1: %Written in Hamburg November 2006
2: %Version as of 04.12.2006
3: %adds by GPV on 04-06.12.2006
4: %Version as of 08.12.2006 (final)
5: %Revised version after submission for citations of 21.12.2006
6: %
7: \documentclass[12pt]{article}
8: \topmargin=-0.5in
9: \oddsidemargin=-0.0in
10: \textheight=8.75in
11: \textwidth=6.5in
12: \baselineskip=20pt
13: %\documentstyle[12pt]{article}
14: %\renewcommand{\baselinestretch}{1.}
15: \input epsf
16: %\textwidth 175mm
17: %\textheight 240mm
18: %\columnsep 38pt
19: %\topmargin -0.5cm
20: %\oddsidemargin 5pt
21: \def\bc{\begin{center}}
22: \def\ec{\end{center}}
23: \def\beq{\begin{equation}}
24: \def\eeq{\end{equation}}
25: \def\bea{\begin{eqnarray}}
26: \def\eea{\end{eqnarray}}
27: \def\noi{\noindent}
28: \def\phid{\phi^{\dagger}}
29: \def\bphi{\bar{\phi}}
30: \def\bphid{\bar{\phi}^{\dagger}}
31: \def\pd{\partial}
32: \def\by{\bar{y}}
33: %
34: %
35: \title{\bf On the "toy model" in the Reggeon Field Theory}
36: %\vskip 0.5 cm
37: %
38: \author{M.A.Braun$^{a)}$, and G.P.Vacca$^{b)}$ \\
39: $^{a)}$ S.Peterburg State University, Russia\\
40: $^{b)}$INFN and Department of Physics, Via Irnerio 46, Bologna, Italy}
41: %
42: \begin{document}
43: \maketitle
44:
45: \begin{abstract}
46:
47: The Reggeon field theory with zero transverse dimensions is studied in the Hamiltonian
48: formulation for both sub-and supercritical pomeron. Mathematical aspects of the model,
49: in particular the scalar products in the space of quantum states, are discussed.
50: Relation to reaction-diffusion processes is derived in absence of pomeron merging.
51: Numerical calculations for different parameters of the models, $\alpha(0)-1=\mu$ and the
52: triple pomeron coupling constant $\lambda$, show that the triple
53: pomeron interaction always makes amplitudes fall with rapidity irrespective of the
54: value of the intercept. The smaller the values of the ratio $\lambda/\mu$ the
55: higher are rapidities $y$ at which this fall starts, so that at small values of
56: $\lambda$ it begins at asymptotically high rapidities
57: (for $\lambda/\mu<1/4$ the fall is
58: noticeable only at $\mu y>100$).
59: No visible singularity is seen for the critical pomeron.
60: A perturbative treatment is proposed which may be useful for more realistic models.
61: \end{abstract}
62:
63: \section{Introduction}
64: At high energies the Quantum Chromodynamics tells us that particle interaction may be
65: mediated by the exchange of hard pomerons, which are non-local entities propagating according to
66: the BFKL equation and splitting into two or merging from two to one with the known triple pomeron
67: verteces. This most complicated picture can be drastically simplified in the quasi-classical
68: treatment supposedly valid when at least one of the colliding hadrons is a heavy nucleus.
69: This leaves only tree diagrams which can be summed by comparatively simple equations, which at least
70: admit numerical solution.
71: However at present much attention is drawn to the contribution of loops. This is obviously a
72: problem of a completely different level of complexity, since it
73: amounts to solving a full-fledged non-local quantum field theory. At present a solution of even some
74: local quantum field theory has not been known except for the one-dimensional case.
75: Naturally for the study of loops in QCD an one-dimensional model has been lately chosen as
76: a starting point.
77:
78: The one-dimensional ("toy") model for the QCD interaction was proposed and solved by A.H.Mueller
79: ~\cite{AHM}. However much earlier a model similar in structure was studied in the framework of
80: a local Reggeon field theory (RFT) with a supercritical pomeron and imaginary triple pomeron interaction
81: in a whole series of papers ~\cite{amati1,alessandrini,jengo,amati2,ciafaloni1,ciafaloni2}
82: As expected the one-dimensional reduction of the RFT
83: admitted more or less explicit solution and lead to definite predictions as to the behaviour of
84: amplitudes both for hA and AB interaction (of course with the understanding of a highly
85: artificial physical picture, very far from the realistic world). In the course of study many
86: ingeneous tricks and techniques were employed and the underlying complicated dynamics was revealed.
87: However some subtle points have remained unclear in our opinion. In particular the nature of
88: correspondence between the functional and Hamiltonial approaches
89: and the origin of restrictions on the form of wave function and the range of its variable
90: have not been fully exposed.
91:
92: To elucidate these points is the primary aim of this note. We also try to follow the line
93: presented in ~\cite{boreskov,bondarenko} to relate the RFT with the
94: probabilistic parton picture and the so-called reaction-diffusion model
95: for the case when there is no merging of pomerons (fan diagrams). This gives us a possibility to
96: find the contribution from the loops with the highest extention in rapidity just by joining two
97: fans propagating from the projectile and target towards the center.
98: We additionally develop a perturbation theory for small $\lambda$ to compute the evolution
99: by noting the PT-symmetry of the model.
100: We finally report on some
101: numerical results on solving the RFT with all loops, which can be obtained quite easily by
102: integrating the corresponding evolution equations.
103:
104: Note that we study exclusively a model with the triple pomeron interaction only.
105: A different model which additionally includes a quartic pomeron
106: interaction with a particular fine-tuned value of the coupling constant
107: has lately received enough attention in literature
108: ~\cite{boreskov, bondarenko, levin1, levin2}. In our opinion the quartic
109: interaction has no immediate counterpart in the realistic pomeron theory
110: with a large number of colours $N_c$, where it is damped by factor $1/N_c^2$.
111: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
112: \section{Functional integral formulation}
113: The toy model of Pomeron interaction with a zero-dimension
114: transverse space can be defined by a generating functional
115: \beq
116: Z=\int D\phi D\phid e^{-S},
117: \label{funint}
118: \eeq
119: where
120: \beq
121: S=\int dy {\cal L},\ \
122: \eeq
123: and
124: \beq
125: {\cal L}=\frac{1}{2}(\phid\phi_y-\phid_y\phi)-\mu\phid\phi
126: +i\lambda\phid\phi(\phid+\phi)-i(j_0\,\phid+j_Y\,\phi).
127: \label{lagrangian}
128: \eeq
129: Here $\mu$ is the pomeron intercept ($\alpha(0)-1$). For the
130: supercritical pomeron $\mu>0$. Triple pomeron coupling constant
131: $\lambda$ is also positive.
132: The sources $j_0=\delta(y)g_1$ amd $j_Y=\delta(y-Y)g_2$ represent the
133: interaction with the
134: target and projectile respectively. In the real world
135: $g_1=g_1(b)=AgT_A(b)$ and $g_2=g_2(b)=BgT_B(b)$ where $g$
136: is the pomeron-nucleon coupling constant, $A$ and $B$ are atomic numbers
137: of the colliding nuclei and $T_A$ and $T_B$ their profile functions.
138:
139: The functional integral (\ref{funint}) converges for $\mu<0$ provided the two
140: field variables $\phi$ and $\phid$ are complex conjugate to one another.
141: In fact putting $\phi=\phi_1+i\phi_2$ and $\phid=\phi_1-i\phi_2$, with
142: both
143: $\phi_{1,2}$ real, we find that all terms in the Lagrangian (\ref{lagrangian}) are pure
144: imaginary except for the mass term proportional to
145: \[ \phid\phi=\phi_1^2+\phi_2^2\geq 0.\]
146: So the integrand in (\ref{funint}) contains a factor
147: $\exp\Big(\mu\int dy (\phi_1^2+\phi_2^2)\Big)$,
148: which guarantees convergence if $\mu<0$ and the integration over
149: the fields $\phi_1$ and $\phi_2$ goes over the real axis. For $\mu>0$
150: the integral does not exist.
151: Note that the integral does not exist for any sign of $\mu$ either,
152: if one changes the
153: integration contour in $\phi_1$ and $\phi_2$ and so over $\phi$ and
154: $\phid$. This means that it is impossible to pass to new fields
155: $u=i\phid$ and $v=i\phi$ and take $u$ and $v$ real directly in the
156: functional integral as in ~\cite{alessandrini},
157: since it requires unlawful rotation of the
158: integration contour in the integration over $\phi_1$.
159:
160: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
161: Obviously the functional formulation can serve only to define the model
162: for the subcritical pomeron with $\mu<0$. To define the theory for
163: positive values of $\mu$ one has to recur to analytic continuation in
164: $\mu$. Note that
165: the properties of the functional integral make one think that there is a
166: singularity
167: at $\mu=0$. In a fully dimensional RFT it was conjectured that this singularity
168: was related to a phase transition. However already in ~\cite{ciafaloni1} it
169: was argued
170: that there was no phase transition in the zero-dimensional world and the
171: amplitudes were analytic in $\mu$ on the whole real axis.
172: As we shall see
173: by direct numerical calculation, there is indeed no singularity at
174: $\mu=0$ in scattering amplitides.
175:
176: In fact the functional formulation is only supported by the fact that it reproduces
177: the perturbative diagrams for the pomeron propagation and interaction.
178: In the following section we introduce an alternative, Hamiltonian formalism, which
179: gives rise to the same perturbative diagrams (such an approach as a starting point
180: was briefly mentioned in ~\cite{jengo}).
181: However in contrast to the
182: functional approach it does not involve any limitation on the value or sign
183: of the intercept $\mu$. Since for $\mu<0$ the perturbative series seems to be
184: convergent,
185: the two formulations are completely equivalent for the subcritical pomeron.
186: Therefore the Hamiltonian formulation gives the desired analytic
187: continuation of the model to positive values of $\mu$.
188: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
189:
190: \section{Hamiltonian formulation}
191: \subsection{Basic definitions}
192: Forgetting the functional integral (\ref{funint}) altogether, we now start from
193: a quasi-Schroedinger equation in rapidity:
194: \beq
195: \frac{d\Psi(y)}{dy}=-H\Psi(y)
196: \eeq
197: and postulate the form of the Hamiltonian $H$
198: \beq
199: H=-\mu \phid\phi +i\lambda \phid(\phi+\phid)\phi
200: \label{firstH}
201: \eeq
202: as a function of two
203: Hermithean conjugate operators
204: $\phi$ and $\phid$
205: with the commutator
206: \beq
207: [\phi,\phid]=1.
208: \label{commrel}
209: \eeq
210: The Hamiltonian is obviously non-Hermithean. We standardly split it
211: into a free and interaction parts:
212: \beq
213: H=H_0+H_I,\ \ H_0=-\mu \phid\phi,\ \ H_I= i\lambda \phid(\phi+\phid)\phi.
214: \label{Hdecomp}
215: \eeq
216:
217: Next step is to define physical observables in this picture.
218: In accordance with the commutation relation (6\ref{commrel}) we define the vacuum state
219: $\Psi_0$, normalized to unity, by the condition
220: \beq
221: \phi\Psi_0=0.
222: \eeq
223: All other states will be built from $\Psi_0$ by application of some number
224: of operators $\phid$. We
225: postulate that the transition amplitude from the initial state
226: $\Psi_i$ at rapidity $y=0$ to the final state $\Psi_f$
227: at rapidity $y$ is given by
228: \beq
229: iA_{fi}=\langle \Psi_f|\Psi_i(y)\rangle=\langle
230: \Psi_f|e^{-Hy}|\Psi_i\rangle.
231: \eeq
232: Here
233: \beq
234: \Psi_i(y)=e^{-Hy}\Psi_i
235: \eeq
236: is the initial state evolved to rapidity $y$.
237: The amplitude $A_{fi}$ is imaginary positive so that the matrix element on
238: the right-hand side of (9) is negative.
239:
240: It will be convenient to define the initial and final states in terms of some operators
241: acting on the vacuum state:
242: \beq
243: \Psi_i=F_i(\phid)\Psi_0,\ \ \Psi_f=F_f(\phid)\Psi_0.
244: \eeq
245: This allows to rewrite the amplitude as a vacuum average
246: \beq
247: iA_{fi}=\langle F_{f}^*(\phi)e^{-Hy}F_i(\phid)\rangle,
248: \eeq
249: where we do not explicitly write the vacuum state $\Psi_0$ in which the
250: average is taken.
251:
252: \subsection{Perturbative expansion}
253: We are going to prove that the perturbative expansion of this theory
254: leads to the same
255: Feynman diagrams as from the functional approach. To this end we present
256: \beq
257: iA_{fi}=\langle
258: e^{H_0y}F_{f}^*(\phi)e^{-H_0y}e^{H_0y}e^{-Hy}F_i(\phid)\rangle.
259: \eeq
260: From (\ref{commrel}) it follows that
261: \beq
262: \phi(y)\equiv e^{H_0y}\phi e^{-H_0y}=e^{\mu y}\phi.
263: \label{phievol}
264: \eeq
265: The operator
266: \beq
267: U(y)\equiv e^{H_0y}e^{-Hy}
268: \eeq
269: satisfies the equation
270: \beq
271: \frac{dU(y)}{dy}=e^{H_0y}(H_0-H)e^{-H(y)}=-e^{H_0y}H_ie^{-H_0y}U(y)
272: =H_I(y)U(y),
273: \label{inteq}
274: \eeq
275: where $H_I(y)$ is the interaction in the interaction representation:
276: \beq
277: H_I(y)=e^{H_0y}H_I(\phi,\phid)e^{-H_0y}=H_I(\phi(y),\phid(y)),
278: \eeq
279: with the rapidity dependent operators (14) and
280: \beq
281: \phid(y)\equiv e^{H_0y}\phid e^{-H_0y}=e^{-\mu y}\phid.
282: \label{phidevol}
283: \eeq
284: Note that at $y>0$ the operators $\phi(y)$ and $\phid(y)$ are no more
285: Hermithean conjugate to each other, since the normal time
286: is changed to the imaginary one. However the similarity transformations
287: (\ref{phievol}) and (\ref{phidevol})
288: do not change the commutation relations.
289: The solution of (\ref{inteq}) with the obvious initial condition $U(0)=1$ is
290: standard
291: \beq
292: U(y)=T_y\exp\Big\{-\int_0^y dy'H_i(y')\Big\},
293: \eeq
294: where $T_y$ means ordering the operators from left to right according to
295: decreasing $y$'s. So we get the expression for the amplitude as
296: \beq
297: iA_{fi}=\langle F^*_f(e^{\mu y}\phi)T_y\exp\Big\{-\int_0^y
298: dy'H_i(y')\Big\}
299: F_i(\phid)\rangle.
300: \eeq
301:
302: Expansion of the $T_y\exp$ gives rise to standard Feynman diagrams.
303: Verteces for pomeron splitting and merging will be given by the expected
304: factors
305: $i\lambda$. The pomeron propagator will be given by
306: \beq
307: P(y_1-y_2)=\langle T_y\Big\{\phi(y_1),\phid(y_2)\Big\}\rangle=
308: e^{\mu(y_1-y_2)}\theta(y_1-y_2),
309: \eeq
310: which is the correct pomeron propagator in our toy model.
311: So indeed the theory defined by this Hamiltonial formulation gives
312: rise to standard Feynman diagrams for the RFT.
313:
314: Note that the Hamiltonian formulation is based on the evolution equation
315: in rapidity and does not require the intercept $\mu$ to have a definite sign.
316: It looks equally good both for positive and negative values of $\mu$.
317: For negative $\mu$ it produces a perturbative expansion which is
318: identical to that in the functional approach. So it gives the desired
319: analytic continuation of the functional approach to the region of
320: positive $\mu$, where the latter looses sense.
321:
322: \subsection{Passing to a real Hamiltonian}
323: It is perfectly possible to continue with the originally
324: defined Schroedinger operators $\phi$ and $\phid$ which are
325: Hermithean conjugate to each other and satisfy the commutation
326: relation (\ref{commrel}). However it is convenient to pass to new operators
327: $u$ and $v$ in terms of which all ingredients of the theory
328: become real. Following the old papers (starting from ~\cite{amati1}) we define
329: \beq
330: u=i\phid,\ \ v=i\phi.
331: \eeq
332: The new fields are anti-Hermithean to each other
333: \beq
334: u^\dagger=-v,\ \ v^\dagger=-u
335: \eeq
336: and satisfy the commutation relation
337: \beq
338: [v,u]=-1.
339: \label{commrel2}
340: \eeq
341: However $v\Psi_0=0$ and the states are generated by action of $u$.
342: So $v$ is the annihilation and $v$ is the creation operator with
343: abnormal commutation relation (\ref{commrel2}).
344:
345: In terms of new fields the Hamiltonian becomes real
346: \beq
347: H=\mu u v-\lambda u(u+v)v.
348: \label{Huv}
349: \eeq
350: It is important that also the operators creating the initial and final
351: states become real. It is normally assumed that for the scattering
352: of two nuclei the initial and final wave functions should correspond to the
353: eikonal picture. In terms of creation operator $u$ this leads to
354: \beq
355: \Psi_{i(f)}=F_{i(f)}(u)\Psi_0,\ \ F_{i(f)}(u)=1-e^{-g_{i(f)}u}.
356: \eeq
357: So the expression for the amplitude is rewritten in terms of
358: real quantities:
359: \[
360: iA_{fi}=\langle F_f(u)\Psi_0|e^{-Hy}|F_i(u)\Psi_0\rangle=
361: \langle\Psi_0|F^*_f(-v)e^{-Hy}F_i(u)\Psi_0\rangle\]\beq=
362: \langle \Big(1-e^{-g_fv}\Big)e^{-Hy}\Big(1-e^{-g_iu}\Big)\rangle.
363: \eeq
364: Since $H\Psi_0=0$ the term independent of $g_i$ and $g_f$ vanishes, so
365: that
366: we can also write
367: \beq
368: iA_{fi}=
369: -\langle e^{-g_fv}e^{-Hy}\Big(1-e^{-g_iu}\Big)\rangle=
370: -\langle\Psi_0|e^{-g_fv}F_i(y,u)\Psi_0\rangle,
371: \label{ampli1}
372: \eeq
373: where $F_i(y,u)$ is the operator which creates the evolved initial state.
374: Since
375: \beq
376: \Psi_i(y)=e^{-Hy}\Psi_i=e^{-Hy}F_i(u)\Psi_0\equiv F_i(y,u)\Psi_0,
377: \eeq
378: we have
379: \beq
380: \frac{\pd F_i(y,u)}{\pd y}=-H(u,v)F_i(y,u)
381: \label{evol1}
382: \eeq
383: with an initial condition
384: \beq
385: F_i(0,u)=1-e^{-g_iu}.
386: \label{initF}
387: \eeq
388: The commutation relation (\ref{commrel2}) allows to represent
389: \beq
390: v=-\frac{\pd}{\pd u}
391: \label{vrepr}
392: \eeq
393: and then (\ref{ampli1}) implies that to find the amplitude one has to substitute
394: $u$ by $g_f$ in $F_i(y,u)$
395: \beq
396: iA_{fi}=-F_i(y,g_f).
397: \eeq
398:
399: This gives a practical recipe for the calculation of the amplitude.
400: One has to solve Eq. (\ref{evol1}) with the Hamiltonian (\ref{Huv}) in which $v$ is
401: represented according to (\ref{vrepr}):
402: \beq
403: H=-\mu u \frac{\pd}{\pd u}+\lambda u^2\frac{\pd}{\pd u}-
404: \lambda u\frac{\pd^2}{\pd u^2}
405: \label{Huvdiff}
406: \eeq
407: and the initial condition (\ref{initF}). After the solution is found,
408: one has to substitute
409: $u$ by $g_f$ in it.
410:
411:
412: \subsection{Target-projectile symmetry}
413: Taking the complex conjugate of (\ref{ampli1}) we find
414: \beq
415: -iA^*_{fi}=
416: \langle \Big(1-e^{g_iv}\Big)e^{-H^{\dagger}y}\Big(1-e^{g_fu}\Big)\rangle
417: =iA_{if}(\lambda\to -\lambda,g_{i(f)}\to -g_{f(i)}).
418: \eeq
419: Having in mind that the amplitude is pure imaginary,
420: we see that interchanging the target and projectile leads to changes of
421: sign in
422: the triple pomeron coupling and the couplings to the external particles.
423: On the other hand changing $u\to -u$ in Eq. (\ref{evol1}) with the Hamiltonian
424: (\ref{Huvdiff}) and the initial condition
425: (\ref{initF}) and also changing signs of $\lambda$ and $g_i$ obviously does not
426: change the solution:
427: \beq
428: F_i(y,-u,-g_i,-\lambda)=F_i(y,u,g_i,\lambda),
429: \eeq
430: from which we find
431: \beq
432: F_i(y,-g_f,-g_i,-\lambda)=F_i(y,g_f,g_i,\lambda).
433: \eeq
434: So the interchange of the target and projectile
435: does not change the amplitude.
436:
437: \subsection{Scalar products}
438: As follows from our derivation the exact mathematical realization of the
439: scalar product in the space of wave functions $\Psi$ is irrelevant for
440: the resulting formulas. The representation of the operator $v$ as a
441: minus derivative in $u$ serves only to express its algebraic action on
442: the operator $u$ in accordance with the commutation relation.
443: It does not require to introduce a representation for the wave function $\Psi$
444: in which $u$ is represented as multiplication by a complex number and
445: correspondingly the scalar product is defined by means of integration
446: over the whole complex plane $u$ with a weight factor $e^{-uu^*}$
447: aas in ~\cite{ciafaloni1}.
448: One may instead represent operators $u$ and $v$ in terms of the standard
449: Hermithean operators $p$ and $q$ as
450: \beq
451: u=\frac{i}{\sqrt{2}}(q-ip),\ \ v=\frac{i}{\sqrt{2}}(q+ip)
452: \eeq
453: and define a scalar product as
454: \beq
455: \langle\Psi_1|\Psi_2\rangle=\int dq\Psi^*_1(q)\Psi_2(q)
456: \eeq
457: with $p=-i\pd/\pd q$. The vacuum state $\Psi_0$ will be given by the
458: oscillator ground state in this representation.
459:
460: However this is not the end of the story. New aspects arise in the
461: process of solution of the Eq. (\ref{evol1}) for $F_i$. The equation itself is
462: regular everywhere except
463: possibly at $u=0$, where however it is also regular having behaviour as a
464: constant or $u$. We are obviously interested in the last behaviour since
465: from the equation it follows that
466: $F_i(y,u=0)=0$.
467: So in principle our solution turns out to be regular in the whole
468: complex $u$-plane. Formally this means that one may choose any continuous
469: interval in this plane and evolve the initial function given in this interval up
470: to the desired values of rapidity. The immediate problem is what interval one has
471: to choose. The final expression for amplitude requires that the value of $g_f$ has
472: to belong to this interval. In particular for positive $g_f$ the interval should
473: include at least a part of the positive axis.
474:
475: Numerical calculations show that it is indeed possible to find a solution to the
476: evolution equation provided the initial interval is restricted to {\it only} positive
477: values of $u$ (including the desired value $u=g_f$).
478:
479: To understand this phenomenon one has to recall some results from the earlier
480: studies.
481: It was found in ~\cite{jengo} that there exists a transformation of both the
482: function and variable
483: which converts the non-Hermithean Hamiltonian (\ref{Huvdiff}) into a
484: Hermithean one.
485: However it is only possible for values of $u$ covering the positive axis rather than
486: the whole real axis where the equation should hold.
487: Namely one introduces variable $x$ by
488: \beq
489: u=x^2>0
490: \label{upos}
491: \eeq
492: and makes a similarity transformation of the Hamiltonian
493: \beq
494: \tilde{H}=W^{-1}HW, \ \
495: W(x)=\sqrt{x}e^{\frac{1}{4}x^4-\frac{\mu}{2\lambda}x^2}
496: \label{iengoform}
497: \eeq
498: to find the transformed Hamiltonian as
499: \beq
500: \tilde{H}=\frac{\lambda}{4}\Big(-\frac{\pd^2}{\pd x^2}+\frac{3}{4x^2}+x^2(x^2-\rho)^2-2x^2\Big)
501: \label{Hiengoform}
502: \eeq
503: where we standardly denote $\rho=\mu/\lambda$.
504: The Hamiltonian $\tilde{H}$ is a well defined Hamiltonian on the whole real axis with
505: a singularity at $x=0$. The interaction potential infinitely grows at $|x|\to \infty$.
506: So restricting to $x>0$ and
507: requiring eigenfunctions to vanish at the origin and at $x\to\infty$ one finds a
508: discrete spectrum, starting from a certain ground state $E_0$ to infinity.
509: The transformed initial function
510: \beq
511: \tilde{F}_i(y=0,x)=\frac{1}{\sqrt{x}}e^{-\frac{1}{4}x^4+\frac{\mu}{2\lambda}x^2}
512: \Big(1-e^{-g_ix^2}\Big)
513: \eeq
514: is well behaved both at $x=0$ and $x=\infty$ and at $0<x<\infty$ can be expanded in the
515: complete set of eigenfunctions $\tilde{F}_n$ of $\tilde{H}$:
516: \beq
517: \tilde{F}_i(y=0,x)=\sum_{n=0}\langle
518: \tilde{F}_i(y=0,x)|\tilde{F}_n\rangle \tilde{F}_n(x).
519: \label{initF2}
520: \eeq
521: Here a new scalar product has been introduced for the transformed functions:
522: \beq
523: \langle\tilde{F}_1|\tilde{F}_2\rangle=\int_0^\infty dx\tilde{F}^*_1(x)\tilde{F}_2(x).
524: \eeq
525: Evolution immediately gives
526: \beq
527: \tilde{F}_i(y,x)=\sum_{n=0}\langle
528: \tilde{F}_i(y=0,x)|\tilde{F}_n\rangle e^{-E_ny}\tilde{F}_n(x)
529: \eeq
530: and applying the operator $W$ one finds the solution as
531: \beq
532: F(y,u)=\sum_{n=0}\langle\tilde{F}_i(y=0,x)|\tilde{F}_n\rangle
533: e^{-E_ny}W(x)\tilde{F}_n(x),
534: \ \ u=x^2.
535: \eeq
536:
537: Note that we could just as well start from negative values of $u$ and define
538: instead of (\ref{upos})
539: \beq
540: u=-x^2,\ \ x>0
541: \eeq
542: Doing the same transformation of the Hamiltonian we then obtain (\ref{Hiengoform}) with the
543: opposite
544: overall sign and an opposite sign of $\rho$,
545: that is the transformed Hamiltonian is now $-\tilde{H}(\rho\to -\rho)$. All eigenvalues now
546: are negative and go down to $-\infty$.
547: If we take the expansion of the initial function analogous to (\ref{initF2}) and try to
548: evolve it we obtain a sum
549: with infinitely growing exponentials in $y$. Obviously such evolution has no sense.
550: This explains why with the positive or negative signs of $\lambda$ and
551: $g$'s one can start evolution
552: only from correspondingly positive or negative values of $u$, but never
553: from intervals including both.
554:
555: Also note that considered as a function of $u$, eigenfunctions are analytic
556: in the whole $u$-plane. Starting from, say, $u>0$ their values at $u<0$ can be obtained by
557: analytic continuation to pure imaguinary values of variable $x$. Obviously eigenfunctions
558: should exponentially grow as $u\to -\infty$ (otherwise we shall find positive eigenvalues
559: of the continued Hamiltonian $-\tilde{H}(\rho\to -\rho)$).
560: Similarly starting from $u<0$ and passing to
561: positive $u$ one finds that eigenfunctions exponentially grow at large positive $u$.
562: So considering the initial Hamiltonian $H$ on the whole real $u$ axis we find that its
563: spectrum goes from $-\infty$ to $+\infty$ with eigenfunctions which vanish only either
564: at large positive $u$ for positive eigenvalues or at large negative $u$ for negative
565: eigenvalues and exponentially grow otherwise. This prevents introduction of the
566: Bargmann scalar product with integration over the whole complex $u$-plane.
567:
568: Finally let us investigate the qualitative behaviour of the system in the
569: large $\lambda/\mu=1/\rho$ limit, which may be of interest in the generalization
570: to more dimensions, where the effective $\lambda$ is (infinitely) large
571: (see \cite{amati2}).
572: For this purpose the representation given in Eq. (\ref{Hiengoform}) is
573: convenient, since it admits perturbative expansion in $\rho$.
574: In the limit $\rho=0$ one has
575: \beq
576: \tilde{H}=
577: \frac{\lambda}{4}\Big(-\frac{\pd^2}{\pd x^2}+\frac{3}{4x^2}+x^6-2x^2\Big)
578: \label{iengoform0}
579: \eeq
580: which means that the amplitude will behave as
581: $e^{- y E_0}$, where $E_0$ is the ground state energy of $\tilde{H}$.
582:
583: One can perform a semiclassical estimation of the ground state of the
584: Hamiltonian in Eq. (\ref{iengoform0}) by imposing the condition
585: \beq
586: \int_{x_1(4E/\lambda)}^{x_2(4E/\lambda)} dx
587: \sqrt{4E/\lambda-V(x)}=\left(n+\frac{1}{2}\right)\pi,
588: \label{semiclassical}
589: \eeq
590: where $V(x)=\frac{3}{4x^2}+x^6-2x^2$ and $x_i$ are the positive real solutions
591: of the equation $V(x)=4 E/\lambda$.
592: For $n=0$ one obtains $E_0 \simeq 3.7 \, \lambda/4 >0$, which is in a
593: very good agreement with the results of numerical
594: evolution for $\lambda/\mu \ge 3$, which indeed shows
595: a decay of the amplitude
596: as $e^{-y E_0}$ after a short initial period of evolution.
597:
598: Note that during this
599: initial period one observes a rise of the amplitude, which may seem strange in view
600: of the fact that all eigenvalues of the transformed Hamiltonian (\ref{iengoform})
601: are in fact positive. This rise is explained by the interference of different
602: contributions which may enter with different signs due to the non-unitarity
603: of the similarity transformation (\ref{iengoform}) from the initial non-Hermithean Hamiltonian.
604:
605: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
606: \section{No pomeron fusion: fan diagrams}
607: \subsection{Solution}
608: A drastic simplification occurs if in the Hamiltonian one drops the term corresponding to
609: fusion of pomerons ('fan case'):
610: \beq
611: H\to H_{fan}=\mu uv-\lambda u^2v=-(\mu u-\lambda u^2)\frac{\pd}{\pd u}.
612: \label{Hfan}
613: \eeq
614: Then, as has been known since ~\cite{schwimmer}, the solution can be
615: immediately obtained in an explicit form.
616: Indeed, following ~\cite{boreskov}, define a new variable $x$ by
617: \beq
618: dz=\frac{du}{\mu u-\lambda u^2},
619: \eeq
620: which gives (up to an irrelevant constant)
621: \beq
622: z=\frac{1}{\mu}\ln\frac{u}{u-\rho},\ \ \rho=\frac{\mu}{\lambda}.
623: \eeq
624: The inverse relation is
625: \beq u=\frac{\rho}{1-e^{-\mu z}}.
626: \eeq
627: Our evolution equation for $F$ takes the form
628: \beq
629: \Big(\frac{\pd}{\pd y}-\frac{\pd}{\pd z}\Big)F(y,z)=0
630: \eeq
631: with a general solution
632: \beq
633: F(y,z)=f(y+z)
634: \label{yshift}
635: \eeq
636: where $f$ is an arbitrary function. It is to be found from the initial
637: condition:
638: \beq
639: F(0,z)=f(z)=1-e^{-g_iu}=1-e^{-g_i\frac{\rho}{1-e^{-\mu z}}}.
640: \eeq
641: From this we obtain~\cite{boreskov}
642: \beq
643: F_{fan}(y,u)=1-\exp\Big(-g_i\frac{\rho}{1-e^{-\mu (y+z)}}\Big)
644: =1-\exp \Big[-\frac{g_iue^{\mu y}}{1+\frac{u}{\rho}\Big(e^{\mu
645: y}-1\Big)}\Big].
646: \label{fansol1}
647: \eeq
648: This solution is evidently non-symmetric in the projectile and target.
649: In practice one considers such an approximation in the study of hA
650: scattering. Then it represents the sum of fan diagrams propagating from
651: the projectile hadron towards the target nucleus. This situation
652: corresponds to the lowest order in powers of $g_i$~\cite{schwimmer}:
653: \beq
654: F^{hA}_{fan}=\frac{g_iue^{\mu y}}{1+\frac{u}{\rho}\Big(e^{\mu y}-1\Big)}
655: \label{fansolhA}
656: \eeq
657: We recall that the amplitude is obtained from (\ref{fansol1}) or (\ref{fansolhA}) by putting
658: $u=g_f$
659:
660: We note that the solution (\ref{fansolhA}) is not analytic in the whole complex $u$
661: plane: it evidently has a pole at some negative value of $u$, which
662: tends to zero from the negative side as $y\to\infty$. So at least for this
663: solution one cannot formulate a scalar product of the Bargmann type with
664: integration over the whole complex $u$ plane. As we have stressed,
665: the existence of such a scalar product is not at all needed.
666:
667: In fact the Hamiltonian $H_{fan}$ has the same spectrum as the free
668: Hamiltonian $H_0$ (Eq. (\ref{Hdecomp})). The two Hamiltonians are related by a
669: similarity transformation
670: \beq
671: H_{fan}=VH_0V^{-1},\ \ V=e^{-u^2v/\rho}.
672: \label{similfan}
673: \eeq
674: It follows that eigenvalues of $H_{fan}$ are all non-positive:
675: $E_n=-\mu n$ with $n=0,1,2,...$. However the amplitudes (\ref{fansol1}) and (\ref{fansolhA})
676: do not grow but rather tend to a constant at large $y$. As seen from the
677: structure of expressions (\ref{fansol1}) and (\ref{fansolhA}) this is a result of interference of
678: contributions with different $E_n$, which separately grow at high
679: $y$.
680: One can reproduce the result of the fan evolution using the form in
681: Eq. (\ref{similfan}) by noting that $e^{-y H_0}$ is an operator which shifts
682: the variable $\ln u$ by $\mu$ while $V$ corresponds to a shift in the variable
683: $1/u$ by $1/\rho$.
684:
685: The fan evolution is to be compared with the situation with the total Hamiltonian
686: $H$, whose eigenvalues are all non-negative for the branch with positive
687: $u$. In this case the solutions generally fall at large $y$ and became
688: nearly constant only at very small $\lambda$ when one of the excited
689: levels goes down practically to zero due to a particular structure
690: of the transformed Hamiltonian (\ref{Hiengoform}) ~\cite{jengo}. So the mechanism
691: for the saturation of the amplitudes at large rapidities is completely
692: different for the complete Hamiltonian $H$ and its fan part $H_{fan}$.
693:
694:
695: \subsection{Relation to the reaction-diffusion approach}
696: The fan case admits a reinterpretation in terms of the so-called
697: reaction-diffusion processes ~\cite{bondarenko}, which may turn out to be helpful for
698: applications.
699: Let us present a solution $F(y,u)$ as a power series in $u$
700: \beq
701: F(y,u)=\sum_{n=1}c_n(y)u^n.
702: \label{powerexp}
703: \eeq
704: The evolution equation gives a system of equations for coefficients
705: $c_n(y)$:
706: \beq
707: \frac {dc_n(y)}{dy}=\mu n c_n(y)-\lambda (n-1)c_{n-1}(y).
708: \eeq
709:
710: Consider first the case $\mu>0$.
711: We rescale $y=\bar{y}/\mu$ and present
712: \beq
713: c_n(\bar{y})=\frac{1}{n!}(-\rho)^{1-n}\nu_n(\bar{y})
714: \label{cn}
715: \eeq
716: to obtain an equation for $\nu_n$
717: \beq
718: \frac{d\nu_n(\bar{y})}{d\bar{y}}= n
719: \nu_n(\bar{y})+n(n-1)\nu_{n-1}(\bar{y}).
720: \label{nuevol}
721: \eeq
722: This equation is identical with the one which is obtained for
723: multiple moments of the probability $P_k(\bar{y})$ to have exactly $k$ pomerons
724: \beq
725: \nu_n(\bar{y})=\sum_{k=n}P_k(\bar{y})\frac{k!}{(k-n)!}
726: \eeq
727: provided the probabilities obey the equation
728: \beq
729: \frac{dP_n(\bar{y})}{d\bar{y}}=-nP_n(\bar{y})+(n-1)P_{n-1}(\bar{y}).
730: \label{probevol}
731: \eeq
732: The equations for $P_n$ are such that they conserve the total probabilty
733: assumed to be equal to unity. Also if the initial probabilities are
734: non-negative, they
735: will remain such during the evolution. This gives the justification
736: for the interpretation of $P_n$ as probabilities.
737:
738: For further purposes it is convenient to introduce a generating functional for
739: the probabilities $P_n$
740: \beq
741: Z(\bar{y},u)=\sum_nP_n(\bar{y})u^n.
742: \eeq
743: In terms of this functional
744: \beq
745: P_n=\frac{1}{n!}\Big(\frac{\pd}{\pd u}\Big)^n Z_{\Big|u=0},\ \
746: \nu_n=\Big(\frac{\pd}{\pd u}\Big)^n Z_{\Big|u=1}.
747: \eeq
748: From the system of equation for $P_n$ one easily finds an equation for $Z(\bar{y},u)$
749: \beq
750: \frac{\pd Z(\bar{y},u)}{\pd\bar{y}}=(u^2-u)\frac{\pd Z(\bar{y},u)}{\pd u}.
751: \eeq
752: Note that this is essentially the same equation as for $F$ with parameters put to unity and
753: the opposite sign of the Hamiltonian. So its solution is immediately
754: obtained from (\ref{yshift}) by
755: changing sign of $y$
756: \beq
757: Z(\bar{y},u)=Z\Big(0,z(u)-\bar{y}\Big),\ \ u=\frac{1}{1-e^z}.
758: \eeq
759:
760: For the simple case of hA scattering this gives
761: \beq
762: Z^{hA}=\frac{ue^{-\by}}{1+u(e^{-\by}-1)}.
763: \eeq
764: Expansion in powers of $u$ gives the probabilities $P_n$ as
765: \beq
766: P_n^{hA}=e^{-\by}\Big(1-e^{-\by}\Big)^{n-1}.
767: \label{PnhA}
768: \eeq
769: For a general case we first construct $Z(y=0)$ as a sum
770: \beq
771: Z(0,u)=1+\sum_{n=1}\nu_n(0) \frac{(u-1)^n}{n!}
772: \eeq
773: with values of $\nu_n(0)$ known from the initial distribution:
774: \beq
775: \nu_n(0)=\rho^{n-1}g_i^n.
776: \eeq
777: Subsequent summation and shift $z\to z-\by$ lead to the final result
778: \beq
779: Z(y,u)=1-\frac{1}{\rho}+\frac{1}{\rho}
780: \exp\Big(g_i\rho
781: \frac{(u-1)e^{\by}}{u-(u-1)e^{\by}}
782: \Big).
783: \eeq
784: The probabilities should be obtained by developing this expression around $u=0$, which is
785: not easily done in the general form.
786:
787: Now let us briefly comment on the case of the subcritical pomeron
788: $\mu=-\epsilon<0$.
789: We again rescale
790: \beq
791: y=y'/\epsilon
792: \eeq
793: and instead of (\ref{cn}) put
794: \beq
795: c_n(y')=\frac{1}{n!}\rho^{1-n}\nu'_n(y')
796: \eeq
797: where now $\rho=-\epsilon/\lambda<0$.
798: The resulting equations for $\nu'_n$ are
799: \beq
800: \frac{d\nu'_n(y')}{d y'}=
801: -n\nu'_n(y')+n(n-1)\nu'_{n-1}(y').
802: \eeq
803: As we see they become similar to the equations for $P_n$
804: in the case of $\mu>0$ and in fact coincide with them for
805: the quantities $\tilde{\nu}_n=\nu'_n/n!$.
806: This means that for negative $\mu$ the relation between the
807: RFT and reaction-diffusion processes is different from the case of
808: positive $\mu$. In fact the interpretation of the
809: amplitude in terms of $P_n$ and $\nu_n$ interchanges: the
810: rescaled coefficients $c_n$ in the amplitude give directly the
811: probabilities to find a given number of pomerons and their
812: multiple moments have to be defined in terms of these quantities
813: in the way similar to the construction of probabilities
814: from their moments for $\mu>0$.
815: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
816: \section{Big loops}
817: One can use the Hamiltonian approach to find the contribution of
818: loops with the maximal extention in rapidity, similar to the
819: approach first taken in~\cite{salam} in the fully dimensional
820: dipole picture. This issue has been also investigated for a different model
821: with a specific quartic interaction where an exact solution can be built~\cite{levin1}.
822: To this end we first split the evolution in two parts:
823: \[
824: iA_{fi}(y)=\langle F^*_f(-v)e^{-Hy}F_i(u)\rangle\]\beq=
825: \langle F^*_f(-v)e^{-Hy/2}e^{-Hy/2}F_i(u)\rangle=
826: \langle \Big(e^{-H^{\dagger}y/2}F_f(u)\Big)^\dagger
827: e^{-Hy/2} F_i(u)\rangle.
828: \eeq
829: Next we neglect fusion of pomerons during the first part of the evolution
830: and merging of them during the second:
831: \beq
832: iA_{fi}(y)=\langle \Big(e^{-H_{fan}^{\dagger}y/2}F_f(u)\Big)^\dagger
833: e^{-H_{fan}y/2} F_i(u)\rangle.
834: \label{amplBL}
835: \eeq
836: Here $H_{fan}$ is given by Eq.(\ref{Hfan}) and $H_{fan}^{\dagger}$ differs from
837: it by the sign of $\lambda$.
838: This approximation takes into account loops which are obtained by joining
839: at center rapidity the two sets of fan diagrams going from
840: the target and projectile.
841:
842: Both operators inside the vacuum average are known and given by Eqs.
843: (\ref{fansol1}) or (\ref{fansolhA}) with the understanding that for the solution with
844: $H_{fan}^{\dagger}$ one has to change the sign of $\lambda$.
845: To simplify we restrict to the lowest order in both coupling constants
846: $g_{i(f)}$. Separating $g_ig_f$ we then obtain the pomeron Green function
847: with loops of the maximal extention.
848: In this case the right factor in (\ref{amplBL}) is given by
849: \beq
850: e^{-H_{fan}y/2}u=\frac{ue^{\mu y/2}}{1+\frac{u}{\rho}
851: \Big(e^{\mu y/2}-1\Big)}
852: \eeq
853: and the right factor is
854: \beq
855: \Big(e^{-H_{fan}^{\dagger}y/2}u\Big)^\dagger=
856: \Big[\frac{ue^{\mu y/2}}{1-\frac{u}{\rho}
857: \Big(e^{\mu y/2}-1\Big)}\Big]^\dagger=
858: \frac{-ve^{\mu y/2}}{1+\frac{v}{\rho}
859: \Big(e^{\mu y/2}-1\Big)}.
860: \eeq
861: So the pomeron Green function is
862: \beq
863: G(y)=-\langle\frac{ve^{\mu y/2}}{1+\frac{v}{\rho}
864: \Big(e^{\mu y/2}-1\Big)}\frac{ue^{\mu y/2}}{1+\frac{u}{\rho}
865: \Big(e^{\mu y/2}-1\Big)}\rangle.
866: \eeq
867: Its calculation is trivial.
868: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
869: Denoting
870: \beq
871: b=\frac{1}{\rho}\Big(e^{\mu y/2}-1\Big)
872: \eeq
873: we present the left operator as
874: \beq
875: \frac{ve^{\mu y/2}}{1+\frac{v}{\rho}
876: \Big(e^{\mu y/2}-1\Big)}=e^{\mu y/2}\frac{v}{1+bv}=
877: e^{\mu y/2}\frac{1}{b}\Big(1-\frac{1}{b}\frac{1}{v+1/b}\Big).
878: \eeq
879: Action of the unit operator gives zero so that we are left with
880: \beq
881: G(y)=e^{\mu y}\frac{1}{b^2}\langle \frac{1}{v+1/b}\frac{u}{1+bu}\rangle.
882: \eeq
883: Now we present
884: \beq
885: \frac{1}{v+1/b}=\int_0^\infty dxe^{-x(v+1/b)}
886: \eeq
887: and, since operator $e^{-xv}$ just substitutes $u$ by $x$ in the vacuum average,
888: obtain
889: \beq
890: G(y)=e^{\mu y}\frac{1}{b^2}\int_0^{\infty}dx e^{-x/b}\frac{x}{1+bx}=
891: e^{\mu y}\frac{1}{b^2}\Big(1+\frac{1}{b^2}e^{1/b^2}{\rm Ei}\,
892: (-1/b^2)\Big).
893: \label{pomBLsol}
894: \eeq
895: In the asymptotic limit $y\to\infty$, in this approximation,
896: the propagator tends to a constant value
897: \beq
898: G(y)_{y\to\infty}\simeq \rho^2\frac{e^{\mu y}}{\Big(e^{\mu y/2}-1\Big)^2}
899: \simeq \rho^2.
900: \eeq
901: The loops tamed its initial exponential growth as $e^{\mu y}$ illustrating the
902: saturation phenomenon in the old Gribov's RFT with the supercritical
903: pomeron.
904:
905: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
906: Note that the same result can be obtained by developing in the number
907: of interactions~\cite{salam,Kovchegov}. We find worth to remind to the reader
908: how one proceeds in this case. In fact we can expand both wave
909: funcions in powers of $u$ and $v$ respectively as in (\ref{powerexp}).
910: Obviously only terms with equal numbers of $u$ and $v$ contribute with
911: \beq
912: \langle v^nu^n\rangle=(-1)^n n!.
913: \eeq
914: So we get
915: \beq
916: G(y)=-\sum_{n=1}(-1)^n n!c_n^2(y/2),
917: \label{Gcn}
918: \eeq
919: where from (\ref{fansolhA}) we find
920: \beq
921: c_n(y/2)=e^{\mu y/2}(-b)^{n-1}.
922: \label{cnyhalf}
923: \eeq
924: We find a divergent series
925: \beq
926: G(y)=-\frac{1}{b^2}e^{\mu y}\sum_{n=1}(-b^2)^n n!
927: \eeq
928: which is however Borel summable. Presenting
929: \beq
930: n!=\int_0^{\infty}dt e^{-t} t^n
931: \eeq
932: we get
933: \beq
934: G(y)=-\frac{1}{b^2}e^{\mu y}\int_0^\infty dt e^{-t}\sum_{n=1}(-b^2t)^n
935: =e^{\mu y}\int_0^\infty dt e^{-t}\frac{t}{1+b^2t}.
936: \eeq
937: This is the same expression (\ref{pomBLsol}) which we obtained before.
938: So these simple but not very rigorous manipulations give the correct answer.
939:
940: It is instructive to express the same result in terms of the probabilities
941: $P_n$. Actually it is a straightforward task. Rescaling the coefficients
942: $c_n$ in (\ref{Gcn}) we rewrite this expression in terms of multiple probability
943: moments
944: $\nu_n$
945: \beq
946: G(y)=-\rho^2\sum_{n=1}\Big(-\frac{1}{\rho^2}\Big)^n\nu_n^2(y/2)\frac{1}{n!}.
947: \label{Gnun}
948: \eeq
949: Now we express the multiple moments in terms of the probabilities according to
950: (\ref{probevol}) to finally obtain
951: \beq
952: G(y)=\sum_{k,l=1}P_k(y/2)P_l(y/2)F_{kl},
953: \eeq
954: where $F_{kl}$ are coefficients independent of energy
955: \beq
956: F_{kl}=\sum_{n=1}n! C_k^nC_l^n\Big(-\frac{1}{\rho^2}\Big)^{n-1}.
957: \eeq
958:
959: It is easy to find the asymptotical values of $F_{kl}$ at large $k$ and $l$.
960: Then we can approximate
961: \beq
962: C_k^n\simeq \frac{k^n}{n!},\ \ n<<k\to\infty.
963: \eeq
964: Under this approximation
965: \beq
966: F_{kl}=-\rho^2\sum_{n=1}\frac{1}{n!}\Big(-\frac{kl}{\rho^2}\Big)^n=
967: \rho^2\Big(1-e^{-kl/\rho^2}\Big).
968: \eeq
969: Then we get from (\ref{Gnun})
970: \beq
971: G(y)=\rho^2\Big(1-\sum_{k,l=1}P_k(y/2)P_l(y/2)e^{-kl/\rho^2}\Big)
972: \eeq
973: At high $y$ and fixed $n$ the probabilities $P_n(y)$ become independent of
974: $n$ and equal to $e^{-\mu y}$
975: (see Eq. (\ref{PnhA})). Then we get
976: \beq
977: G(y)=\rho^2\Big(1-e^{-\mu y}\sum_{k,l=1}e^{-kl/\rho^2}\Big),
978: \eeq
979: with the correct limiting value $\rho^2$ at $y\to\infty$.
980: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
981: \section{Perturbative analysis for small $\lambda$ and PT-symmetry}
982:
983: Having in mind possible generalizations to the realistic two-dimensional world,
984: in this section we study perturbative treatment of the model at
985: small values of $\lambda$. As follows from the previous studies
986: ~\cite{jengo,amati2,ciafaloni1}
987: actually the point $\lambda=0$ corresponds to an essential singularity
988: in the Hamiltonian spectrum corresponding to the existence of a low-
989: lying excited state with a positive energy
990: \beq
991: \Delta E=\frac{\mu^2}{\sqrt{2\pi} \lambda} e^{-\rho^2/2}.
992: \label{DeltaE}
993: \eeq
994: This states dominates evolution at very high rapidities.
995: However at earlier stages of evolution this unperturbative component
996: may play a secondary role as compared to purely perturbative contributions.
997: We also point out in this section some general properties of the Hamiltonian $H$
998: given by Eq. (5),which
999: belongs to a class of non-Hermithean Hamiltonians widely studied in
1000: literature.
1001:
1002: Inspecting the Hamiltonian $H$ we can see that even
1003: if it is not Hermithean, it describes a PT-symmetric system
1004: evolving in the imaginary time. This property has been intensively studied mainly for
1005: anharmonic oscillators \cite{bender}.
1006: Indeed $H$ is invariant under the product of parity and ''time''-reversal
1007: transformations: $\phi \to -\phi$, $\phi^\dagger \to - \phi^\dagger$
1008: and $i \to -i$.
1009:
1010: As the transformation to the form (42) has shown, the spectrum of $H$ is real.
1011: This fact is
1012: true whenever for a
1013: diagonalizable operator $H$ certain conditions are fullfilled \cite{mostafazadeh}.
1014: The most important is that $H$ should be Hermithean with respect to some
1015: positive-definite scalar product $\langle \cdot | \cdot \rangle_\eta =
1016: \langle \cdot | \eta \cdot \rangle$,
1017: constructed using a positive-defined metric operator $\eta=e^{-Q}$
1018: with Hermithean $Q$.
1019: This defines the $\eta$-pseudo-Hermiticity property of $H$, which
1020: may be conveniently written as
1021: \beq
1022: H^\dagger=\eta\, H \,\eta^{-1} = e^{-Q} H \, e^Q.
1023: \label{pseudoH}
1024: \eeq
1025: Considering the set all positive-definite metric operators, it is
1026: easy to see that the parity operator $P$ belongs to it.
1027: Then one can define a $C$ operator by $C=P^{-1} \eta= \eta^{-1} P$ which
1028: commutes with the Hamiltonian $H$ and the $PT$ symmetry generator.
1029: In this formalism one should therefore introduce as a scalar product the
1030: CPT-inner scalar product
1031: \beq
1032: \langle \Psi_1 |\Psi_2 \rangle_{CPT}=\langle \Psi_1|e^{-Q}\Psi_2\rangle
1033: =\langle e^{-Q/2}\Psi_1|e^{-Q/2}\Psi_2\rangle
1034: \eeq
1035:
1036: The $\eta$ operator is in general unique up to symmetries of $H$. It
1037: describes the nature of the physical Hilbert space, the observables being
1038: $\eta$-pseudo-Hermithean operators.
1039: Using a symmetric (second) form of the scalar product (105)
1040: we define
1041: \beq
1042: h=e^{-Q/2} H \,e^{Q/2}\,,
1043: \label{hdef}
1044: \eeq
1045: which is Hermithean with respect to the scalar product in this form.
1046: The eigenvalues of $h$ and $H$ obviously coincide.
1047:
1048: The operator $Q$ can be
1049: determined perturbatively.
1050: In full generality we write
1051: \beq
1052: H=H_0 +\lambda H_1
1053: \eeq
1054: and in the the perturbative expansion $Q=\lambda Q_1 + \lambda^3 Q_3 + \cdots$
1055: we may derive the terms $Q_i$ using
1056: Eq. (\ref{pseudoH}) and
1057: matching the terms at each order in $\lambda$. Here we shall restrict to the
1058: first two orders in perturbation theory
1059: \bea
1060: &&\left[ H_0,\frac{Q_1}{2} \right]=-H_1 \nonumber \\
1061: &&\left[ H_0,\frac{Q_3}{2} \right]=
1062: - \frac{1}{3} \left[\left[ H_1,\frac{Q_1}{2} \right],\frac{Q_1}{2}\right]\, .
1063: \eea
1064: Inserting this into (106) for for $h=h^{(0)}+\lambda^2
1065: h^{(2)}+\lambda^4 h^{(4)}+\cdots$
1066: one then obtains
1067: \bea
1068: && h^{(0)}=H_0 \nonumber \\
1069: && h^{(2)}=\frac{1}{2} \left[H_1,\frac{Q_1}{2}\right] \nonumber \\
1070: && h^{(4)}=\frac{1}{2} \left[H_1,\frac{Q_3}{2}\right]+
1071: \frac{1}{8} \left[ \left[H_0,\frac{Q_3}{2}\right],\frac{Q_1}{2}\right] \,.
1072: \label{hpert}
1073: \eea
1074:
1075: Let us now use these results for the system under investigation.
1076: We define the operators
1077: \beq
1078: N=\phi^\dagger \, \phi \quad,\quad R_n=N^n \phi+\phi^\dagger N^n \quad,\quad
1079: S_n=N^n \phi-\phi^\dagger N^n
1080: \eeq
1081: and write $H_0=-\mu N$ and $H_1=i R_1$. One easily finds that
1082: \beq
1083: \frac{Q_1}{2}=-\frac{i}{\mu} S_1 \quad , \quad
1084: \frac{Q_3}{2}=\frac{4 i}{\mu^3}\left( S_2 + \frac{1}{3} S_1 \right)
1085: \eeq
1086: Inserting this expressions into Eq.~(\ref{hpert}) one finally obtains
1087: for $h$ up
1088: to order $\lambda^4$:
1089: \beq
1090: h= - \mu \left[ N + \frac{\lambda^2}{\mu^2} \left( 3N^2 -N\right)
1091: + \frac{\lambda^4}{\mu^4}\left(-12
1092: N^3+6N^2-2N+\Delta h^{(4)}\right) +{\cal O}
1093: \left(\frac{\lambda^6}{\mu^6}\right)\right]\,,
1094: \label{Epert}
1095: \eeq
1096: where
1097: \beq
1098: \Delta h^{(4)}=\frac{5}{2}\left(N(N+1)\phi^2+\phi^{\dagger 2}
1099: N(N+1)\right)
1100: \eeq
1101: is a term not diagonal in the number operator $N$ and thus contributes to the
1102: energy eigenvalues only at a higher order in $\lambda/\mu$.
1103:
1104:
1105: Note that the transformation (\ref{iengoform}) for positive $u$ and $\lambda$
1106: leads to Hamiltonian (\ref{Hiengoform}), which has all eigenvalues positive.
1107: In contrast the perturbative eigenvalues are seemingly non-positive
1108: for small enough $\lambda$ and certainly such at $\lambda=0$.
1109: This shows that the perturbative series is not convergent.
1110: Note however that it provides a reasonable asymptotic expansion
1111: at small values of $\lambda$ for not very high rapidities and
1112: describes well the initial period of evolution when the amplitudes rise.
1113: As an illustration, using this perturbatively approximated spectrum one may compute
1114: the pomeron propagator at two loops using
1115: \beq
1116: G(y)=\langle 0| \phi\, e^{-y H}\, \phi^\dagger |0\rangle=
1117: \langle 0| \phi\, e^{Q/2}e^{-y h} e^{-Q/2}\, \phi^\dagger |0\rangle \,,
1118: \eeq
1119: expanded in the basis $|n \rangle_h$ of the eigenstates of $h$ ($h |n\rangle_h = E_n
1120: |n\rangle_h$).
1121: Comparison to the exact numerical evolution in the next section
1122: shows a perfect agreement up to some
1123: value of $\mu y$, much below the saturation region, beyond which the
1124: perturbative approximation, which continues to grow in rapidity,
1125: fails. In particular for $\lambda/\mu=1/10$ the perturbative approximation is
1126: very good up to $\mu \,y\le 3$.
1127:
1128: In the limit $\lambda\to 0$ one finds that the initial period of evolution, when
1129: the amplitude grows, extends to infinitely high rapidities
1130: and the propagator continuosly passes into the free one, growing as $\exp(\mu y)$.
1131: So in spite of the essential singularity at $\lambda=0$, the amplitudes seem
1132: to have a well defined limit as $\lambda\to 0$, when they
1133: continuously pass into free ones.
1134:
1135: One might expect reasonable to apply a similar perturbative approach for some
1136: interval of rapidity also in the case for non zero transverse dimensions.
1137: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1138: \section{Numerical calculations}
1139: As mentioned in the introduction,
1140: most of the qualitative results in the toy model were obtained thirty
1141: years ago ~\cite{amati1,alessandrini, jengo,amati2,ciafaloni1,ciafaloni2}.
1142: They were restricted to the case of very
1143: small triple pomeron coupling $\lambda$, when the dynamics becomes
1144: especially transparent and admits asymptotic estimates. However
1145: the smallness of $\lambda$ in the zero-dimensional case actually does not
1146: correspond to the physical situation in the world with two transverse
1147: dimensions. If the transversce space is approximated by a
1148: two-dimensional lattice with intersite distance $a$ then
1149: the effective coupling constant in the zero-dimensional
1150: world is inversely proportional to $a$ and thus goes to infinity in the
1151: continuum space (see ~\cite{amati2}). So the behaviour of the model at
1152: any values of $\mu$ and
1153: $\lambda$ has a certain interest. In particular it may be interesting
1154: to know the behaviour at $\mu\to 0$, since the functional integral
1155: defining the
1156: theory for $\mu<0$ diverges at $\mu>0$ and one can expect a certain
1157: singularity at $\mu=0$. The validity of different approximate methods to
1158: find the solution
1159: may also be of interest in view of applications to more realistic models.
1160:
1161: Modern calculational facilities allow to find solutions for the model
1162: at any values of $\mu$ and $\lambda$ without any difficulty.
1163: Our approach was to evolve the wave function in rapidity directly from
1164: its initial value at $y=0$ using Eq. (\ref{evol1}) by the Runge-Cutta technique.
1165: This straightforward approach proved to be simple and powerful enough to
1166: give reliable results for a very short time. True, the step in rapidity
1167: had
1168: to be chosen rather small and correlated with the step in $u$.
1169: In practice we chose the initial interval of $u$ as
1170: \beq
1171: 0<u<20
1172: \eeq
1173: divided in 2000 points. The corresponding step in rapidity had to be
1174: taken not greater than $2.5\cdot 10^{-3}$
1175:
1176: Presenting our results we have in mind that amplitudes depend
1177: not only on the dynamics but also on the two coupling constants
1178: $g_{i(f)}$.
1179: To economize we therefore restrict ourselves to show either
1180: the full pomeron propagator
1181: \beq
1182: P(y)=-\langle ve^{-Hy}u\rangle
1183: \eeq
1184: or the amplitude for the process which mimics hadron-nucleus scattering
1185: \beq
1186: A(y,g)=-\langle e^{-gv}e^{-Hy}u\rangle
1187: \eeq
1188: (with $g_f\equiv g>0$ and $g_i=1$).
1189:
1190: \subsection{Pomeron propagator}
1191: Straightforward evolution according to Eq. (\ref{evol1}) from the initial
1192: wave function $F_i(0,u)=u$ gives the results for the pomeron propagator
1193: shown in Fig. 1 for different sets of $\mu$ and $\lambda$.
1194: We recall that the free propagator is just
1195: \beq
1196: P_0(y)=e^{\mu y}.
1197: \label{freepom}
1198: \eeq
1199: Inspection of curves in Fig. 1 allows to make the following conclusions.
1200:
1201: 1) Inclusion of the triple-pomeron interaction of any strength
1202: makes the propagator fall with rapidity. This fall is the stronger
1203: the larger the coupling.
1204:
1205: 2) At very small values of the coupling ($\rho\sim 10$ or greater)
1206: the fall is not felt at maximal rapidity $y=50$ chosen for the evolution.
1207: This is in full accord with the predictions of earlier studies
1208: ~\cite{jengo,amati2,ciafaloni1},
1209: in which it
1210: was concluded that the behaviour should be $\sim \exp(-\Delta E y)$
1211: with $\Delta E$ given by (\ref{DeltaE}).
1212: With $\mu=1$ and $\rho=10$ one finds $\Delta E \sim 10^{-21}$, so that the fall
1213: of the propagator should be felt only at fantastically large rapidities.
1214:
1215: 3) However already with $\mu=1$ and $\lambda=1/3$ the decrease of the
1216: propagator with the growth of $y$ is visible (see the non-logarithmic
1217: plot in Fig. 3).
1218:
1219: 4) With small values of $\mu$ at $\lambda=1$ the propagator goes down
1220: practically as $\exp(-y)$, so that the triple pomeron coupling takes the
1221: role of the damping mechanism
1222:
1223: 5) No singularity is visible at $\mu=0$ and fixed
1224: $\lambda=1$.
1225: The curves with very small $\mu=\pm 10^{-4}$ look completely identical.
1226: Of course this does not exclude discontinuities in higher derivatives,
1227: which we do not see numerically.
1228:
1229: We next compared our exact results with two approximate estimates.
1230: The first is the propagator obtained in the approximation of
1231: 'big loops', Eq. (\ref{pomBLsol}). The second is the asymptotic expression at small
1232: $\lambda$ obtained in earlier studies
1233: \beq
1234: P(y)=\rho^2e^{- y\Delta E}
1235: \label{asymptpom}
1236: \eeq
1237: with $\Delta E$ given by (\ref{DeltaE}). The results for $\mu=1$ and $\lambda
1238: =1/10$
1239: and 1/3 are shown in Figs. 2 and 3 respectively. As one observes, both
1240: approximations are not quite satisfactory.
1241: For quite small $\lambda=1/10$ the asymptotic limit, which we followed up to
1242: $y\sim 50$, is
1243: correctly reproduced
1244: in both approximations. However the exact propagator seems to reach this
1245: limit at still higher values of $y$ (to start falling at fantastically
1246: large $y\sim 10^{20}$). The behaviour at small values of $y$ is of course
1247: not described by (\ref{asymptpom}) at all. The 'big loop' approximation works better
1248: and describes the rising part of the curve with an error of less than 10\%,
1249: which goes down to 2\% for $y \ge 15$.
1250: At larger values of $\lambda$ the situation becomes worse. As seen from
1251: Fig. 3 ($\lambda = 1/3$) both approximations work very poorly.
1252: The 'big loop' approximation fails to reproduce both the magnitude of
1253: the propagator and its fall at large $y$. It describes the propagator
1254: more or less satisfactorily only at small values of $y\leq 2$.
1255: The approximation (\ref{asymptpom}) describes the trend of the propagator
1256: at higher $y$ better but also fails to reproduce its magnitude.
1257: For still higher values of $\lambda$ both approximations do not work at
1258: all.
1259: So our conclusion is that both approximations are applicable only at quite
1260: small values of $\lambda/\mu<1/5$ and that the approximation of 'big loop' is
1261: better, since it also reproduces the growth of the propagator
1262: at smaller $y$ from its initial value $P(0)=1$ to its asymptotic value
1263: $\rho^2$ at maximal rapidities considered.
1264:
1265: In Fig. 4 we compare the exact pomeron propagator with the calculations based on the
1266: perturbative expansion (up to two loops) and with the free propagator (\ref{freepom}).
1267: One observes that there exists a region of intermediate values of $\mu y$,
1268: from 1 to approximately 3 where, on the one hand, the influence of loops is already
1269: noticeable and, on the other hand, the perturbative approach works with a reasonable
1270: precision.
1271:
1272: \subsection{hA amplitude}
1273: In Fig. 5 we show our results for the amplitude $A(y,g)$ at
1274: comparatively large rapidity $y=10$ as a function of coupling $g$ to the
1275: target for different sets of $\mu$ and $\lambda$. This dependence mimics
1276: that of the realistic hA amplitude on the atomic number of the nucleus.
1277: In all cases except for $(\mu,\lambda)=(-1,1/10)$ the amplitude $A(g)$
1278: rapidly grows as
1279: $g$ rises from zero to approximately unity and then continues to
1280: grow very
1281: slowly at higher values of $g$ clearly showing signs of saturation. To
1282: make this
1283: behaviour visble we use the log-log plot. In the exceptional case
1284: $(\mu,\lambda)=(-1,1/10)$ one expects a similar saturation but at higher
1285: values of $g$, as will become clear in the following.
1286:
1287: We again compare our exact results with predictions of pure fan
1288: approximation, Eq.(\ref{fansolhA}) with $g_i=1$ and $u=g$, and the asymptotic
1289: formula similar to Eq. (\ref{asymptpom})
1290: \beq
1291: A(y,g)=\rho\Big(1-e^{-g\rho}\Big)e^{-y\Delta E}.
1292: \eeq
1293: The results are
1294: shown in Figs. 6,7 and 8 for
1295: $(\mu,\lambda)=(1,1/10), (1,1/3)$ and (-1,1/10) respectively.
1296: With $\mu$ positive the effect of loops is considerable, so that pure fan
1297: model gives a bad description, especially at low values of $g$ and not very
1298: small values of $\lambda$.
1299: The asymptotic formula (120) works much better. At $\mu=1$ and
1300: $\lambda=1/10$ it gives very good results, as seen from Fig. 6. However at
1301: larger values of $\lambda/\mu$ its precision rapidly goes down.
1302: The case $\mu=-1,\lambda=1/10$ illustrated in Fig. 8 was chosen to see the
1303: influence of loops for the subcritical pomeron, where such influence
1304: should be minimal. In this case the pure fan formula gives
1305: the amplitude
1306: \beq
1307: A(y=10,g)=\frac{ge^{-y}}{1+10g(1-e^{-y})}\Big|_{y=10}
1308: \eeq
1309: which grows linearly with $g$ up to values around $g=10$ and only
1310: then saturates at the value $0.1e^{-10}$. This explains a somewhat
1311: exceptional behaviour of $A(y=10,g)$ for such $\mu$ and $\lambda$.
1312: As follows from Fig. 8 the influence of loops is in fact small. However
1313: it is greater than $\sim\lambda^2$=1\% as one could expect on simple
1314: estimates of a single loop contribution. In fact the pure fan formula
1315: overshoots the exact result by about 20\%.
1316:
1317: \section{Conclusions}
1318: As mentioned, the zero-dimensional RFT was studied in detail some thirty
1319: years ago. Our aim was to clarify some mathematical aspects of the model and also to
1320: study it at different values of the triple pomeron coupling and not only at small ones.
1321:
1322: We have found that the model can be consistently formulated in terms of the
1323: quantum Hamiltonian. This formulation is valid both for negative and positive values
1324: of $\mu$, in contrast to the functional approach which does not admit positive $\mu$.
1325: The Hamiltonial formulation does not need to introduce a scalar product in
1326: the representation in which the creation operators $u$ are diagonal. The wave functions
1327: seem to be not integrable in the whole complex $u$ plane. Rather the standard scalar
1328: product should be used in which the creation and annihilation operators are expressed
1329: by Hermithean operators.
1330:
1331: A clear physical content of the theory is well seen after transformation to an
1332: Hermithean Hamiltonian made in ~\cite{jengo}. However such transformation can only be
1333: done separately for positive and negative $u$, leading to two branches of the spectrum
1334: for the
1335: initial Hamiltonian, which include both positive anfd negative eigenvalues.
1336: The choice between two branches is determined by the signs of $\lambda$ and
1337: external coupling constants.
1338:
1339: For $\mu>0$ a simple case of pure fan diagrams admits an explicit
1340: solution and reinterpretation in terms of
1341: fashionable reaction-diffusion approach. Using this one can obtain an approximate
1342: expression for the pomeron propagator with loops of the largest extension in rapidity
1343: taken ito account, which gives a reasonable approximation at small values of $\lambda$.
1344:
1345: Finally we performed numerical calculations of the amplitudes.
1346: They show that in all cases the triple pomeron interaction makes amplitudes fall
1347: at high rapidities. This fall starts later for smaller $\lambda$ and at very
1348: small $\lambda$ begins at asymptotically high rapidities (for $\lambda/\mu<1/4$
1349: it is noticeable only at $\mu y>100$). At small $\lambda$ the behaviour in $y$ is well
1350: described by formulas obtained in earlier studies.
1351: However when $\lambda$ is not so small all approximate formulas work poorly.
1352: Our numerical calculations have also shown that there is no visible singularity at
1353: $\mu=0$, in spite of the fact that the functional formulation meets with
1354: a divergence at this point.
1355:
1356: We have proposed a new method for the perturbative treatment of the theory,
1357: which may have appliciations to more sophisticated models in the world of more
1358: dimensions
1359:
1360: \section{Acknowledgments}
1361: M.A.B. and G.P.V. gratefully acknowledge the hospitality of the II
1362: Institut for Theoretical Physics of the University of Hamburg, where part of
1363: this work was done. M.A.B. also thanks the Bologna Physics department and
1364: INFN for hospitality. The authors thank J. Bartels,
1365: S.Bondarenko, L.Motyka and A.H. Mueller for very interesting and constructive
1366: discussions.
1367: G.P.V. thanks the Alexander Von Humboldt foundation for partial support.
1368: This work was also partially supported by grants RNP 2.1.1.1112 and RFFI 06-02-16115a
1369: of Russia.
1370:
1371: \begin{thebibliography}{99}
1372: %
1373: \bibitem{AHM} A.H.Mueller, Nucl. Phys. {\bf B 437} (1995) 107.
1374: %
1375: \bibitem{amati1} D.Amati, L.Caneshi and R.Jengo, Nucl. Phys. {\bf B 101}
1376: (1975) 397.
1377: %
1378: \bibitem{alessandrini} V.Alessandrini, D.Amati and R.Jengo, Nucl. Phys.
1379: {\bf B 108} (1976) 425.
1380: %
1381: \bibitem{jengo} R.Jengo, Nucl. Phys. {\bf B 108} (1976) 447.
1382: %
1383: \bibitem{amati2} D.Amati, M.Le Bellac, G.Marchesini and M.Ciafaloni, Nucl.
1384: Phys. {\bf B 112} (1976) 107.
1385: %
1386: \bibitem{ciafaloni1} M.Ciafaloni, M. Le Bellac and G.C.Rossi, Nucl. Phys.
1387: {\bf B 130} (1977) 388.
1388: %
1389: \bibitem{ciafaloni2} M.Ciafaloni, Nucl. Phys. {\bf B 146} (1978) 427.
1390: %
1391: \bibitem{boreskov} K.G.Boreskov, arXiv: hep-ph/0112325.
1392: %
1393: \bibitem{bondarenko} S.Bondarenko, L.Motyka, A.H.Mueller, A.I.Shoshi
1394: and B.-W.Xiao, hep-ph/0609213.
1395: %
1396: \bibitem{levin1} M.~Kozlov and E.~Levin,
1397: %``Solution for the BFKL pomeron calculus in zero transverse dimensions,''
1398: Nucl.\ Phys.\ A {\bf 779} (2006) 142.
1399: %%CITATION = HEP-PH 0604039;%%
1400: %
1401: \bibitem{levin2} M.Kozlov, E.Levin, V.Khachatryan and J.Miller,
1402: arXiv: hep-ph/0610084.
1403: %
1404:
1405: \bibitem{schwimmer} A.Schwimmer, Nucl. Phys. {\bf B 94} (1975) 445.
1406: %
1407: \bibitem{salam} G.P.Salam, Nucl. Phys. {\bf b 461} (1996) 512.
1408: %
1409: \bibitem{Kovchegov}
1410: Y.~V.~Kovchegov,
1411: %``Inclusive gluon production In high energy onium onium scattering,''
1412: Phys.\ Rev.\ D {\bf 72} (2005) 094009.
1413: %%CITATION = HEP-PH 0508276;%%
1414: %
1415: \bibitem{bender}
1416: C.~M.~Bender and S.~Boettcher,
1417: %``Real Spectra in Non-Hermitian Hamiltonians Having PT Symmetry,''
1418: Phys.\ Rev.\ Lett.\ {\bf 80}, 5243 (1998);
1419: %%CITATION = PHYS-ICS 9712001;%%
1420:
1421: C.~M.~Bender, D.~C.~Brody and H.~F.~Jones,
1422: %``Complex Extension of Quantum Mechanics,''
1423: Phys.\ Rev.\ Lett.\ {\bf 89}, 270401 (2002)
1424: [Erratum-ibid.\ {\bf 92}, 119902 (2004)].
1425: %%CITATION = QUANT-PH 0208076;%%
1426: %
1427: \bibitem{mostafazadeh}
1428: A.~Mostafazadeh,
1429: %``Pseudo-Hermiticity versus PT-Symmetry III: Equivalence of
1430: %pseudo-Hermiticity and the presence of anti-linear symmetries,''
1431: J.\ Math.\ Phys.\ {\bf 43}, 3944 (2002);
1432: %%CITATION = MATH-PH 0203005;%%
1433:
1434: A.~Mostafazadeh,
1435: %``PT-Symmetric Cubic Anharmonic Oscillator as a Physical Model,''
1436: J.\ Phys.\ A {\bf 38}, 6557 (2005)
1437: [Erratum-ibid.\ A {\bf 38}, 8185 (2005)].
1438: %%CITATION = QUANT-PH 0411137;%%
1439: %
1440: \end{thebibliography}
1441: %
1442: %\newpage
1443: %
1444: %Fig.1
1445: \begin{figure}[ht]
1446: \epsfxsize 4in
1447: \centerline{\epsfbox{fig1.eps}}
1448: \caption{The full pomeron propagator as a function of rapidity for
1449: different sets of $\mu$ and $\lambda$ indicated in the figure}
1450: \label{Fig1}
1451: \end{figure}
1452: %
1453: %Fig.2
1454: \begin{figure}[ht]
1455: \epsfxsize 4in
1456: \centerline{\epsfbox{fig2.eps}}
1457: \caption{The full pomeron propagator as a function of rapidity
1458: for $\mu=1$ and $\lambda=0.1$
1459: (the lower curve) as compared to the 'big loop' formula Eq. (88)
1460: (the middle curve) and asymptotic expression (119) (the upper curve)}
1461: \label{Fig2}
1462: \end{figure}
1463: %
1464: %Fig. 3
1465: \begin{figure}[ht]
1466: \epsfxsize 4in
1467: \centerline{\epsfbox{fig3.eps}}
1468: \caption {The full pomeron propagator as a function of rapidity
1469: for $\mu=1$ and $\lambda=1/3$
1470: (the lower curve) as compared to the 'big loop' formula Eq. (88)
1471: (the upper curve on the right) and asymptotic expression
1472: (119) (the middle curve on the right)}
1473: \label{Fig3}
1474: \end{figure}
1475: %
1476: %Fig. 4
1477: \begin{figure}[ht]
1478: \epsfxsize 4in
1479: \centerline{\epsfbox{perturba.eps}}
1480: \caption{The Pomeron Green's function as a function of $\mu\, y$ for $\mu=1$ and $\lambda=1/10$: from
1481: bottom to top: the exact numerical evolution, the two loop
1482: perturbative approximation, the free case ($e^{y \mu}$). }
1483: \label{Fig4}
1484: \end{figure}
1485: %
1486: %Fig. 4
1487: \begin{figure}[ht]
1488: \epsfxsize 4in
1489: \centerline{\epsfbox{fig4.eps}}
1490: \caption{ The hA amplitude at $y=10$ as a function of the coupling
1491: to the target. Curves from top to bottom on the right correspond to
1492: $(\mu,\lambda)$=(1,0.1),(1,1/3), (1,1),(-1,0.1),(0.1,1),(0,1)}
1493: \label{Fig5}
1494: \end{figure}
1495: %
1496: %Fig.6
1497: \begin{figure}[ht]
1498: \epsfxsize 4in
1499: \centerline{\epsfbox{fig5.eps}}
1500: \caption{the hA amplitude $A(y,g)$ at $y=10$ for $\mu=1$, $\lambda=0.1$
1501: (the lower curve) as compared to the pure fan prediction (the upper
1502: curve) and the asymptotic expression Eq. (120) (middle curve)}
1503: \label{Fig6}
1504: \end{figure}
1505: %
1506: %Fig.7
1507: \begin{figure}[ht]
1508: \epsfxsize 4in
1509: \centerline{\epsfbox{fig6.eps}}
1510: \caption{Same as Fig. 5 for $\mu=1$ and $\lambda=1/3$}
1511: \label{Fig7}
1512: \end{figure}
1513: %
1514: %Fig. 8
1515: \begin{figure}[ht]
1516: \epsfxsize 4in
1517: \centerline{\epsfbox{fig7.eps}}
1518: \caption{The hA amplitude $A(y,g) $at $y=10$ for $\mu=-1$, $\lambda=0.1$
1519: (the lower curve) as compared to the pure fan prediction (the upper
1520: curve)}
1521: \label{Fig8}
1522: \end{figure}
1523:
1524: \end{document}
1525: