this divergence can show up only in the last
814: of these integrations. We compute the self-energy in the mixed
815: representation. Hence, we cannot see the UV divergence explicitly but we
816: {\em must} already see (e.g., in ${\rm Im}\Sigma$) the various infrared
817: divergences which emerge due to real processes with massless fields. A
818: corresponding analysis for the case of the null-plane dynamics was done in
819: paper [I]. These divergences must be regularized (or even removed from the
820: theory, as is done by means of dimensional regularization) before the UV
821: renormalization. The primary goal of this section is to demonstrate that in
822: the wedge form of dynamics the quark self-energy is completely protected
823: from collinear problems, and that this is not a surprise. Indeed, in the
824: theory with massless fermions and gauge bosons, the infrared singularities
825: show up in a different way depending on the type of Hamiltonian dynamics
826: (including the gauge condition) which is used to describe the process. In
827: the gauge $A^+ =0$ they look like collinear divergences. In the gauge $A^0
828: =0$, they look like an infrared problem of the proper field of the charged
829: particle. In both cases, the problem emerges due to the incomplete gauge
830: fixing, and manifests itself through spurious poles of the gauge field
831: propagators. As has been shown in paper [III], the gauge $A^\tau =0$ is
832: fixed completely, and therefore, the quark self-energy that we compute here
833: is totally free of these problems.
834:
835: In order to demonstrate this appealing feature we shall compute (the most
836: dangerous in this respect) the vacuum part of the fermion self-energy,
837: concentrating on the terms where the integrand as a function of the
838: rapidity $\alpha$ is not suppressed at $~|\alpha|\to\infty~$. A
839: self-consistent piece of this type is the contribution of the transverse
840: electric mode of the gluon field. The tensor part of any gluon correlator
841: for this mode is of a very simple form,
842: \begin{eqnarray}
843: D_{[\alpha]rs}^{TE}=(\delta_{rs} -k_rk_s/k_t^2)~{\cal D}_{[\alpha]}^{TE}~;
844: \label{eq:E5.1}
845: \end{eqnarray}
846: it has no $\eta$ components, and the scalar functions
847: ${\cal D}_{[\alpha]}^{TE}$ can be computed exactly since they
848: have simple integral representations.
849: We use this piece of the self-energy to explain the principles we base our
850: calculations on.
851: Using Eqs.~(\ref{eq:E3.13}) and (\ref{eq:E3.13a}), we get
852: \begin{eqnarray}
853: \big[\Sigma_{[ret]}^{L(\pm)}\big]^{TE}_{vac}(\tau_1,\tau_2;\eta,\vec{p_t})=
854: {i\alpha_sC_F\over 2\pi}\theta(\tau_{12}^2)\theta(\tau_1-\tau_2)
855: \int d^2\vec{k_t}\big[q_t g^{L(\pm)}_{[0]}(q_t){\cal D}_{[1]}^{TE}(k_t)-
856: q_t g^{L(\pm)}_{[1]}(q_t){\cal D}_{[0]}^{TE}(k_t)\big]~,
857: \label{eq:E5.3}
858: \end{eqnarray}
859: where $\vec{q_t}=\vec{k_t}+\vec{p_t}$ and the minus sign in the second term
860: is due to the definition (\ref{eq:E3.2}) of $D_{[adv]}$. The
861: $\Sigma_{[ret]}^{TE}$ is fully confined within the light wedge
862: $\tau_{12}^2 >0$.
863: Then, the vacuum quark and gluon correlators have the following form,
864: \begin{eqnarray}
865: {\cal D}_{[0]}^{TE}(\tau_2,\tau_1;\eta_2-\eta_1;\vec{k_t})=
866: 2^{-1}\theta(\tau_{12}^{2})J_0(\tau_{12}k_t)~,~~~~
867: {\cal D}_{[1]}^{TE}(\tau_2,\tau_1;\eta_2-\eta_1;\vec{k_t})=
868: 2^{-1}i~Y_0(\tau_{12}k_t)~,
869: \label{eq:E5.2a}
870: \end{eqnarray}
871: \begin{eqnarray}
872: g^{L(\pm)}_{[0]}=
873: i~{\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over 4\sqrt{|\tau_{12}^2|}}
874: \theta(\tau_{12}^2)~J_1(q_t\sqrt{|\tau_{12}^2|})~,~~~~
875: g^{L(\pm)}_{[1]}=
876: {\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over 4\sqrt{|\tau_{12}^2|}}
877: ~Y_1(q_t\sqrt{|\tau_{12}^2|})~,
878: \label{eq:E5.2b}
879: \end{eqnarray}
880: where $\tau_{12}^2=\tau_{1}^2+\tau_{2}^2 - 2\tau_1\tau_2\cosh\eta$, and
881: the commutators $D_{[0]}^{TE}$ and $G_{[0]}$ include the causal
882: $\theta(\tau_{12}^{2})$ by definition, and the terms proportional
883: to $\theta(-\tau_{12}^{2})$ in the densities $D_{[1]}^{TE}$ and
884: $G_{[1]}$ are omitted.
885: Integration over the angle $\varphi$ between the vectors $\vec{p_t}$ and
886: $\vec{k_t}$ involves only the invariants $g_{[\alpha]}$. According to
887: Eqs.(\ref{eq:E5.3}) and (\ref{eq:E5.2b}), we have to integrate,
888: \begin{eqnarray}
889: \int_{0}^{2\pi}q_t J_1(\tau_{12}q_t)d\varphi=
890: 2\pi~[ k_t J_0(\tau_{12}p_t)J_1(\tau_{12}k_t)+
891: p_t J_1(\tau_{12}p_t)J_0(\tau_{12}k_t)]~,\nonumber\\
892: \int_{0}^{2\pi}q_t Y_1(\tau_{12}q_t)d\varphi=
893: 2\pi~\{\theta(k_t-p_t)[ k_t J_0(\tau_{12}p_t)Y_1(\tau_{12}k_t)+
894: p_t J_1(\tau_{12}p_t)Y_0(\tau_{12}k_t)]+\nonumber\\
895: +\theta(p_t-k_t)[ k_t J_1(\tau_{12}k_t)Y_0(\tau_{12}p_t)+
896: p_t Y_1(\tau_{12}p_t)J_0(\tau_{12}k_t)] \}~.
897: \label{eq:E5.4}
898: \end{eqnarray}
899: This integration is done with the aid of the so called addition
900: theorems \cite{Watson} for Bessel functions of the argument
901: $~q_t=[k_t^2+p_t^2+2k_tp_t\cos\varphi]^{1/2}$. Starting from this point, we
902: can continue in two ways. The most straightforward option is to use the
903: gluon correlators in the integrated form of the Bessel functions, thus
904: sweeping under the rug the singular behavior stemming from
905: $\alpha\to\infty$. This leads to the integral
906: \begin{eqnarray}
907: \big[\Sigma_{[ret]}^{L(\pm)}\big]^{TE}=
908: {i\alpha_sC_F\over 8}\theta(\tau_{12}^2)\theta(\tau_1-\tau_2)
909: {\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over \tau_{12}}\nonumber\\
910: \times\bigg[ - J_0(\tau_{12}p_t)
911: \int_{0}^{\infty}k_t^2[J_1(\tau_{12}k_t)Y_0(\tau_{12}k_t)+
912: J_0(\tau_{12}k_t)Y_1(\tau_{12}k_t)]dk_t -
913: 2p_tJ_1(\tau_{12}p_t)
914: \int_{0}^{\infty}k_tJ_0(\tau_{12}k_t)Y_0(\tau_{12}k_t)dk_t\nonumber\\
915: +p_tJ_1(\tau_{12}p_t)
916: \int_{0}^{p_t}k_tJ_0(\tau_{12}k_t)Y_0(\tau_{12}k_t)dk_t+
917: J_0(\tau_{12}p_t)
918: \int_{0}^{p_t}k_t^2J_0(\tau_{12}k_t)Y_1(\tau_{12}k_t)dk_t\nonumber\\
919: -p_t Y_1(\tau_{12}p_t)
920: \int_{0}^{p_t}k_tJ_0(\tau_{12}k_t)J_0(\tau_{12}k_t)dk_t
921: -Y_0(\tau_{12}p_t)
922: \int_{0}^{p_t}k_t^2J_0(\tau_{12}k_t)J_1(\tau_{12}k_t)dk_t\bigg]
923: \label{eq:E5.5}
924: \end{eqnarray}
925: Here, all the integrals can be computed explicitly as indefinite integrals.
926: The integrals from $0$ to $p_t$ yield the regular part of the answer
927: below. Taking the upper limit of improper integrals to be $\Lambda\to\infty$,
928: we get the singular part as the limit,
929: \begin{eqnarray}
930: \lim_{\Lambda\to\infty} {\Lambda^2\over\tau}\bigg[[J_0(\tau_{12}p_t)
931: +\tau p_tJ_1(\tau_{12}p_t)]J_1(\tau\Lambda)Y_1(\tau\Lambda)+
932: \tau p_tJ_1(\tau_{12}p_t)J_0(\tau\Lambda)Y_0(\tau\Lambda)\bigg]~.
933: \label{eq:E5.6}
934: \end{eqnarray}
935: Using the asymptotic expansion of Bessel functions, we find that
936: the singular part is built from $\delta(\tau_{12})$ and its derivative.
937: The full answer reads as
938: \begin{eqnarray}
939: \big[\Sigma_{[ret]}^{L(\pm)}\big]^{TE}=
940: -{i\alpha_sC_F\over 16\pi}\theta(\tau_{12}^2)\theta(\tau_1-\tau_2)
941: {\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over \tau_{12}}\nonumber\\
942: \times\bigg\{ p_t^2 {J_2(\tau_{12}p_t)\over\tau_{12}}
943: +\pi\bigg[{J_0(\tau_{12}p_t)\over\tau_{12}}\bigg(\delta'(\tau_{12})
944: -{2\over\tau_{12}}\delta(\tau_{12})\bigg)-
945: {4p_tJ_1(\tau_{12}p_t)\over\tau_{12}}\delta(\tau_{12})\bigg]\bigg\}~.
946: \label{eq:E5.7}
947: \end{eqnarray}
948: Thus, the answer is singular at the null planes $\tau_{12}^2=0$. In order to
949: find the true source of this singularity, we shall proceed in a different
950: manner, keeping the gluon invariants $g^{L(\pm)}$ in the integral form. We
951: shall integrate over $k_t$ first and leave the integral over the gluon
952: rapidity $\alpha$ for the end of calculation. This leads to
953: \begin{eqnarray}
954: \big[\Sigma_{[ret]}^{L(\pm)}\big]^{TE}=
955: {i\alpha_sC_F\over 8}\theta(\tau_{12}^2)\theta(\tau_1-\tau_2)
956: {\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over \tau_{12}}
957: \bigg\{-~{ p_t^2\over 2\pi} {J_2(\tau_{12}p_t)\over\tau_{12}}\nonumber\\
958: +\int_{-\infty}^{\infty}{d\alpha\over\pi}~
959: \bigg[ J_0(\tau_{12}p_t)
960: \bigg(\int_{0}^{\infty}k_t^2J_1(\tau_{12}k_t)\cos (T_{12}k_t) dk_t-
961: \int_{0}^{\infty}k_t^2 Y_1(\tau_{12}k_t)
962: \sin (T_{12}k_t) dk_t\bigg)\nonumber\\
963: +p_tJ_1(\tau_{12}p_t)
964: \bigg(\int_{0}^{\infty}k_tJ_0(\tau_{12}k_t)\cos (T_{12}k_t) dk_t-
965: \int_{0}^{\infty}k_t Y_0(\tau_{12}k_t)
966: \sin (T_{12}k_t) dk_t\bigg)\bigg]\bigg\}~,
967: \label{eq:E5.8}
968: \end{eqnarray}
969: where, we remind the reader that,
970: $T_{12}=T_1-T_2=\tau_{12}\cosh(\alpha-\psi)$. The integrals
971: in this expression are the well-known Fourier transforms of the Bessel
972: functions,
973: \begin{eqnarray}
974: \int_{0}^{\infty}k_t^2J_1(\tau_{12}k_t)\cos (T_{12}k_t) dk_t=
975: \int_{0}^{\infty}k_t^2 Y_1(\tau_{12}k_t)\sin (T_{12}k_t) dk_t=
976: 3T_{12}\tau_{12}\big[T_{12}^{2}-\tau_{12}^2\big]_{+}^{-5/2}~,\nonumber\\
977: \int_{0}^{\infty}k_tJ_0(\tau_{12}k_t)\cos (T_{12}k_t) dk_t=
978: \int_{0}^{\infty}k_t Y_0(\tau_{12}k_t)\sin (T_{12}k_t) dk_t=
979: -T_{12}\big[T_{12}^{2}-\tau_{12}^2\big]_{+}^{-3/2}~,
980: \label{eq:E5.9}
981: \end{eqnarray}
982: where the distribution $x^{\lambda}_{+}$ is defined in a standard way with
983: the due number of subtracted terms of the Taylor expansion in the integral
984: $\int f(x) x^{\lambda}_{+}dx$ \cite{Gel'fand}. Three different issues are
985: important here. First, each of the integrals (\ref{eq:E5.9}) is a well
986: defined distribution that includes all necessary regulators which provide
987: the convergence of subsequent integrations. The Bessel functions themselves
988: are {\em defined} as the Fourier transforms of the (+)-distributions and we
989: just recover the original regular form by doing the inverse Fourier
990: transform (\ref{eq:E5.9}) (see Ref.\cite{Gel'fand})~. Second, as it will be
991: shown in the next section, the singular behavior of the integrals
992: (\ref{eq:E5.9}) originates from the collinear domain. The (+)-prescription
993: that emerges here eliminates them term-by-term. Third, after the result of
994: the term-by-term integration is put back into Eq.~(\ref{eq:E5.8}), the
995: singular collinear terms just cancel everywhere, including the null-plane
996: $\tau_{12}^2=0$. This type of cancelation of collinearly singular terms
997: takes place in all other pieces of the vacuum part of the quark self-energy.
998:
999: All these observations lead us to the conclusion, that even in its vacuum
1000: part, the self-energy does not suffer from collinear problems, which seems
1001: to be a unique property of the expanding system. We do not continue to study
1002: the vacuum part of the self-energy here, since we are currently interested
1003: only in its material part which is discussed in the next section. [~The full
1004: analysis of this part, including the issue of its renormalization, will be
1005: published elsewhere.~]
1006:
1007:
1008:
1009: \section{Radiation fields in the material part of the self-energy}
1010: \label{sec:S6}
1011:
1012: We found that the major contribution to the one-loop effective mass of a
1013: ``soft'' quark mode in the background of ``hard'' quarks and gluons comes
1014: from the quark-quark forward scattering mediated by the magnetic component
1015: of the longitudinal field. The purpose of this section is to demonstrate
1016: that the interactions via transverse fields (including the forward
1017: scattering of soft quark on hard gluons) is a secondary effect at least at
1018: the very early stage of the nuclear collision. This conclusion may look
1019: counter-intuitive, since, namely in the interactions of the transverse
1020: fields, we expect to encounter the collinear enhancement of the radiation
1021: amplitude. As it has been shown in Sec.\ref{sec:S5}, in the vacuum part of
1022: the self-energy, the integrals of this type (taken in the limits from $0$ to
1023: $\infty$) cancel each other leaving the vacuum sector free from collinear
1024: divergences. The statistical weights ${\cal N}_g$ and ${\cal N}_f$, which
1025: are different for the different terms, prevent such a cancelation in the
1026: material part. Thus we have to analyze each term of the material part
1027: separately.
1028:
1029: As in Sec.\ref{sec:S5}, we consider an isolated piece which corresponds to
1030: TE-gluons. The gluon correlators of this piece are the most singular and are
1031: known not only in the integral representation, but in closed analytic form
1032: also. The last circumstance is very helpful for the analysis of the multiple
1033: integrals we meet below. (The terms identical to those computed below, also
1034: appear in the part of self-energy due to the TM-gluons; the remaining terms
1035: of TM-sector are less singular and, eventually, smaller than considered
1036: here.) The corresponding fragment of the quark self-energy (\ref{eq:E3.13})
1037: in the dispersion equation (\ref{eq:E3.12}) is
1038: \begin{eqnarray}
1039: \bigg[\Sigma^{0}(\tau_1,\tau_2)\bigg]^{(TE)}_{mat}=
1040: {i\alpha_s C_F\over 4\pi}~\theta(\tau_1-\tau_2)\int d^2 \vec{k_t}
1041: \int_{-\infty}^{\infty}d\eta~ q_t \big[ g^{0}_{[0]} {\cal D}^{(TE)}_{[1]}
1042: - g^{0}_{[1]} {\cal D}^{(TE)}_{[0]}\big] ,
1043: \label{eq:E6.01}\end{eqnarray}
1044: \begin{eqnarray}
1045: \bigg[\Sigma^{T}(\tau_1,\tau_2)\bigg]^{(TE)}_{mat}=
1046: {i\alpha_s C_F\over 4\pi}~
1047: \theta(\tau_1-\tau_2)\int d^2\vec{k_t}\int_{-\infty}^{\infty}d\eta~
1048: \bigg[{(\vec{p_t} \vec{q_t})\over p_t^2}
1049: -2{(\vec{k_t} \vec{p_t})(\vec{k_t} \vec{q_t})\over k_t^2 p_t^2}\bigg]
1050: \big[ g^{T}_{[0]}{\cal D}^{(TE)}_{[1]}-g^{T}_{[1]}{\cal D}^{(TE)}_{[0]}\big].
1051: \label{eq:E6.02}\end{eqnarray}
1052: The invariants of (anti-)commutators are the same as in the vacuum case,
1053: \begin{eqnarray}
1054: g^{0}_{[0]}(\tau_1,\tau_2;\eta;\vec{q_t})=
1055: i~{(\tau_1 -\tau_2)\cosh(\eta/2)\over 2\tau_{12}}
1056: \theta(\tau_{12}^2)~J_1(q_t\tau_{12})~,~~~~
1057: g^{T}_{[0]}(\tau_1,\tau_2;\eta;\vec{q_t})=
1058: -{\cosh(\eta/2)\over 2}
1059: \theta(\tau_{12}^2)~J_0(q_t\tau_{12})~,
1060: \label{eq:E6.03}\end{eqnarray}
1061: and the invariant ${\cal D}_{[0]}^{TE}(\tau_2,\tau_1;\eta;\vec{k_t})$ is
1062: given by the first of Eqs.~(\ref{eq:E5.2a}). They all differ from zero only
1063: for the timelike $\tau_{12}$. Hence, the material part of the invariants
1064: $g_{[1]}$ and ${\cal D}^{(TE)}_{[1]}$ will be needed only in this domain. As
1065: it has been discussed earlier, the distributions include only ``hard''
1066: particles which are defined with respect to the soft mode with the
1067: transverse momentum $p_t$ by the inequalities, $k_t>p_\ast$, and
1068: $q_t>p_\ast$, where $p_\ast\geq p_t$. Now, we are interested only in the
1069: material part with occupation numbers given by the equations,
1070: \begin{eqnarray}
1071: n_f(q_t,\theta)\approx {{\cal N}_f\over \pi R_\bot^2}
1072: {\theta (q_t-p_\ast) \over q_t^2} ,~~~~
1073: n_g(k_t,\alpha)\approx {{\cal N}_g\over \pi R_\bot^2}
1074: {\theta (k_t-p_\ast) \over k_t^2} ~,
1075: \label{eq:E6.04}
1076: \end{eqnarray}
1077: and we must keep in mind the width $2Y$ of the rapidity plateau with the goal
1078: to study if this is a significant parameter for the calculation of local
1079: quantities. We may also question the validity of these formulae at
1080: sufficiently large $k_t$ and $q_t$, since without a cutoff, the integral
1081: $\int dk_t/k_t$ diverges.
1082:
1083: The material part of the densities will be employed in two different forms,
1084: \begin{eqnarray}
1085: g^{0}_{[1]}(\tau_1,\tau_2;\eta;\vec{q_t})=
1086: \int_{-\infty}^{\infty} {d\theta\over \pi}
1087: n_f(\theta;q_t)\cosh\theta \sin q_t T_{12}(\theta)
1088: =-~{(\tau_1 -\tau_2)\cosh(\eta/2)\over \tau_{12}}
1089: n_f(q_t)~Y_1(q_t\tau_{12})~,
1090: \label{eq:E6.05}
1091: \end{eqnarray}
1092: \begin{eqnarray}
1093: g^{T}_{[1]}(\tau_1,\tau_2;\eta;\vec{q_t})=
1094: -~i\cosh{\eta\over 2}\int_{-\infty}^{\infty} {d\theta\over \pi}
1095: n_f(\theta;q_t)~\cos q_t T_{12}(\theta)
1096: =-~i\cosh(\eta/2)~n_f(q_t)~Y_0(q_t\tau_{12})
1097: \label{eq:E6.06}
1098: \end{eqnarray}
1099: \begin{eqnarray}
1100: {\cal D}_{[1]}^{(TE)}(\tau_2,\tau_1;\eta;\vec{k_t})=
1101: (\pi i)^{-1}\int_{-\infty}^{\infty} d\alpha
1102: n_g(\alpha;k_t) \cos k_t T_{12}(\alpha)
1103: =i~Y_0(\tau_{12}k_t)~n_g(k_t)~,
1104: \label{eq:E6.07}
1105: \end{eqnarray}
1106: where the second equation in (\ref{eq:E6.05})-(\ref{eq:E6.07}) is valid
1107: only when $n_g$ and $n_f$ are rapidity-independent, and we employ the
1108: following notation,
1109: \begin{eqnarray}
1110: T_{12}(\alpha)=\tau_1\cosh(\alpha-\eta/2)-\tau_2\cosh(\alpha+\eta/2)=
1111: \tau_{12}\cosh(\alpha -\psi)~,\nonumber\\
1112: \tau_{12}^2=\tau_{1}^2+\tau_{2}^2 - 2\tau_1\tau_2\cosh\eta > 0,~~~~
1113: \tanh\psi(\eta)=
1114: {\tau_1+\tau_2\over\tau_1-\tau_2}\tanh{\eta\over 2},\nonumber\\
1115: |\eta|<\eta_0 =\ln{\tau_1\over\tau_2}
1116: \approx{\tau_1-\tau_2\over \sqrt{\tau_1\tau_2}}=\xi , ~~~
1117: \tanh\psi(\pm\eta_0)=\pm 1,~~~\psi(\pm\eta_0)=\pm \infty~,
1118: \label{eq:E6.08}
1119: \end{eqnarray}
1120:
1121: where $\xi=(\tau_1-\tau_2)/\sqrt{\tau_1\tau_2}\approx\eta_0$ is the main
1122: parameter of our calculations. This parameter is supposed to be small in
1123: order that the notion of the current transverse mass $\mu(p_t,\tau_1)$ has
1124: the expected meaning of a slowly varying parameter. The geometric mean time
1125: $\tau_m= \sqrt{\tau_1\tau_2}$ has a simple interpretation. The two
1126: characteristics, one connecting the points $(\tau_2,-\eta_0)$ and
1127: $(\tau_1,\eta_0)$, and the second one connecting the points
1128: $(\tau_2,\eta_0)$ and $(\tau_1,-\eta_0)$, intersect at the point
1129: $(\tau_m,0)$. The proper time $\tau_m$ is always inside the domain of the
1130: ``causal interaction.''
1131:
1132: Let us start the analysis of the radiation-dominated terms with the invariant
1133: $[\Sigma^{0}]^{(TE)}_{mat}$. (The invariant $[\Sigma^{T}]^{(TE)}_{mat}$
1134: appears to have an extra small factor $\xi$.) According to
1135: Eqs.~(\ref{eq:E6.04}), (\ref{eq:E6.05}), and (\ref{eq:E6.07}) it can be
1136: written as a multiple integral,
1137: \begin{eqnarray}
1138: \bigg[\Sigma^{0}(\tau_1,\tau_2)\bigg]^{(TE)}_{mat}=
1139: {i\alpha_sC_F\over 8\pi^2}
1140: \int_{-\infty}^{+\infty} d\eta \theta(\tau_{12}^2)
1141: \bigg\{{{\cal N}_f\over \pi R_\bot^2}\int_{p_\ast}^{\infty} d q_t
1142: \int_{0}^{2\pi}d\varphi ~J_0(\tau_{12}k_t)
1143: \int_{-\infty}^{\infty}d \theta~\cosh \theta
1144: \sin T_{12}(\theta)q_t \nonumber\\
1145: +{{\cal N}_g\over \pi R_\bot^2}
1146: {(\tau_1-\tau_2)\cosh\eta/2\over\tau_{12}}
1147: \int_{p_\ast}^{\infty} {d k_t\over k_t}\int_{0}^{2\pi}d\varphi
1148: ~q_t J_1(\tau_{12}q_t)\int_{-\infty}^{\infty}d\alpha~
1149: \cos T_{12}(\alpha)k_t ~\bigg\}\theta(\tau_1-\tau_2)~,
1150: \label{eq:E6.09}
1151: \end{eqnarray}
1152: where we choose the integral form of the densities $g_{[1]}$ and ${\cal
1153: D}^{(TE)}_{[1]}$ in order to find the domain in the multidimensional space
1154: where the dominant contribution comes from. Since the two terms in
1155: (\ref{eq:E6.09}) are not expected to interfere (or UV- diverge), we are free
1156: to change variables in these terms independently. We leave $d^2\vec{k_t}$ in
1157: the second term, and change for $d^2\vec{q_t}$ in the first one. The next
1158: step is to integrate over the azimuthal angle between $\vec{q_t}$ and
1159: $\vec{p_t}$ in the first term of Eq.~(\ref{eq:E5.3}), and over the angle
1160: between $\vec{k_t}$ and $\vec{p_t}$ in the second term. This integration
1161: deals only with the retarded propagators inside the self-energy loop and
1162: selects the lowest angular harmonics,
1163: \begin{eqnarray}
1164: \int_{0}^{2\pi}q_t J_1(\tau_{12}q_t)d\varphi=
1165: 2\pi~[ k_t J_0(\tau_{12}p_t)J_1(\tau_{12}k_t)+
1166: p_t J_1(\tau_{12}p_t)J_0(\tau_{12}k_t)]~,~~~k_t>p_t~,
1167: \label{eq:E6.10a}
1168: \end{eqnarray}
1169: \begin{eqnarray}
1170: \int_{0}^{2\pi}~ J_0(\tau_{12}k_t)d\varphi=
1171: 2\pi~ J_0(\tau_{12}p_t)J_0(\tau_{12}q_t)~,~~~q_t>p_t~.
1172: \label{eq:E6.10b}
1173: \end{eqnarray}
1174: Only the first of the two terms in Eq.~(\ref{eq:E6.10a}), corresponding to
1175: the collinear geometry in the transverse plane survives in the limit of
1176: $k_t\gg p_t$ and has to be retained by our major assumption. The second term
1177: describes the deviation from collinearity and is small. However, it is
1178: instructive to keep it for a while. After these angular integrations,
1179: Eq.~(\ref{eq:E6.09}) becomes
1180: \begin{eqnarray}
1181: \bigg[\Sigma^{0}(\tau_1,\tau_2)\bigg]^{(TE)}_{mat}=
1182: {i\alpha_sC_F\over 4\pi}\theta(\tau_1-\tau_2)
1183: \int_{-\infty}^{+\infty} d\eta \theta(\tau_{12}^2)
1184: \bigg\{{{\cal N}_f\over \pi R_\bot^2}\int_{p_\ast}^{\infty} d q_t
1185: J_0(\tau_{12}p_t)J_0(\tau_{12}q_t)
1186: \int_{-\infty}^{\infty}d \theta~\cosh \theta
1187: \sin T_{12}(\theta)q_t \nonumber\\
1188: +{{\cal N}_g\over \pi R_\bot^2}
1189: {(\tau_1-\tau_2)\cosh\eta/2\over\tau_{12}}
1190: \int_{p_\ast}^{\infty} {d k_t\over k_t}
1191: [ k_t J_0(\tau_{12}p_t)J_1(\tau_{12}k_t)
1192: +p_t J_1(\tau_{12}p_t)J_0(\tau_{12}k_t)]
1193: \int_{-\infty}^{\infty}d\alpha
1194: \cos T_{12}(\alpha)k_t ~\bigg\},
1195: \label{eq:E6.11}
1196: \end{eqnarray}
1197: where the actual limits of integration over $\eta$, $\theta$, and $\alpha$
1198: have yet to be put in agreement with the model we employ. Now, we have
1199: approached the most subtle point of our analysis. This expression includes
1200: triple integrations, any of which (if performed formally) yields singular
1201: functions. For the sake of definiteness, let us start with the second
1202: term in Eq.(\ref{eq:E6.11}) (which corresponds to the forward scattering of
1203: soft quark on a hard gluon from the distribution $n_g(\alpha,k_t)$),
1204: rewriting it in its most expanded form,
1205: \begin{eqnarray}
1206: {\cal J}_2={{\cal N}_g\over \pi R_\bot^2}\int d\eta \theta(\tau_{12}^2)
1207: {(\tau_1-\tau_2)\cosh\eta/2\over\tau_{12}}J_0[\tau_{12}(\eta)p_t]
1208: \int_{-\infty}^{\infty}d\alpha
1209: \int_{p_\ast}^{\infty}J_1[\tau_{12}(\eta)k_t]
1210: \cos [k_t\tau_{12}(\eta)\cosh(\alpha-\psi(\eta))]~d k_t~.
1211: \label{eq:E6.12}
1212: \end{eqnarray}
1213:
1214: The first observation is that at large $k_t$ (which is the condition that
1215: the distribution $n_g(\alpha,k_t)$ can be measured within a short time) the
1216: main contribution to the $\alpha$-integration comes from the domain
1217: $\alpha\approx\psi(\eta)$ where the phase of the $\cos T_{12}(\alpha)k_t$ is
1218: stationary. This is the domain of collinear interaction when the hard gluon
1219: from the distribution $n_g(\alpha,k_t)$ has almost the same rapidity as the
1220: virtual quark with transverse momentum $~\vec{q_t}=\vec{k_t}+\vec{p_t}~$ in
1221: the self-energy loop. Obviously, this quark is also hard. Furthermore, its
1222: propagator, $G_{[ret]}(\tau_1,\tau_2;q_t)= \theta
1223: (\tau_1-\tau_2)G_{[0]}(\tau_1,\tau_2;q_t)$, is devised only from the free
1224: on-mass-shell partial waves which themselves are well localized in the
1225: rapidity direction. Hence, we deal with the intuitively very clear case of
1226: collinear absorption and emission of the gauge field quantum by a charged
1227: particle. All participants of the process are moving with the same
1228: velocity. According to the property of localization of states in wedge
1229: dynamics studied in paper [II], such a fine tuning of $\alpha$ to $\psi$
1230: is indeed possible. This is illustrated by the left-hand figure on
1231: Fig.\ref{fig:fig2}. where the grey segments of the hyperbolas $\tau=\tau_2$
1232: and $\tau=\tau_1$ correspond to the rapidity intervals occupied by the soft
1233: quark mode, $\tau p_t<1$, at the beginning and at the end of the scattering
1234: process, respectively. The bold black and the dashed segments show the
1235: rapidity intervals where the hard virtual quark and the hard gluon are
1236: localized at the same times. All three fields effectively overlap around
1237: rapidity $\eta_2=-\eta/2$ at $\tau=\tau_2$ and around $\eta_1=+\eta/2$ at
1238: $\tau=\tau_1$. The rapidity direction between these points is exactly
1239: $\psi(\eta)$. The rapidity $\alpha$ of the external gluon is sufficiently
1240: small and is close to the rapidity $\psi(\eta)$.
1241:
1242: \begin{figure}[htb]
1243: \begin{center}
1244: \mbox{
1245: \psfig{file=./col1.ps,height=2.2in,bb=105 535 385 715}
1246: \hspace{.2cm}
1247: \psfig{file=./col2.ps,height=2.2in,bb=105 535 385 715}
1248: }
1249: \end{center}
1250: \caption{Geometry of fields in the forward scattering amplitude $qg\to qg$.}
1251: \label{fig:fig2}
1252: \end{figure}
1253:
1254: The maximal rapidity width of the interaction domain is defined by the
1255: causality condition $\tau_{12}^2 >0$, which immediately establishes the
1256: upper boundary $|\eta|<\eta_0$. Since the collinear interaction
1257: corresponds to the condition $\alpha\approx\psi(\eta)$, the rapidity of the
1258: hard gluon must be within this geometrically defined interval as well. The
1259: opposite case is depicted in the right-hand figure of Fig.\ref{fig:fig2}.
1260: The rapidity $\psi(\eta)$ is so large, that the external gluon is not
1261: localized within the causal boundaries $\pm\eta_0/2$ of the interaction
1262: domain. In order to avoid this, we have to impose an even stronger
1263: requirement that $|\psi(\eta)|<\eta_0$. According to (\ref{eq:E6.08}), we
1264: have $|\psi(\eta)|>\eta$. Hence, we must further take $|\eta|<\eta_\ast$,
1265: where the boundary $\eta_\ast$ is defined by the equation
1266: $\psi(\eta_\ast)=\eta_0$,
1267: \begin{eqnarray}
1268: {\tau_1+\tau_2\over\tau_1-\tau_2}\tanh{\eta_\ast\over 2}=\tanh\eta_0
1269: \equiv {\tau_1^2-\tau_2^2\over\tau_1^2+\tau_2^2}~,
1270: \label{eq:E6.13}
1271: \end{eqnarray}
1272: which has a solution,
1273: \begin{eqnarray}
1274: \tanh{\eta_\ast\over 2}
1275: = {(\tau_1-\tau_2)^2\over\tau_1^2+\tau_2^2}~,~~~~~~
1276: \eta_\ast\approx{(\tau_1-\tau_2)^2\over \tau_1\tau_2}~=\xi^2~.
1277: \label{eq:E6.14}
1278: \end{eqnarray}
1279: We remind the reader that $\xi\ll 1$; only this condition allows one to
1280: introduce the the time-dependent transverse mass $\mu(p_t,\tau)$.
1281: In order to simplify further analysis, it is convenient to present the
1282: internal integral over $k_t$ as the difference,
1283: \begin{eqnarray}
1284: -~{\tau_{12}\over (T^2_{12}(\alpha)-\tau_{12}^2)_+^{1/2}
1285: [T_{12}(\alpha)+(T^2_{12}(\alpha)-\tau_{12}^2)^{1/2}]}~
1286: -\int_{0}^{p_\ast} J_1(\tau_{12}k_t)
1287: \cos [T_{12}(\alpha)k_t]~d k_t ~,
1288: \label{eq:E6.15}
1289: \end{eqnarray}
1290: where the first singular term is the integral over $k_t$, computed from $0$
1291: to $\infty$, and thus, it completely accounts for the domain
1292: $k_t\to\infty$. It includes the function
1293: \begin{eqnarray}
1294: f(\eta,\alpha) =[T^2_{12}(\alpha)-\tau_{12}^2]_+^{-1/2}=
1295: [\tau_{12}^2(\eta)\sinh^2(\alpha-\psi(\eta))]_+^{-1/2}~, \nonumber
1296: \end{eqnarray}
1297: which is singular at $\alpha = \psi(\eta)$, thus fully accounting for the
1298: expected collinear enhancement. This function, however,
1299: is a canonical distribution with respect to both its arguments $\alpha$
1300: and $\eta$, and it carries the standard regulators for the subsequent
1301: integrations. We shall consider the singular and the regular terms
1302: separately. Using the above found limits, we may write the singular term as
1303: \begin{eqnarray}
1304: I_{2}^{sing}=\int_{-\eta_\ast}^{\eta_\ast} d\eta
1305: {(\tau_1-\tau_2)\cosh\eta/2\over\tau_{12}(\eta)}
1306: {1\over [\tau_{12}(\eta)]_+ }~
1307: \int_{-\eta_0}^{\eta_0}d\alpha~
1308: {e^{-|\alpha-\psi |}\over [\sinh^2|\alpha -\psi |]_{+}^{1/2}}~,
1309: \label{eq:E6.16}
1310: \end{eqnarray}
1311: where, since $\tau_{12}p_t\ll 1$, we put $J_0[\tau_{12}p_t]\approx 1$. After
1312: an obvious change of variable, the internal integral of Eq.~(\ref{eq:E6.16})
1313: can be split into two,
1314: \begin{eqnarray}
1315: \int_{-\eta_0}^{\eta_0} d\alpha~
1316: {e^{-|\alpha-\psi |}\over [\sinh^2|\alpha -\psi |]_{+}^{1/2}}=
1317: \bigg[ \int_{0}^{\eta_0+\psi}+\int_{0}^{\eta_0-\psi}\bigg]
1318: {\alpha~e^{-\alpha}\over \sinh\alpha }~{d\alpha\over \alpha_{+}}~.\nonumber
1319: \end{eqnarray}
1320: Since by the definition of the (+)-distribution,
1321: \begin{eqnarray}
1322: \int_{0}^{\beta} {\alpha~e^{-\alpha}\over \sinh\alpha }
1323: ~{d\alpha\over \alpha_{+}}=
1324: \int_{0}^{\beta}{e^{-\alpha }d\alpha\over\sinh\alpha}
1325: -\int_{\epsilon}^{1}{d\alpha\over\alpha}=
1326: \ln{1-e^{-2\beta}\over 2}~,
1327: \label{eq:E6.17}
1328: \end{eqnarray}
1329: we obtain the singular part in the form,
1330: \begin{eqnarray}
1331: I_2^{sing}=2~\int_{0}^{\eta_\ast}
1332: {(\tau_1-\tau_2)\cosh\eta/2\over\tau_{12}(\eta)}
1333: { d\eta \over [\tau_{12}(\eta)]_+ }~
1334: \ln\big\{{1\over 4}[1+e^{-4\eta_0}-2e^{-2\eta_0}\cosh 2\psi(\eta)]\big\}~.
1335: \label{eq:E6.18}
1336: \end{eqnarray}
1337: Next, it is convenient to trade $\eta$ for a new variable $y$,
1338: $\tau_{12}(\eta)=(\tau_1 -\tau_2)y$. The helpful relations for this change
1339: of variables are
1340: \begin{eqnarray}
1341: {\cosh(\eta/2)d\eta \over\tau_{12}(\eta)}={-1\over\sqrt{\tau_1\tau_2}}~
1342: {d y\over \sqrt{1-y^2}}~,~~~~
1343: \tau_{12}^2(\eta_\ast)=
1344: {(\tau_1 -\tau_2)^2\over 1+(\tau_1 -\tau_2)^2/\tau_1\tau_2}~,\nonumber\\
1345: y_\ast\approx 1-{(\tau_1 -\tau_2)^2\over 2\tau_1\tau_2}~,~~~~
1346: \cosh 2\psi={(\tau_1 +\tau_2)^2\over 2\tau_1\tau_2}{1\over y^2}-
1347: {\tau_1^2 +\tau_2^2\over 2\tau_1\tau_2}~.\nonumber
1348: \end{eqnarray}
1349: Taking into account that $\cosh 2\psi\to 1$, when $y\to 1$, we obtain,
1350: \begin{eqnarray}
1351: I_2^{sing}={4\over\sqrt{\tau_1\tau_2}}~\int_{y_\ast}^{1}
1352: {d y\over y\sqrt{1-y^2}}
1353: \ln{1-e^{-2\eta_0}\over 2} \approx {4\over\sqrt{\tau_1\tau_2}}
1354: ~{\tau_1-\tau_2\over\sqrt{\tau_1\tau_2}}
1355: ~\ln{\tau_1-\tau_2\over\sqrt{\tau_1\tau_2}}=
1356: {4\over\sqrt{\tau_1\tau_2}}~\xi\ln\xi~.
1357: \label{eq:E6.19}
1358: \end{eqnarray}
1359:
1360: This formula has two distinctive elements. The first element is the large
1361: $\ln\xi$, which is due to the collinear geometry of the interaction. This
1362: would lead to a divergence if the interaction domain were unlimited. The
1363: second element is the small factor $\xi$ which is due to the small volume
1364: occupied by the interaction and it completely suppresses the potential
1365: divergence. One may notice that when the mean time $\sqrt{\tau_1\tau_2}$
1366: increases, then $\xi\to 0$, and the corresponding part of the self-energy
1367: also tends to zero. This is easy to understand, since with the mean time
1368: growing, the system becomes more and more diluted locally.
1369:
1370: The regular part of Eq.~(\ref{eq:E6.12}) is given by the integral,
1371: \begin{eqnarray}
1372: I_2^{reg}=\int_{-\eta_\ast}^{\eta_\ast} d\eta
1373: {(\tau_1-\tau_2)\cosh\eta/2\over\tau_{12}(\eta)}
1374: \int_{-\eta_0}^{\eta_0}d\alpha~
1375: \int_{0}^{p_\ast} J_1(\tau_{12}k_t)
1376: \cos [T_{12}(\alpha)k_t]~d k_t ~,
1377: \label{eq:E6.20}
1378: \end{eqnarray}
1379: where, when $~-\eta_0<\eta<\eta_0$, then $T_{12}(\alpha)$ varies between
1380: its minimal and maximal values,
1381: \begin{eqnarray}
1382: {(\tau_1 -\tau_2)^2\over\sqrt{\tau_1\tau_2} }e^{-|\alpha|} <
1383: T_{12}(\alpha)<{(\tau_1 -\tau_2)^2\over
1384: \sqrt{\tau_1\tau_2} }e^{+|\alpha|}~. \nonumber
1385: \end{eqnarray}
1386: Therefore, when $k_t<p_t$,
1387: we have $T_{12}k_t\sim (\tau_1-\tau_2)k_t\ll 1$, and both functions under
1388: the integral over $k_t$ can be expanded in Taylor series. All integrations
1389: become trivial and yield \footnote{ Even if we impose no limitations on
1390: $\alpha$ the estimate is still as small as
1391: $${\tau_1\tau_2 p_\ast^2\over\sqrt{\tau_1\tau_2}}~\xi^2\ln\xi~.$$}
1392: \begin{eqnarray}
1393: I_2^{reg}= {\tau_1\tau_2 p_\ast^2\over\sqrt{\tau_1\tau_2}}~
1394: \bigg[{\tau_1-\tau_2\over\sqrt{\tau_1\tau_2}}\bigg]^4
1395: ={\tau_1\tau_2 p_\ast^2\over\sqrt{\tau_1\tau_2}}~\xi^4~.
1396: \label{eq:E6.21}
1397: \end{eqnarray}
1398: We have chosen this form of the answer, because our major assumption is
1399: valid only as long as $\tau p_t \leq 1$ and because the dispersion equation
1400: (\ref{eq:E3.12}) has a kinematic factor $\sqrt{\tau_1\tau_2}$ in it.
1401:
1402: The other two terms in Eq.~(\ref{eq:E6.11}) can be studied along the same
1403: guidelines. The third term is suppressed with respect to ${\cal J}_2$ by the
1404: factor $p_t/k_t$, which is small by our major model agreement and it could
1405: have been discarded on this ground only. To be on safe side, let us rewrite
1406: it as
1407: \begin{eqnarray}
1408: {\cal J}_3={{\cal N}_g\over \pi R_\bot^2}\int_{-\eta_\ast}^{\eta_\ast}
1409: d\eta ~{(\tau_1-\tau_2)\cosh\eta/2\over\tau_{12}}
1410: ~p_t~J_1[\tau_{12}(\eta)p_t]\nonumber\\
1411: \times\int_{-\eta_0}^{\eta_0} d\alpha ~
1412: \bigg\{ -\gamma_E -\ln\big[{\tau_{12}p_t\over 2}
1413: e^{|\alpha-\psi(\eta)|}\big]+
1414: \int_{0}^{p_\ast}{1- J_0[\tau_{12}(\eta)k_t]\cos (T_{12}(\alpha)k_t)
1415: \over k_t}~d k_t~\bigg\}~,
1416: \label{eq:E6.22}
1417: \end{eqnarray}
1418: where the first term is the integral over $k_t$ from $0$ to $\infty$. As i
1419: could be expected, the integrand is regular. Since there is an accounted for
1420: difference between $k_t$ and $q_t$, the exactly collinear regime becomes
1421: impossible and we do not have the large collinear logarithm in ${\cal J}_3$.
1422: Overall, this term is also suppressed at least by a factor $\xi$ stemming
1423: from $J_1[\tau_{12}(\eta)p_t]$ in the integrand.
1424:
1425: The first term in Eq.(\ref{eq:E6.11}) corresponds to the forward quark-quark
1426: scattering with high momentum transfer,
1427: \begin{eqnarray} {\cal J}_1={{\cal N}_f\over \pi R_\bot^2}
1428: \int_{-\eta_\ast}^{\eta_\ast} d\eta
1429: ~J_0[\tau_{12}(\eta)p_t] \hspace{10cm}\nonumber\\
1430: \times \int_{-\eta_0}^{\eta_0} d\theta ~
1431: \bigg\{ {\cosh\theta\over [\sinh^2(\theta-\psi(\eta))]_{+}^{1/2}}
1432: {1\over[\tau_{12}(\eta)]_+}
1433: -\cosh\theta~\int_{0}^{p_\ast} J_0[\tau_{12}(\eta)q_t]
1434: \sin (T_{12}(\theta)q_t)~d q_t~\bigg\}~.
1435: \label{eq:E6.23} \end{eqnarray}
1436: Here, we again recognize the collinear singularity which is, as previously,
1437: regulated by the (+)-prescription. All further calculations for
1438: ${\cal J}_1$ are similar to the case of ${\cal J}_2$ and the answer reads,
1439: \begin{eqnarray}
1440: I_1^{sing} \leq {-8\xi + 4 \xi^2 \ln\xi\over \sqrt{\tau_1\tau_2}}~,~~~~
1441: I_1^{reg}\approx 2{p_\ast^2 \tau_1\tau_2 \over \sqrt{\tau_1\tau_2}}\xi^5~.
1442: \label{eq:E6.24}
1443: \end{eqnarray}
1444: These results will serve for us as a reference point for the estimates of
1445: the mathematically more complicated part connected with the radiation field
1446: of the transverse magnetic TM-modes. Before we address this issue, it is
1447: expedient to look at the obtained results more attentively and trace the
1448: correspondence between the calculations and physical picture in more
1449: details:
1450:
1451: {\em ~~1.} It has been observed in Sec.~\ref{sec:S5} (for the vacuum part of
1452: the quark self-energy) that in the framework of wedge dynamics, the
1453: collinear problems do not jeopardize the field theory. In ``material part''
1454: of the self-energy, the collinear interactions were proved to be the most
1455: intensive and to lead to a visible enhancement of the interaction between
1456: the quark and {\em radiation} field. However, this enhancement never turns
1457: into a disaster of collinear divergence. One of the trivial reasons is that
1458: the space-time domain of the interaction is now limited, and large
1459: logarithms are multiplied by small phase volumes.
1460:
1461: {\em ~~2.} A deeper insight into the wedge dynamics, shows that even
1462: intermediate collinear singularities observed in the terms ${\cal J}_1$
1463: and ${\cal J}_2$ are spurious. In order to reveal this fact, one can notice
1464: that the singularity at $\alpha = \psi(\eta)$ is present only in ${\cal
1465: J}_1$ and ${\cal J}_2$. It is absent in ${\cal J}_3$, because of the extra
1466: negative power of $k_t$ brought by the subleading term of the angular
1467: integral $d\varphi$. This extra $k_{t}^{-1}$ effectively suppresses the
1468: distribution $n_g(\alpha,k_t)$ at large $k_t$. Next, one may ask, what
1469: minimal change of $n_g(\alpha,k_t)$ at large $k_t$ is necessary in order
1470: that the intermediate collinear singularity does not appear at all. This can
1471: be learned by changing the order of integration in Eq.~(\ref{eq:E6.12}).
1472: One can start from the integral $d\eta$ with an assumption that the
1473: integrand only slowly varies within some interval of $\alpha$ around $\alpha
1474: =0$. Then, it is easy to see that the singular term $I_2^{sing}$ of
1475: Eq.~(\ref{eq:E6.19}) {\em totally} originates from the domain of
1476: $k_t\to\infty$. It comes from a logarithmic integral between two infinite
1477: limits. This residual piece emerges only because we extend the distribution
1478: $n_g\sim k_t^{-2}$ (obtained from a dimensional estimate) to an arbitrarily
1479: large $k_t$. As we have already mentioned, this dependence is unphysical,
1480: e.g., because the distribution $d^2\vec{k_t}/k_t^2$ is not normalizeable. It
1481: has to be modified above some value of $k_t$ and, therefore, the singular
1482: term must vanish completely. Thus, in wedge dynamics, the phenomenon of
1483: collinear enhancement is intrinsically connected with the basic property of
1484: localization inherent in the one-particle states. Only the states with
1485: infinitely large $k_t$ can have a precisely given rapidity and be
1486: responsible for the singularities like we encounter in Eqs.~(\ref{eq:E6.15})
1487: and (\ref{eq:E6.23}).
1488:
1489: {\em ~~3.} Our way to pick out the leading contributions (in the mixed
1490: representation of wedge dynamics) from the space-time domains, where the
1491: phases of the interacting fields are stationary, is a generalization of the
1492: known method of isolating the leading terms using the pinch-poles in the
1493: plane of complex energy. The similarity of two methods can be easily
1494: understood since, e.g., the quark density correlator in the self-energy can
1495: be presented as a sum of two propagators,
1496: $G_{[1]}(q)=G_{[00]}(q)+G_{[11]}(q)$. In the plane of the complex energy
1497: $q^0=k^0+p^0$, the (Feynman-type) propagator $G_{[00]}(q)$ has poles in the
1498: second and fourth quadrants, while the (anti-Feynman-type) propagator
1499: $G_{[11]}(q)$ has poles in the first and third quadrants. The radiation
1500: part of the retarded gauge field propagator $D_{[ret]}(k)$ has poles in the
1501: third and fourth quadrants. Therefore, in both terms of $G_{[1]}(q)
1502: D_{[ret]}(k) =G_{[00]}(q)D_{[ret]}(k) +G_{[11]}(q)D_{[ret]}(k)$, the
1503: integration path along the real axis of the complex $k^0$ plane is pinched
1504: between two poles (one from $D_{[ret]}$, and the second from $G_{[00]}$ or
1505: $G_{[11]}$ ) giving the leading contribution when $p^0$ is small, and the
1506: three-momenta $\bbox{k}$ and $\bbox{q=k+p}$ are are very close to each
1507: other. Similar arguments are valid for the second part, $G_{[ret]}D_{[1]}$,
1508: of the quark self-energy. The term $G_{[1]}(q)D_{[long]}(k)$ is exceptional,
1509: because the propagator $D_{[long]}(k)$ of the longitudinal field has no
1510: poles corresponding to the propagation.
1511:
1512: The wedge dynamics does not allow for a standard momentum representation,
1513: since its geometric background is not homogeneous in the $t$- and
1514: $z$-directions; accordingly, we do not have familiar pinch-poles in our
1515: calculations. Nevertheless, the patches of phase space where the phases of
1516: certain field fragments are stationary and effectively overlap, do now the
1517: same job as the pinch-poles, and yield the same answers when the homogeneity
1518: required for the momentum representation is restored. The way wedge dynamics
1519: tackles the problem is genuinely more general, because it addresses the
1520: space-time picture of the interacting fields.\footnote{It is well known,
1521: that the threshold behavior of the imaginary part of the photon self-energy
1522: can be derived from the pinch geometry of the poles of the electron
1523: propagators \cite{Berest}. Since at the threshold, the $e^+ e^-$ pair is
1524: created with zero relative velocity, the pinch indeed corresponds to the
1525: overlap of the stationary phases of the $e^+$ and $ e^-$ wave functions in
1526: the maximal possible volume.} The momentum space is now split into the
1527: subspaces of rapidity and transverse momentum; the correlation between the
1528: particle's rapidity and its location is increasing with the increase of its
1529: transverse momentum . The role of pinch-poles is taken over by the
1530: geometrical overlap of the field patterns with the same rapidity. This
1531: observation can serve as a footing for the future development of an
1532: effective technique for perturbative calculations in wedge dynamics. The
1533: arguments of the localization are not applicable to the longitudinal part
1534: of the gluon field. In the term $G_{[1]}D_{[long]}$, no patch in space-time
1535: is dynamically selected, since $D_{[long]}$ is not assembled from the
1536: propagating waves that could match the virtual quark in the loop by their
1537: phase. This is in line with the absence of pinch-poles due to
1538: $D_{[long]}(k)$ in the momentum picture.
1539:
1540: The arguments presented above allow one to estimate the contribution of the
1541: radiation fields of the TM-mode in a very economical way. Let us consider
1542: the group ${\cal D}_{[0]}^{(2)}g^0_{[1]}$, which is very similar to the term
1543: ${\cal J}_{1}$ studied above, as an example. Now, the invariant ${\cal
1544: D}_{[0]}^{(2)}$, as can be inferred from Eq.~(\ref{eq:A1.4}), is known only
1545: in the integral representation, and not in an analytic form. Let us,
1546: therefore, employ the analytic form of $g^0_{[1]}$, given by
1547: Eq.~(\ref{eq:A1.20}). It is easy to see, that the integration over $\alpha$
1548: in ${\cal D}_{[0]}^{(2)}$ leads to the same causal step-function
1549: $\theta(\tau_{12}^2)$ and, as previously shown, we have $|\eta|<\eta_0$.
1550: The expanded form of this term is
1551: \begin{eqnarray}
1552: {\cal J}_4={{\cal N}_f\over \pi R_\bot^2}\int_{-\eta_0}^{\eta_0}
1553: d\eta \theta(\tau_{12}^2)
1554: {(\tau_1-\tau_2)\cosh\eta/2\over\tau_{12}}\int d\alpha
1555: \tanh\big(\alpha-{\eta\over 2}\big) \tanh\big(\alpha+{\eta\over 2}\big)
1556: \int_{p_\ast}^{\infty} Y_1[\tau_{12}(\eta)q_t]
1557: \sin [ T_{12}(\alpha)q_t]dq_t ,
1558: \label{eq:E6.25}
1559: \end{eqnarray}
1560: where we have integrated out the azimuthal angle $\varphi$ assuming that
1561: $k_t,q_t\gg p_t$ (the first correction is smaller by the factor
1562: $(p_t/k_t)^2\ll 1$). The internal integral over $q_t$ can be transformed into
1563: \begin{eqnarray}
1564: -~{\cosh (\alpha-\psi(\eta))]\over
1565: [\tau_{12}^2(\eta)\sinh^2(\alpha-\psi(\eta))]_+^{1/2}}~
1566: -~\int_{0}^{p_\ast} Y_1(\tau_{12}q_t) \sin [T_{12}(\alpha)k_t]~d k_t ~,
1567: \label{eq:E6.26}
1568: \end{eqnarray}
1569: which brings us very close to Eq.~(\ref{eq:E6.23}) for ${\cal J}_1$. Once
1570: again, we encounter a collinear singularity at $\alpha=\psi(\eta)$, and
1571: exactly the same arguments force us to set the same limits in the integrals,
1572: as in Eq.~(\ref{eq:E6.23}). We do not have to continue the calculations to
1573: understand the smallness of ${\cal J}_4$, mention only, that due to the
1574: narrow limits of two rapidity integrations in Eq.~(\ref{eq:E6.25}), the
1575: product of the two hyperbolic tangents in the integrand will add extra
1576: $\xi^2$ to the order of smallness of ${\cal J}_4$. By the same argument as
1577: used previously, we can easily learn that the singular term in
1578: Eq.~(\ref{eq:E6.26}) is spurious.
1579:
1580: In this group, associated with the $TM$-mode of the radiation field, the
1581: leading (still parametrically small, and equal to ${\cal J}_1$)
1582: contribution comes from the $D^{\eta\eta}$ component of the gluon
1583: correlators. We may summarize by the observation that only the overlap of the
1584: domains of stationary phase in two correlators matters. It can be visualized
1585: via the partial-wave expansion of any of the correlators in the self-energy
1586: loop.
1587:
1588:
1589:
1590: \section{ Nonlocal part of the longitudinal propagator
1591: in the material part of the self-energy.}
1592: \label{sec:SA2}
1593:
1594: The longitudinal part of the gluon propagator contributes, to the invariant
1595: $\Sigma^{0}$, the term
1596: \begin{eqnarray}
1597: \bigg[\Sigma^{0}(\tau_1,\tau_2)\bigg]^{[long]}_{mat}=
1598: {i\alpha_s C_F\over 4\pi}~\theta(\tau_1-\tau_2)\int d^2 \vec{k_t}
1599: \int_{-\infty}^{\infty}d\eta~ q_t ~g^{0}_{[1]}
1600: \bigg[{\cal D}^{(2)}_{[long]}
1601: +{1\over \tau_1\tau_2}{\cal D}^{(\eta\eta)}_{[long]}\bigg]~.
1602: \label{eq:A2.01}\end{eqnarray}
1603: Using Eqs.~(\ref{eq:A1.14a}) and (\ref{eq:A1.15a}) for the gluon invariants,
1604: and the first of Eqs.~(\ref{eq:A1.20}) for the quark invariant $g^{0}_{[1]}$,
1605: we arrive at the following expression which accounts for the non-local part
1606: of the longitudinal propagator (the contact part was studied in
1607: Sec.~\ref{sec:S3a}).
1608: \begin{eqnarray}
1609: \bigg[\Sigma^{0}(\tau_1,\tau_2)\bigg]^{[long]}_{mat}=
1610: {i\alpha_s C_F{\cal N}_f\over 2\pi^2R_\bot^2}~\theta(\tau_1-\tau_2)~
1611: \int_{p_\ast}^{\infty} d q_t
1612: \int_{-\infty}^{\infty}d\eta~
1613: {(\tau_1-\tau_2)\cosh\eta/2\over |\tau_{12}^2|^{1/2}} \nonumber\\
1614: \times {k_t\cosh \eta\over 2} \int_{\tau_2}^{\tau_1} e^{-t q_t\sinh |\eta|}
1615: \bigg(1-{t^2\over\tau_1\tau_2}\bigg) ~ dt
1616: \bigg[\theta(\tau_{12}^2)~Y_1(\tau_{12}q_t)
1617: +{2\over\pi}\theta(-\tau_{12}^2)~
1618: K_1(\tilde{\tau}_{12}q_t)\bigg] ,
1619: \label{eq:A2.02}\end{eqnarray}
1620:
1621: where we have integrated out the dependence on the azimuthal angle
1622: $\varphi$ in the approximation of $q_t\gg p_t$. The key observation that
1623: allows one to judge about the smallness of this term is that the limits of
1624: integration over variable $t$ are very close, and the factor
1625: $[1-t^2/\tau_1\tau_2]$ is very small. This factor reflects a known
1626: competition between the electric and magnetic interaction of moving charges
1627: which reduces the net yield almost to zero. Let us replace $t$ by the
1628: dimensionless $u=t/\sqrt{\tau_1\tau_2}$. Then this part of the integral
1629: becomes
1630: \begin{eqnarray}
1631: {1\over\sqrt{\tau_1\tau_2}}
1632: \int_{\tau_2}^{\tau_1}
1633: \bigg[1-{t^2\over\tau_1\tau_2}\bigg]\cdot\cdot\cdot dt
1634: = \int_{\sqrt{\tau_2/\tau_1}}^{\sqrt{\tau_1/\tau_2}}
1635: (1-u^2)\cdot\cdot\cdot du
1636: =\int_{\sqrt{1-\xi^2/4}-\xi/2}^{\sqrt{1-\xi^2/4}+\xi/2}
1637: (1-u^2)\cdot\cdot\cdot du \approx -{\xi^3\over 3}f(\xi)~,
1638: \label{eq:A2.03}\end{eqnarray}
1639: where the limiting behavior $f(\xi)\sim const/\xi$ when $\xi\to 0$ can be
1640: conjectured from the behavior of the functions $Y_1(x)$ and $K_1(x)$ at
1641: small $x$. However, the Laplace transforms of these functions in
1642: Eq.~(\ref{eq:A2.02}) are singular functions of $\eta$ and we have to be
1643: careful in estimating these terms. In fact, the $\xi^2$ order of smallness
1644: is not altered by the remaining integrations. First, it is useful to notice
1645: that the last factor in square brackets in Eq.~(\ref{eq:A2.02}) is nothing
1646: but the invariant $g^{0}_{[1]}$ which, according to its integral
1647: representation (\ref{eq:A1.20}), equals zero at $\xi=0$. Next, it is
1648: profitable to change the variables of integration in the following way. We
1649: trade $\eta$ for $y$ according to $\tau_{12}(\eta)=(\tau_1 -\tau_2)y$ in the
1650: domain $\tau_{12}^2>0$. In the complimentary domain
1651: $\tau_{12}^2=-\tilde{\tau}_{12}^2<0$, we change $\eta$ for $y$ using
1652: $\tilde{\tau}_{12}(\eta)=(\tau_1 -\tau_2)y$. We also replace $\xi q_t$ by a
1653: new variable $q$.
1654: \begin{eqnarray}
1655: \bigg[\Sigma^{0}(\tau_1,\tau_2)\bigg]^{[long]}_{mat}=
1656: {i\alpha_s C_F{\cal N}_f\over 2\pi^2R_\bot^2}~\theta(\tau_1-\tau_2)
1657: ~{1\over\xi}~\int_{\sqrt{1-\xi^2/4}-\xi/2}^{\sqrt{1-\xi^2/4}+\xi/2} (1-u^2)
1658: \int_{\xi p_\ast}^{\infty} q d q \nonumber \\
1659: \times\bigg\{\int_0^1 {dy\over\sqrt{1-y^2}}
1660: \bigg[1+{\xi^2\over 2}(1-y^2)\bigg]
1661: ~{\pi\over 2}~Y_1(\tau_m qy)
1662: ~e^{-\tau_m q u\sqrt{1-y^2}~\sqrt{1*\xi^2(1-y^2)/2}}\nonumber \\
1663: +\bigg(\int_0^1+\int_1^\infty\bigg)
1664: {dy\over\sqrt{1+y^2}}\bigg[1+{\xi^2\over 2}(1+y^2)\bigg]
1665: K_1(\tau_m qy)
1666: ~e^{-\tau_m q u\sqrt{1+y^2}~\sqrt{1*\xi^2(1+y^2)/2}}\bigg\}~.
1667: \label{eq:A2.04}\end{eqnarray}
1668: where, we remind the reader that $\tau_m=\sqrt{\tau_1\tau_2}$. When the
1669: argument is small, the functions $Y_1(x)$ and $K_1(x)$ are,
1670: $${\pi\over 2}~Y_1(x)\approx -{1\over x} +\ln{x\over 2}\bigg({x\over 2}
1671: +O(x^3)\bigg),~~~~K_1(x)\approx {1\over x} +\ln{x\over 2}\bigg({x\over 2}
1672: +O(x^3)\bigg)~. $$
1673: It is easy to see now that in the sum of the two integrals $dy$ over the
1674: interval $(0,1)$, the leading singularities $dy/y$ exactly cancel each
1675: other. Furthermore, it is safe to take the limits of $\xi\to 0$, $u\to 1$ in
1676: the integrand, and even to set the lower limit of the integral $dq$ to be
1677: zero. The resulting integral is convergent, $\xi$-independent, and yields a
1678: term of the order $\xi^2$. In the remaining integral, the variable $y$ runs
1679: from $1$ to $\infty$, the integrand is not singular at finite $y$ and $q$,
1680: and it is exponentially suppressed at large $y$ and $q$. The behavior of the
1681: integral at $\xi\to 0$ is not singular and the basic upper estimate
1682: $~const\times\xi^2$ remains unchanged for the entire integral
1683: (\ref{eq:A2.04}).
1684:
1685:
1686: \vspace{2cm}
1687:
1688:
1689: \noindent {\bf ACKNOWLEDGMENTS}
1690:
1691: The authors are grateful to Berndt Muller and Edward Shuryak for
1692: helpful discussions at various stages in the development of this work,
1693: and appreciate the help of Scott Payson who critically read the
1694: manuscript.
1695:
1696:
1697:
1698: \bigskip
1699:
1700:
1701:
1702:
1703:
1704: \renewcommand{\theequation}{A\arabic{equation}}
1705: \setcounter{equation}{0}
1706:
1707: \section{Appendix. Wightman functions and propagators of wedge dynamics.}
1708: \label{sec:SA1}
1709:
1710: In this Appendix, we put all field correlators into a form which is needed
1711: for the practical calculation of the self-energy. The density of states
1712: $D_{[1]}$ and the causal part $D_{[0]}$ of the gluon propagator are used in
1713: the form of decomposition over the transverse modes,
1714: \begin{eqnarray}
1715: D_{[10]}^{lm}(\tau_2,\tau_1;\eta_2-\eta_1;\vec{k_t})= -i(2\pi)^2
1716: \int d\alpha \sum_{\lambda} v^{(\lambda)l}_{\alpha,\vec{k_t}}(\tau_2,\eta_2)
1717: \overstar{v}^{(\lambda)m}_{\alpha,\vec{k_t}}(\tau_1,\eta_1)~,\nonumber\\
1718: D_{[01]}^{lm}(\tau_2,\tau_1;\eta_2-\eta_1;\vec{k_t})= -i(2\pi)^2
1719: \int d\alpha \sum_{\lambda} v^{(\lambda)l}_{\alpha,-\vec{k_t}}(\tau_2,\eta_2)
1720: \overstar{v}^{(\lambda)m}_{\alpha,-\vec{k_t}}(\tau_1,\eta_1)~,
1721: \label{eq:A1.1}
1722: \end{eqnarray}
1723: where
1724: \begin{eqnarray}
1725: v^{(TE)}_{{\vec k},\alpha}(x)={1\over 4\pi^{3/2} k_{t}}
1726: \left[ \begin{array}{c}
1727: k_y \\
1728: -k_x \\
1729: 0
1730: \end{array} \right]
1731: e^{-ik_{t}\tau\cosh (\alpha-\eta)};~~~
1732: v^{(TM)}_{{\vec k},\alpha}(x)={1\over 4\pi^{3/2} k_{t}}
1733: \left[ \begin{array}{c}
1734: k_x f_1 \\
1735: k_y f_1 \\
1736: - f_2
1737: \end{array} \right] ~~,
1738: \label{eq:A1.2}\end{eqnarray}
1739: are the transverse electric and transverse magnetic modes of the radiation
1740: field found previously in paper [II]. Here, we denoted,
1741: \begin{eqnarray}
1742: f_1(\tau ,\eta)= i \tanh (\alpha-\eta)
1743: (e^{-ik_{t}\tau \cosh (\alpha-\eta)} - 1)~~, \nonumber\\
1744: f_2(\tau ,\eta)= {e^{-ik_{t}\tau \cosh (\alpha-\eta)} -1 \over
1745: \cosh^{2} (\alpha-\eta)} +
1746: i k_{t}\tau {e^{-ik_{t}\tau\cosh (\alpha-\eta)}
1747: \over \cosh (\alpha-\eta)}~~.
1748: \label{eq:A1.3}
1749: \end{eqnarray}
1750: Starting from this form, we get the components of the commutator
1751: $D_{[0]}(\tau_2,\tau_1;\eta_2-\eta_1;\vec{k_t})$, \footnote{In all formulae
1752: below, the gluon rapidity $\alpha$ is counted from the reference point
1753: $(\eta_1+\eta_2)/2$, the geometric center of the coordinates $\eta_1$ and
1754: $\eta_2$ of the vertices in the self-energy loop. Thus, it corresponds to the
1755: rapidity $\theta'$ in the integral representation of the quark correlators
1756: in paper [II].}
1757: \begin{eqnarray}
1758: D_{[0]}^{rs}=
1759: \int {d\alpha\over 2\pi} \bigg\{\bigg[\delta_{rs}-{k_rk_s\over k_t^2}\bigg]
1760: \sin k_tT_{12}
1761: +{k_rk_s\over k_t^2}\tanh\big(\alpha+{\eta\over 2}\big)
1762: \tanh\big(\alpha-{\eta\over 2}\big)
1763: \big[\sin k_tT_{12}\underline{-\sin k_tT_1 +\sin k_tT_2}~\big]\bigg\}~,
1764: \label{eq:A1.4}
1765: \end{eqnarray}
1766: \begin{eqnarray}
1767: D_{[0]}^{\eta\eta}=
1768: \int {d\alpha\over 2\pi}
1769: {1\over k_t^2\cosh^2(\alpha+{\eta\over 2}) \cosh^2(\alpha-{\eta\over 2})}
1770: \big[(1+k_t^2T_1T_2)\sin k_tT_{12}
1771: -k_tT_{12}\cos k_tT_{12}\nonumber\\
1772: \underline{+\sin k_tT_2-\sin k_tT_1-k_tT_2\cos k_tT_2
1773: +k_tT_1\cos k_tT_1}~\big]~,
1774: \label{eq:A1.5}
1775: \end{eqnarray}
1776: where $ T_1=\tau_1\cosh (\alpha-\eta/2)$, $ T_2=\tau_2\cosh (\alpha+\eta/2)$,
1777: $T_{12}=T_1-T_2$. In the first of these equations, the coefficients of the
1778: tensors $(\delta_{rs}-k_rk_s/k_t^2)$ and $k_rk_s/k_t^2$ are the invariants
1779: ${\cal D}^{(TE)}_{[0]}$ and ${\cal D}^{(2)}_{[0]}$ of Eq.~(\ref{eq:E3.8}),
1780: respectively. The latter is due to the $TM$-mode of the radiation field. Up to
1781: the factor $k_t^{-2}$, Eq.~(\ref{eq:A1.5}) defines the invariant ${\cal
1782: D}^{(\eta\eta)}_{[0]}$. The underlined terms are connected with the boundary
1783: conditions imposed on the $TM$-mode at $\tau=0$. They cancel with the
1784: underlined terms in the longitudinal part of the gauge field propagator
1785: given by Eqs.~(\ref{eq:A1.14}) and (\ref{eq:A1.15}) thus
1786: providing the causal behavior of the components $D_{[adv]}^{rs}$ and
1787: $D_{[adv]}^{\eta\eta}$ of the retarded propagator $D_{[adv]}(\tau_2,\tau_1)$.
1788: In the body of the paper, we call the residues of this cancelation as
1789: the diagonal components of $D^{[0]}$.
1790: The ``off-diagonal'' components of the commutator $D_{[0]}^{ij}$ are
1791: \begin{eqnarray}
1792: D_{[0]}^{r\eta}= {-ik_r\over k_t^2}
1793: \int {d\alpha\over 2\pi}
1794: {\tanh(\alpha+{\eta\over 2})\over\cosh^2(\alpha-{\eta\over 2})}
1795: \big[\sin k_tT_{12}- k_tT_1\cos k_tT_{12}
1796: +\sin k_tT_2-\sin k_tT_1+k_tT_1\cos k_tT_1\big]~,
1797: \label{eq:A1.6}
1798: \end{eqnarray}
1799: \begin{eqnarray}
1800: D_{[0]}^{\eta r}= {ik_r\over k_t^2}
1801: \int {d\alpha\over 2\pi}
1802: {\tanh(\alpha-{\eta\over 2})\over\cosh^2(\alpha+{\eta\over 2})}
1803: \big[\sin k_tT_{12}+ k_tT_2\cos k_tT_{12}
1804: +\sin k_tT_2-\sin k_tT_1-k_tT_2\cos k_tT_2\big]~.
1805: \label{eq:A1.7}
1806: \end{eqnarray}
1807: The commutator is not a symmetric tensor. However, by examination, these
1808: components are odd with respect to the rapidity difference
1809: $\eta=\eta_1-\eta_2$, and hence they do not contribute to the effective quark
1810: mass we are computing in this paper.
1811:
1812: The tensor of the gluon density $D_{[1]}^{ij}(\tau_2,\tau_1; \eta_2-\eta_1;
1813: \vec{k_t})$ has the ``diagonal'' components,
1814: \begin{eqnarray}
1815: D_{[1]}^{rs}=
1816: -i \int {d\alpha\over 2\pi} \bigg\{\bigg[\delta_{rs}-{k_rk_s\over k_t^2}\bigg]
1817: \cos k_tT_{12} \hspace{8cm}\nonumber \\
1818: +{k_rk_s\over k_t^2}\tanh(\alpha+{\eta\over 2}) \tanh(\alpha-{\eta\over 2})
1819: (\cos k_tT_{12}-\cos k_tT_1 -\cos k_tT_2 +1)\bigg\}~,
1820: \label{eq:A1.8}
1821: \end{eqnarray}
1822: \begin{eqnarray}
1823: D_{[1]}^{\eta\eta}=
1824: -i\int {d\alpha\over 2\pi}
1825: {1\over k_t^2\cosh^2(\alpha+{\eta\over 2}) \cosh^2(\alpha-{\eta\over 2})}
1826: \big[(1+k_t^2T_1T_2)\cos k_tT_{12}
1827: +k_tT_{12}\sin k_tT_{12}\nonumber\\
1828: -\cos k_tT_2-\cos k_tT_1-k_tT_2\sin k_tT_2 -k_tT_1\sin k_tT_1 +1\big]~,
1829: \label{eq:A1.9}
1830: \end{eqnarray}
1831: from which one can infer the invariants ${\cal D}^{(TE)}_{[1]}$ and
1832: ${\cal D}^{(2)}_{[1]}$ of Eq.~(\ref{eq:E3.8}) exactly in the same way as it
1833: was done for the invariants of the commutator $D_{[0]}$. The off-diagonal
1834: components,
1835: \begin{eqnarray}
1836: D_{[1]}^{r\eta}= {-k_r\over k_t^2}
1837: \int {d\alpha\over 2\pi}
1838: {\tanh(\alpha+{\eta\over 2})\over\cosh^2(\alpha-{\eta\over 2})}
1839: \big[\cos k_tT_{12}- k_tT_1\sin k_tT_{12}
1840: -\cos k_tT_2-\cos k_tT_1-k_tT_1\sin k_tT_1 +1\big]~,
1841: \label{eq:A1.10}
1842: \end{eqnarray}
1843: \begin{eqnarray}
1844: D_{[1]}^{\eta r}= {k_r\over k_t^2}
1845: \int {d\alpha\over 2\pi}
1846: {\tanh(\alpha-{\eta\over 2})\over\cosh^2(\alpha+{\eta\over 2})}
1847: \big[\cos k_tT_{12}- k_tT_2\sin k_tT_{12}
1848: -\cos k_tT_2-\cos k_tT_1-k_tT_2\sin k_tT_2 +1\big]~,
1849: \label{eq:A1.11}
1850: \end{eqnarray}
1851: are also non-symmetric and odd with respect to the rapidity difference
1852: $\eta$. They also do not contribute to the effective quark mass. Equations
1853: (\ref{eq:A1.8})-- (\ref{eq:A1.11}) give the components of the vacuum density
1854: of states of the gauge field in the wedge dynamics. In order to incorporate
1855: the ``material'' part given by the distribution of real gluons, the integrand
1856: of each of Eqs.~(\ref{eq:A1.8})-- (\ref{eq:A1.11}) must be multiplied by the
1857: common factor $[1+2n_g(k_t,\alpha)]$.
1858:
1859: The full tensor of the longitudinal part of the propagator that defines
1860: the field $A(\tau_1)$ via the current $j(\tau_2)$ at all preceding times,
1861: \begin{eqnarray}
1862: A^{[long]}_l(\tau_1)=\int_{0}^{\tau_1}\tau_2~d\tau_2~d\eta_2
1863: D^{[long]}_{lm}(\tau_2, \tau_1;\eta_1-\eta_2,{\vec k_t} ) j^m(\tau_2)~,
1864: \label{eq:A1.12}
1865: \end{eqnarray}
1866: was found in paper [III] in the following form,
1867: \begin{eqnarray}
1868: D^{[long]}_{lm}(\tau_2, \tau_1;\eta_1-\eta_2, {\vec k_t})
1869: =\hspace{12cm}\\ \label{eq:A1.13}
1870: = \int {d\nu d^2{\vec k} \over (2\pi)^3 k_{\bot}^2}
1871: \left[ \begin{array}{cc}
1872: k_rk_s [Q_{-1,i\nu}(k_{\bot}\tau_2)- Q_{-1,i\nu}(k_{\bot}\tau_1)] &
1873: k_r \nu [ Q_{1,i\nu}(k_{\bot}\tau_2) -Q_{-1,i\nu}(k_{\bot}\tau_1)]\\
1874: \nu k_s [ Q_{-1,i\nu}(k_{\bot}\tau_2) -Q_{1,i\nu}(k_{\bot}\tau_1)] &
1875: \nu^2 [ Q_{1,i\nu}(k_{\bot}\tau_2) -Q_{1,i\nu}(k_{\bot}\tau_1)]
1876: \end{array} \right]_{lm}
1877: e^{-i\nu(\eta_1-\eta_2) }~.\nonumber
1878: \end{eqnarray}
1879: The diagonal components of this longitudinal part of the gluon propagator
1880: are just the differences between the vector potentials of the ``static''
1881: gluon fields at the final time $\tau_1$ and the initial time $\tau_2$,
1882: \begin{eqnarray}
1883: D^{[long]}_{rs}= {k_rk_s\over k_t^2}
1884: \bigg\{ {\coth |\eta|\over 2} (e^{-\tau_1 k_t\sinh |\eta|}-
1885: e^{-\tau_2 k_t\sinh |\eta|}) \hspace{6cm}\nonumber \\
1886: -~ \underline{\int {d\alpha\over 2\pi}
1887: \tanh(\alpha+{\eta\over 2}) \tanh(\alpha-{\eta\over 2})
1888: \big[~\sin k_tT_1-\sin k_tT_2~\big]}~\bigg\}~,
1889: \label{eq:A1.14}
1890: \end{eqnarray}
1891: \begin{eqnarray}
1892: D^{[long]}_{\eta\eta}= - {\tau_1^2-\tau_2^2\over 2}\delta(\eta)
1893: -\bigg\{\bigg[\bigg({\cosh\eta\over k_t^2\sinh^3|\eta|}+
1894: {\tau_1\cosh\eta\over k_t\sinh^2\eta}+
1895: {\tau_1^2\cosh\eta\over 2\sinh |\eta|}\bigg)e^{-\tau_1 k_t\sinh |\eta|}\bigg]
1896: -\bigg[\tau_1\to\tau_2\bigg]\bigg\} \nonumber\\
1897: -\underline{\int {d\alpha\over 2\pi}{1\over k_t^2\cosh^2(\alpha+{\eta\over 2})
1898: \cosh^2(\alpha-{\eta\over 2})}
1899: \big[~ \sin k_tT_1-\sin k_tT_2-k_tT_1\cos k_tT_1
1900: +k_tT_2\cos k_tT_2~\big]} ~.
1901: \label{eq:A1.15}
1902: \end{eqnarray}
1903: By the derivation, these components include $\theta(\tau_1-\tau_2)$ of the
1904: following origin. The source current which acts at the moment $\tau_2$
1905: produces the simultaneous longitudinal electric field $E(\tau_2)$. The gauge
1906: field potential is rebuilt from the electric field at the time $\tau_1>\tau_2$
1907: by integrating the electric field $E(\tau)$ over all times from $0$ to
1908: $\tau_1$. The underlined terms in Eqs.~(\ref{eq:A1.14}) and (\ref{eq:A1.15})
1909: cancel out in the full assembly of the retarded propagator
1910: $D_{[adv]}(\tau_2,\tau_1)$ with the underlined terms in the radiation part,
1911: Eqs.~(\ref{eq:A1.4}) and (\ref{eq:A1.5}). In the body of the paper, we
1912: call the residue of this cancelation as the diagonal components of
1913: $D^{[long]}$, which can be conveniently written as
1914: \begin{eqnarray}
1915: D^{[long]}_{rs}= - {k_rk_s\over k_t^2}
1916: ~{k_t\cosh \eta\over 2}
1917: \int_{\tau_2}^{\tau_1} e^{-t k_t\sinh |\eta|} dt~,
1918: \label{eq:A1.14a}
1919: \end{eqnarray}
1920: \begin{eqnarray}
1921: D^{[long]}_{\eta\eta}= - {\tau_1^2-\tau_2^2\over 2}\delta(\eta)
1922: + {k_t\cosh \eta\over 2}
1923: \int_{\tau_2}^{\tau_1} e^{-t k_t\sinh |\eta|} t^2~ dt~.
1924: \label{eq:A1.15a}
1925: \end{eqnarray}
1926: Once again, the ``off-diagonal'' components of the longitudinal part of
1927: propagator,
1928: \begin{eqnarray}
1929: D^{[long]}_{r\eta}= {i k_r\over k_t^2}~
1930: \big\{ -{{\rm sign}\eta\over 2\sinh^2\eta} (e^{-\tau_1 k_t\sinh |\eta|}-
1931: e^{-\tau_2 k_t\sinh |\eta|})
1932: +{k_t\tau_2\over\sinh\eta}e^{-\tau_2 k_t\sinh |\eta|}-
1933: {k_t\tau_1\cosh^2\eta\over\sinh\eta}e^{-\tau_1 k_t\sinh |\eta|}\nonumber\\
1934: -\int {d\alpha\over 2\pi}
1935: {\tanh(\alpha-{\eta\over 2})\over \cosh^2(\alpha+{\eta\over 2})}
1936: [\sin k_tT_1-\sin k_tT_2+ k_t T_2\cos k_t T_2]\big\}
1937: \label{eq:A1.16}
1938: \end{eqnarray}
1939: \begin{eqnarray}
1940: D^{[long]}_{\eta r}= {i k_r\over k_t^2}
1941: \big\{ -{{\rm sign}\eta\over 2\sinh^2\eta} (e^{-\tau_1 k_t\sinh |\eta|}-
1942: e^{-\tau_2 k_t\sinh |\eta|})
1943: -{k_t\tau_1\over\sinh\eta}e^{-\tau_1 k_t\sinh |\eta|}
1944: +{k_t\tau_2\cosh^2\eta\over\sinh\eta}e^{-\tau_2 k_t\sinh |\eta|}\nonumber\\
1945: -\int {d\alpha\over 2\pi}
1946: {\tanh(\alpha+{\eta\over 2})\over \cosh^2(\alpha-{\eta\over 2})}
1947: [-\sin k_tT_1+\sin k_tT_2+ k_t T_1\cos k_t T_1]\big\}
1948: \label{eq:A1.17}
1949: \end{eqnarray}
1950: are odd with respect to $\eta$ and do not contribute the effective quark mass.
1951:
1952: The gauge-field correlators have several distinctive features. First, the
1953: lengthy expression for each component is such that the gauge field correlators
1954: obey the boundary condition $A_{\eta}(\tau =0, \vec{r_t})=0$ which provides
1955: continuity of the field at $\tau=0$, and allows for a complete fixing of the
1956: gauge. Second, in the $rs-$ and $\eta\eta$-components of the propagator
1957: \begin{eqnarray}
1958: D^{lm}_{[adv]}(\tau_2,\tau_1,\eta;k_t)=
1959: -\theta(\tau_1-\tau_2)D^{lm}_{[0]}(\tau_2,\tau_1,\eta;k_t)
1960: +D^{lm}_{L}(\tau_2,\tau_1,\eta;k_t)~,
1961: \label{eq:A1.18}
1962: \end{eqnarray}
1963: the boundary terms cancel between the transverse and longitudinal parts.
1964: This fact provides causal behavior of the $\Sigma(\tau_1,\tau_2)$ that
1965: defines the dispersion law.
1966:
1967: The fermion invariants $g_{[\alpha]}$ were derived in paper [II].
1968: For the sake of completeness, we reproduce the final answers here,
1969: \begin{eqnarray}
1970: g^{L(\pm)}_{[0]}=
1971: i~{\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over 4\sqrt{|\tau_{12}^2|}}
1972: \theta(\tau_{12}^2)~J_1(q_t\sqrt{|\tau_{12}^2|})~,~~~~
1973: g^{T(\pm)}_{[0]}= -~{ e^{\mp\eta/2}\over 4}
1974: \theta(\tau_{12}^2)~J_0(q_t\sqrt{|\tau_{12}^2|})~,
1975: \label{eq:A1.19}\end{eqnarray}
1976: \begin{eqnarray}
1977: g^{L(\pm)}_{[1]}= - \int{d\theta' \over 4\pi}
1978: \big[1- 2 n_f\big({\eta_1+\eta_2\over 2}+ \theta',p_t\big)\big]
1979: ~e^{\mp\theta'}
1980: ~\sin\big( p_t[\tau_1\cosh(\theta-\eta/2)-
1981: \tau_2\cosh(\theta+\eta/2)]\big)\nonumber\\
1982: ={\tau_1 e^{\mp\eta/2}-\tau_2e^{\pm\eta/2}\over 4\sqrt{|\tau_{12}^2|}}
1983: \bigg[\theta(\tau_{12}^2)~Y_1\bigg(q_t\sqrt{|\tau_{12}^2|}\bigg)
1984: +{2\over\pi}\theta(-\tau_{12}^2)
1985: K_1\bigg(q_t\sqrt{|\tau_{12}^2|}\bigg)\bigg]
1986: ~\big[ 1-2n_f(q_t) \big] ~,\nonumber\\
1987: g^{T(\pm)}_{[1]}= -i e^{\mp\eta/2} \int{d\theta' \over 4\pi}
1988: \big[1- 2 n_f\big({\eta_1+\eta_2\over 2}+ \theta',p_t\big)\big]
1989: ~\cos\big( p_t[\tau_1\cosh(\theta-\eta/2)-
1990: \tau_2\cosh(\theta+\eta/2)]\big)\nonumber\\
1991: = i{ e^{\mp\eta/2}\over 4}
1992: \bigg[\theta(\tau_{12}^2)~Y_0\bigg(q_t\sqrt{|\tau_{12}^2|}\bigg)
1993: -{2\over\pi}\theta(-\tau_{12}^2)
1994: K_0\bigg(q_t\sqrt{|\tau_{12}^2|}\bigg)\bigg]~\big[ 1-2n_f(q_t) \big]~.
1995: \label{eq:A1.20}
1996: \end{eqnarray}
1997:
1998:
1999:
2000:
2001: \begin{references}
2002: \bibitem{tev} A. Makhlin and E. Surdutovich, Phys.Rev. {\bf C 58} (1998) 389
2003: (quoted as paper [I]).
2004: \bibitem{Xiong}L. Xiong, E.V. Shuryak, Nucl. Phys {\bf A590} (1995) 589.
2005: \bibitem{Eskola} Kari J. Eskola, Berndt Muller, and Xin-Nian Wang,
2006: Selfscreened parton cascades, Preprint DUKE-TH-96-120 ( nucl-th/9608013).
2007: \bibitem{gqm} A. Makhlin, {\em Scenario for Ultrarelativistic Nuclear
2008: Collisions: II.~ Geometry of quantum states at the earliest stage.}
2009: hep-ph/0007300 (quoted as paper [II])
2010: \bibitem{wdg} A. Makhlin, {\em Scenario for Ultrarelativistic Nuclear
2011: Collisions: III.~ Gluons in the expanding geometry.}
2012: hep-ph/0007301 (quoted as paper [III])
2013: \bibitem{Keld} L.V. Keldysh, Sov. Phys. JETP {\bf 20} (1964) 1018;
2014: E.M. Lifshits, L.P. Pitaevsky, Physical kinetics,
2015: Pergamon Press, Oxford, 1981.
2016: \bibitem{QFK} A. Makhlin, Phys.Rev. {\bf C 51} (1995) 34
2017: \bibitem{QGD} A. Makhlin, Phys.Rev. {\bf C 52} (1995) 995.
2018: \bibitem{Watson} G.N. Watson, Treatise on the theory of Bessel functions
2019: \bibitem{Gel'fand} I.M. Gel'fand and G.E. Shilov, Generalized functions,
2020: New York, Academic Press, 1964
2021: \bibitem{Landau} L.D. Landau and E.M. Lifshits, Quantum mechanics, Sec.1,
2022: Oxford ; New York : Pergamon Press, 1977.
2023: \bibitem{Dirac} P.A.M. Dirac, Rev.Mod. Phys, {\bf 21}, 392 (1949).
2024: \bibitem{Fock} V.A.Fock, Z. f. Phys. {\bf 57}, 261 (1929).
2025: \bibitem{Witten} M.B. Green, J.H. Schwarz, E. Witten, Superstring
2026: theory, Cambridge University press, 1987.
2027: \bibitem{Berest}V.B. Berestetskii, E.M. Lifshitz, and L.P. Pitaevskii,
2028: Quantum electrodynamics, Oxford ; New York :
2029: Pergamon Press, 1982.
2030: \end{references}
2031: \end{document}
2032:
2033:
2034: