1: \documentstyle[aps,preprint,tighten,epsfig,amssymb]{revtex}
2: \begin{document}
3: %\draft
4:
5: \title{The Canonical Transformation and Duality in the 1+1 dimensional
6: $\phi^4$ and $\phi^6$ theory}
7:
8: \author{
9: Chueng-Ryong Ji$^{a}$
10: \footnote{e-mail address: ji@ncsu.edu},
11: Joon-Il Kim$^{b}$
12: \footnote{e-mail address: jikim@phya.snu.ac.kr},
13: Dong-Pil Min$^{b}$
14: \footnote{e-mail address: dpmin@phya.snu.ac.kr},
15: Andrey V. Vinnikov$^{a,b,c}$
16: \footnote{e-mail address: vinnikov@phya.snu.ac.kr}}
17:
18: \address{
19: \bigskip
20: $^a$ Department of Physics, North Carolina State University, Raleigh, NC
21: 27695-8202, USA\\
22: $^b$ School of Physics and Center for Theoretical Physics,
23: Seoul National University, Seoul 151-742, Korea\\
24: $^c$ Bogoliubov Laboratory of Theoretical Physics, JINR, 141980,
25: Dubna, Russia}
26:
27: \date{\today}
28:
29: \maketitle
30:
31: \begin{abstract}
32: We investigate the self-organizing nature of relativistic quantum field
33: theory in terms of canonical transformation and duality presenting simple
34: but explicit examples of $(\phi^4)_{1+1}$ and $(\phi^6)_{1+1}$ theories.
35: Our purpose is fulfilled by applying the oscillator representation (OR) method
36: which allows us to convert the original strong interaction theory
37: into a weekly interacting quasiparticle theory that is equivalent
38: to the original theory.
39: We discuss advantages of the OR method and compare the results with what
40: was already obtained by the method
41: of Gaussian effective potential (GEP) and the Hartree approximation (HA).
42: While we confirm that the GEP results are identical to the Hartree results
43: for the ground state energy, we found that the OR method
44: gives the quasiparticle mass different from the GEP and HA
45: results. In our examples, the self-organizing nature is revealed by
46: the vacuum energy density that gets lowered when the quasiparticles are
47: formed.
48: In the $(\phi^6)_{1+1}$ theory, we found two
49: physically meaningful duality-related quasiparticle solutions which have
50: different symmetry properties under the transition of quasiparticle
51: field $\Phi \rightarrow -\Phi$. However, these two quasiparticle
52: solutions yield the identical effective
53: potential in the strong coupling limit of the
54: original theory.
55: \end{abstract}
56:
57: PACS: 11.10.Gh, 11.10.Kk, 11.10.Lm
58:
59: %Keywords: Self-organizing Quantum Field Theory, Oscillator Representation
60:
61: \section{INTRODUCTION}
62: With the current advances of Relativistic Heavy Ion Collision(RHIC)
63: physics, there is a growing interest in discussing
64: the self-organizing nature of relativistic quantum field theory(RQFT).
65: The phase transition and the spontaneous symmetry breaking
66: anticipated to be observed in the RHIC facilities are the paramount
67: examples of the physical phenomena due to the self-organization of
68: quantum fields. Since these novel phenomena cannot be easily predicted in
69: the ordinary perturbation series, they form highly nontrivial examples
70: that ought to be analyzed by various nonperturbative methods available
71: for the RQFTs.
72:
73: Although the problem of quantization of interacting fields
74: is not yet completely solved, it appears that the duality existing in the
75: RQFTs is the key to handle the strong interaction theories.
76: According to the duality, the original particle theory of strong couplings
77: may be equivalent to the quasi-particle theory of weak couplings. As the
78: interactions get stronger, the associate quantum fluctuations get larger
79: and consequently the formation of a nontrivial vacuum
80: often accompanied by the condensation of fields
81: may be energetically more favorable than the adherence to an original trivial
82: vacuum without any condensation. The particles moving in the ambient
83: nontrivial vacuum condensates may be described by the quasi-particles
84: which carry different masses
85: compare to their original masses defined in the trivial vacuum.
86: Then, the couplings among the quasi-particles may become weak as the
87: couplings among the original particles grow strong.
88: When such duality works self-consistently within the RQFT, one may
89: be able to solve the corresponding strong interaction problem utilizing
90: the weak coupling developed for the quasi-particles.
91: For the phenomena involving only weak interactions, the perturbation
92: theory may provide systematic predictions and even some physical insight
93: how the interaction can affect the properties of the system.
94: Thus, the duality allows us to convert the original nonperturbative
95: problem into the quasiparticle perturbative problem.
96: However, the key issue here is to check if the RQFTs indeed have the
97: self-organizing characteristics that lead to the duality within
98: themselves. To investigate such nontrivial characteristics, one cannot
99: just rely on the perturbation series from the beginning but need to
100: develop a nonperurbative method which may be useful to analyze the
101: self-organizing nature of the RQFTs.
102:
103: In this work, we discuss such a nonperturbative method that appears to be
104: much simpler than other rather well-known methods such as the Hartree
105: Approximation (HA) and the Gaussian Effective Potential (GEP) method.
106: Following the nomenclature in the literature we call this method as the
107: Oscillation Representation (OR) method, although in our view all of these
108: nonperturbative methods (OR,GEP,HA) stem from the same principle of
109: quantum effective action approach\cite{CJT}.
110: The OR method was explicitly formulated by Efimov\cite{efimov} and
111: was, in fact, used earlier by Chang \cite{chang2} and
112: Magruder \cite{magruder}, though in some indirect way.
113: The method is at length described in a monograph \cite{efbook}.
114: The basic idea of the OR method is to redefine the mass of interacting field
115: relative to the free one and simultaneously introduce a shift of the field
116: quantization point leading to a nonzero value of its vacuum condensate.
117: This effect is realized in the nature represented by a spontaneous symmetry
118: breaking mechanism and thus one can expect that it may be a general
119: feature of quantum field systems with interactions.
120: However, the OR method has not been utilized as extensively as the methods
121: of GEP\cite{ft7,gep,gep4,gep6,mp} and HA\cite{Chang,hartree}. Even in the
122: 1+1 dimensional scalar field theory such as
123: the one we present in this work, not much of the OR results distinguished
124: from the GEP or HA results have been known or discussed to the best of
125: our knowledge.
126: Although the OR results for the $(\phi^4)_{1+1}$ theory were presented
127: in Ref.\cite{efbook}, the distinction between the OR and GEP (or HA)
128: results for this simple model was not pointed out at length.
129: Thus, for the present work, we analyze the details of the distinction
130: in this most simple case first and extend the application to the
131: $(\phi^6)_{1+1}$ theory:
132: \begin{eqnarray}
133: \label{lagrangian}
134: {\cal L} = \frac{1}{2} \partial_\mu \phi \partial^\mu \phi
135: -\frac{1}{2} m^2 \phi^2 - \frac{g}{4} \phi^4 - h \phi^6,
136: \end{eqnarray}
137: where the results from the variational method of the GEP have
138: also been studied \cite{gep6}.
139: The GEP is defined as
140: \begin{equation}
141: \label{gep}
142: {\bar V}_G(\phi_0) = min_\Omega <\phi_0,\Omega|{\cal H}|\Omega,\phi_0>,
143: \end{equation}
144: where ${\cal H}$ is the Hamiltonian density corresponding to
145: Eq.(\ref{lagrangian}). Here,
146: $\Omega$ is the mass parameter and $|\Omega,\phi_0>$ is a normalized
147: Gaussian wave functional centered on $\phi = \phi_0$, {\it i.e.}
148: \begin{equation}
149: <\phi_0,\Omega|\Omega,\phi_0> = 1,
150: <\phi_0,\Omega|\phi|\Omega,\phi_0> = \phi_0.
151: \end{equation}
152: As we will discuss the details in the next two sections (Sections II and
153: III), the GEP method can in principle give the two OR equations
154: that we solve simultaneously in this work. One of them is known
155: as the ${\bar \Omega}$ equation in the GEP,{\it i.e.}
156: $\partial <\phi_0,\Omega|{\cal H}|\Omega,\phi_0>/\partial \Omega = 0$
157: at $\Omega = {\bar \Omega}$, and the other OR equation is equivalent
158: to $d{\bar V}_G(\phi_0)/{d \phi_0} = 0$. However, in the GEP method,
159: these two equations are not solved simultaneously in practice.
160: It turns out that the quasiparticle mass defined by $d^2 {\bar
161: V}_G(\phi_0)/{d \phi_0^2}$ in the GEP is not same with ${\bar \Omega}$
162: which is the simultaneous solution of the quasiparticle mass from the two
163: OR equations.
164: In this paper, we therefore pay more
165: attention to the distinct aspects of the OR results compared to the GEP
166: and
167: HA results and discuss the property of the OR results which was not
168: mentioned in the more great detail before.
169:
170: Formally, the OR is settled by requiring that the Hamiltonian operator
171: should be written in terms of creation and annihilation operators
172: of an oscillator basis with an appropriate frequency and in the correct
173: form defined by\\
174: (1) all field operators in the total Hamiltonian $H = H_0 +H_I$ are
175: written in the normal ordered product,\\
176: (2) the Hamiltonian $H_0$ is quadratic over the field operators,\\
177: (3) the interaction Hamiltonian $H_I$ contains field operators in powers
178: more than two.\\
179: Using this method with the cannonical transformation of quantum fields,
180: one may allow to build a standard perturbation theory
181: for the quasiparticles in the region of the coupling constants where
182: the original theory is nonperturbative. Thus, we utilize the OR method
183: to transform the original theory which is intrinsically nonperturbative
184: to the equivalent theory of the quasiparticle
185: representation that can be solved rather easily by the perturbation
186: method.
187:
188: As we will discuss, the exact OR results include several
189: vacuum solutions yielding different mass values of the quantum
190: field for fixed values of coupling constants.
191: Among those multiple solutions, however, it is certainly not difficult
192: to select the non-trivial solutions that are physically meaningful
193: because of the duality consideration. In this work, we will thus focus
194: only those duality-related solutions which are not hindered from the
195: physical interpretation.
196: Nevertheless, we note that the OR method generates also the solutions that
197: cannot be interpreted as straightforward as the ones we focus in this
198: paper. In particular, we find a class of solution that doesn't exhibit
199: the anticipated duality property and another class of
200: solution that we call spurious in the sense that it yields
201: the pure imaginary value for the vacuum condensation.
202: We do not yet know all the physical meaning of these classes of solutions,
203: but the generation of multiple solutions are expected because the OR
204: equations are the nonlinear algebraic equations relating the quasiparticle
205: mass and the field condensation with the parameter of coupling constant
206: as we will present in details for the rest of the paper.
207:
208: In Section II, we present more details of the OR method using the
209: $\phi^4$ theory which can be easily deduced from the Lagrangian
210: density of $\phi^6$ theory by taking $h=0$. We compare the OR results
211: with the results obtained by the other available methods such as the GEP
212: and the HA as well as the available lattice result.
213: In Section III, we apply the OR method to
214: the $(\phi^6)_{1+1}$ theory and discuss the effects from the presence of
215: $\phi^6$ interaction to the duality-related vacuum solutions.
216: We analyze the quasiparticle mass and the energy density in
217: the entire domain of coupling parameter space and
218: present the classical potential for the nontrivial solutions.
219: We discuss two physically meaningful duality-related quasiparticle
220: solutions with different symmetry properties and show that they however
221: generate the identical effective potential in the strong coupling
222: limit of the original theory.
223: Summary and conclusions follow in Section IV. In the Appendix A,
224: we briefly summarize some details of the HA in the
225: $(\phi^4)_{1+1}$ theory and discuss an equivalence between the two
226: nontrivial vacuum solutions in this approximation. In the Appendix B,
227: we present a proof of unitary inequivalence between the bifurcated
228: nontrivial solutions from the OR method in the infinite volume limit.
229: In the Appendix C, we summarize the results of the spurious solutions
230: with some discussion.
231:
232: \section{THE OSCILLATOR REPRESENTATION (OR) METHOD}
233: The usual recipe to produce a quantum system comes from the correspondence
234: principle. Namely, a quantum system can be obtained from
235: its classical analog by assuming that the field is not a $c$-number
236: but an operator satisfying relevant commutation relations.
237: The oscillator representation method is a generalization
238: of this scheme. To start with, let us consider the $\phi^4$ theory
239: as a simple example.
240:
241: The Hamiltonian density of the classical field is given by
242: \begin{equation}
243: {\cal H} = \frac{1}{2}\pi^2+\frac{1}{2}(\nabla \phi)^2+\frac{1}{2}
244: m^2\phi^2+\frac{g}{4}\phi^4.
245: \end{equation}
246: To plug the quantum physics in, we postulate that the field $\phi$ and
247: the corresponding conjugate momentum $\pi$ are operators which satisfy
248: canonical equal time commutation relations
249: \begin{equation}
250: [\phi(x),\pi(y)]_{x_0=y_0}= i\delta({\bf x}-{\bf y}).
251: \label{commut}
252: \end{equation}
253:
254: The usual approach to quantize the field, {\it i.e.} to find the operators
255: $\phi$ and $\pi$ which satisfy the condition (\ref{commut}) is to start from
256: the free field, where $g$=0. Then the operators are
257: \begin{eqnarray}
258: \phi_m(x) = \int \frac{d {\bf k}}{\sqrt{2\pi}} \frac
259: {1}{\sqrt{2\omega_m({\bf k})}}
260: (a_m({\bf k})e^{ik\cdot x} + a_m^+({\bf k})e^{-ik\cdot x}), \nonumber \\
261: \pi_m(x) = \frac{1}{i}\int \frac{d {\bf k}}{\sqrt{2\pi}}
262: {\sqrt{\frac{\omega_m({\bf k})}{2}}}(a_m({\bf k})e^{ik\cdot x} - a_m^+({\bf k})
263: e^{-ik\cdot x}),
264: \label{fieldop}
265: \end{eqnarray}
266: where $k=(\omega_m,{\bf k})$, $\omega_m=\sqrt{{\bf k}^2+m^2}$ and the
267: operators of creation and annihilation of the particle of mass $m$ and
268: momentum ${\bf k}$ satisfy the commutation relations
269: \begin{eqnarray}
270: [ a_m({\bf k}),a_m^+({\bf k'})] = \delta({\bf k} -{\bf k'}), \nonumber \\
271: {[}a_m({\bf k}),a_m({\bf k'}) ] = [ a_m^+({\bf k}),a_m^+({\bf k'})] =0.
272: \label{operat}
273: \end{eqnarray}
274: With this definition the state $a_m^+({\bf k})|0\rangle$ is an eigenstate
275: of both
276: free Hamiltonian and momentum operators with corresponding eigenvalues
277: $\omega_m$ and ${\bf k}$.
278:
279: In the case of interacting fields the states $a_m^+({\bf k})|0\rangle$ are
280: not
281: already eigenstates of the full Hamiltonian. Nevertheless,
282: in the case of small interaction strength, the free representation
283: of field operators can be a good starting point to develop a perturbation
284: theory. The Hamiltonian density operator is then written as
285: \begin{equation}
286: {\cal H}\rightarrow {\cal H}_m = \frac{1}{2}\pi^2_m+
287: \frac{1}{2}(\nabla \phi_m)^2+\frac{1}{2}m^2\phi^2_m+\frac{g}{4}\phi_m^4.
288: \label{quantham}
289: \end{equation}
290:
291: Now, we use the canonical transformation of the field operators
292: to build the quasiparticle representation of the same Hamiltonian.
293: We consider the essential point of the reparameterizations,
294: {\it i.e.} the change of mass, which can affect the choice of
295: representation for the free field quantization.
296: For the free field operators, the change of mass $m\rightarrow M$ of
297: particles is described by the Bogoliubov-Valatin transformation:
298: \begin{eqnarray}
299: a_M({\bf k})=\frac{1}{2}\left( \sqrt{\frac{\omega_m({\bf k})}
300: {\omega_M({\bf k})}} + \sqrt{\frac{\omega_M({\bf k})}
301: {\omega_m({\bf k})}} \right) a_m({\bf k})-
302: \frac{1}{2}\left( \sqrt{\frac{\omega_m({\bf k})}
303: {\omega_M({\bf k})}} - \sqrt{\frac{\omega_M({\bf k})}
304: {\omega_m({\bf k})}} \right) a^+_m({\bf -k}), \nonumber \\
305: a^+_M({\bf k})=\frac{1}{2}\left( \sqrt{\frac{\omega_m({\bf k})}
306: {\omega_M({\bf k})}} + \sqrt{\frac{\omega_M({\bf k})}
307: {\omega_m({\bf k})}} \right) a^+_m({\bf k})-
308: \frac{1}{2}\left( \sqrt{\frac{\omega_m({\bf k})}
309: {\omega_M({\bf k})}} - \sqrt{\frac{\omega_M({\bf k})}
310: {\omega_m({\bf k})}} \right) a_m({\bf -k}).
311: \end{eqnarray}
312: Here, the free field operators with mass $M$ are expressed in terms of
313: the field operators with mass $m$.
314:
315: A consequence of the ambiguity of the mass definition due to interactions
316: comes from the ambiguity in the choice of the initial representation of the
317: interacting field. This was emphasized in Coleman's paper \cite{coleman}.
318: He also proposed a useful technique on how to redefine a normal
319: ordered product of any number of field operators with respect to a new
320: value of mass in the case of (1+1) scalar field theory. The redefinition
321: after the mass change is given by the formulas \cite{chang2,coleman}:
322: \begin{equation}
323: N_m\left(\frac{1}{2}\pi^2_m+\frac{1}{2}(\nabla\phi_m)^2\right)=
324: N_M\left(\frac{1}{2}\pi^2_M+\frac{1}{2}(\nabla\phi_M)^2\right)+
325: \frac{1}{8\pi}(M^2-m^2),
326: \label{normfree}
327: \end{equation}
328: \begin{equation}
329: N_m(e^{\imath \beta\phi_m})=\Bigl ( \frac{M^2}{m^2} {\Bigr )}^{\beta^2/8\pi}
330: N_M(e^{\imath \beta\phi_M}),
331: \label{normexp}
332: \end{equation}
333: where $\beta$ is some arbitrary $c$-number, $N_m$($N_M$) stands for
334: the normal ordering of an operator with respect to
335: the initial mass $m$ (the new mass $M$), $\phi_m$
336: is the free quantum field defined by Eq.(\ref{fieldop}) and $\phi_M$ is
337: the quantum field of independent quasi-particles for which
338: the representation
339: given by Eq.(\ref{fieldop}) is valid after changing $m\rightarrow M$.
340: The correspondence between the canonical transformation and the
341: renormalization group equation is discussed in Ref.\cite{efbook}.
342: %The change of normal ordering mass from $m$ to $M$ can be interpreted
343: %in terms of the renormalization group change of scale from $\mu$ to
344: %$\nu$.
345: %This translates into $m(\mu)/\mu = M(\nu)/\nu$ for the solutions found
346: %by the OR method\cite{efbook}.
347: The expression (\ref{normexp}) can be rewritten as
348: \begin{equation}
349: N_m(\phi^n_m)=n!\sum\limits^{[\frac{n}{2}]}_{j=0} \Bigl ( \frac{1}{8\pi}
350: \ln(\frac{M^2}{m^2}) \Bigr ) ^j\frac{(-1)^j}{j!(n-2j)!}N_M(\phi^{n-2j}_M),
351: \label{normord}
352: \end{equation}
353: where $\bigl [ \frac{n}{2} \bigr ] $ is the integer part of $\frac{n}{2}$.
354:
355: The most simple illustration of nontrivial solutions in this case is the 1+1
356: dimensional $\phi^4$ theory. Rewriting the Hamiltonian (\ref{quantham})
357: in terms of the quasi-particles using Coleman's formula of normal ordering
358: rearrangement, one gets
359: \begin{eqnarray}
360: N_m({\cal H}_m) &=& N_M({\cal H}_M) \nonumber \\
361: &+&\frac{1}{2}(m^2-M^2) N_M(\phi^2_M)
362: +\frac{1}{8\pi}(M^2-m^2) -\frac{m^2 t}{8\pi}
363: -\frac{3gt}{8\pi} N_M(\phi^2_M)+\frac{3gt^2}{64\pi^2},
364: \label{phim}
365: \end{eqnarray}
366: where $t=\ln\bigl ( \frac{M^2}{m^2} \bigr ) $.
367:
368: For a consistent
369: description of the dynamics in terms of the quasi-particles one
370: should provide the right form of the Hamiltonian, {\it i.e.} the free part
371: should correspond to that of the field with mass $M$ and the
372: interacting part should contain
373: only terms with powers of $\phi_M$ larger than 2 \cite{efbook}. As it can
374: be seen, the expression (\ref{phim}) itself can not provide this requirement
375: for the field with mass other than $m$. Revising just the mass does not
376: lead to a nontrivial solution and thus an additional transformation
377: is obviously required. The latter can be done by shifting the quantization
378: point of the field, which makes the quanta
379: of the field to be produced around some vacuum expectation value
380: $\langle \phi \rangle \ne 0$.
381:
382: Thus, let us now change the field variable from $\phi_M(x)$ to $\Phi_M(x)$
383: for the inclusion of the field condensation denoted by $b (=\langle
384: \phi_M \rangle)$:{\it i.e.},
385: \begin{equation}
386: \phi_M(x) = \Phi_M(x) + b.
387: \end{equation}
388: Then the Hamiltonian density is written as \cite{efbook}:
389: \begin{equation}
390: N_M({\cal H})=N_M({\cal H}_M^{right})+{\cal H}_1+\varepsilon_M,
391: \end{equation}
392: where ${\cal H}_M^{right}$ is the right form of Hamiltonian density for
393: the field with mass $M$,
394: \begin{equation}
395: {\cal H}_M^{right}=\frac{1}{2}\pi^2_M + \frac{1}{2}(\nabla\Phi_M)^2 +
396: \frac{M^2}{2}\Phi^2_M + \frac{g}{4}\Phi^4_M +
397: g b\Phi^3_M,
398: \label{phi4ham}
399: \end{equation}
400: $\varepsilon_M$ is the energy density of the vacuum for the quasiparticle
401: of mass $M$ up to the zeroth
402: order of the dimensionless effective coupling $g/M^2$,
403: \begin{equation}
404: \varepsilon_M=\frac{m^2b^2}{2}+\frac{gb^4}{4}-\frac{m^2 t}{8\pi}-\frac{3gb^2 t}{8\pi}
405: +\frac{3gt^2}{64\pi^2}+\frac{M^2-m^2}{8\pi},
406: \label{energy}
407: \end{equation}
408: and
409: \begin{equation}
410: {\cal H}_1 = \Bigl \{ \frac{1}{2}(m^2-M^2)+ \frac{3}{2}g b^2 -\frac{3gt}{8\pi}
411: \Bigr \} \Phi^2_M + \Bigl \{m^2 b +g b^3 -\frac{3g b t}{4\pi}\Bigr \} \Phi_M
412: \end{equation}
413: is a remainder. To have the Hamiltonian in the right form the remainder
414: must be equal to 0. Therefore, we have two equations of two variables,
415: $t$ and $b$, as follows:
416: \begin{equation}
417: \frac{1}{2}(1-\frac{M^2}{m^2})+ \frac{3}{2}G b^2 -\frac{3Gt}{8\pi} = 0,
418: \label{OR1}
419: \end{equation}
420: and
421: \begin{equation}
422: b \Bigl (1 +G b^2 -\frac{3G t}{4\pi} \Bigr ) = 0,
423: \label{OR2}
424: \end{equation}
425: where $G$ is the dimensionless coupling of the initial theory, {\it i.e.}
426: $G=g/{m^2}$.
427:
428: We note here the correspondence to the GEP discussed in Section I.
429: The energy density $\varepsilon_M$ given by Eq.(\ref{energy}) corresponds
430: to $<\phi_0,\Omega|{\cal H}|\Omega,\phi_0>$ presented in Eq.(\ref{gep}).
431: Minimizing $<\phi_0,\Omega|{\cal H}|\Omega,\phi_0>$ (or $\varepsilon_M$)
432: with respect to $\Omega$ (or $t$), one gets the GEP given by
433: Eq.(\ref{gep}) with the so-called ${\bar \Omega}$ equation
434: \footnote{ See
435: Eq.(4.12) of Ref.\cite{gep4} which
436: was rederived as Eq.(2.16a) of Ref.\cite{mp} using Bogoliubov
437: transformations.}
438: which is identical to Eq.(\ref{OR1}) that is one of the OR equations.
439: The other OR equation, Eq.(\ref{OR2}), can be obtained by making the
440: total (not partial) derivative of the GEP ({\it i.e.}, Eq.(4.11) of
441: Ref.\cite{gep4} or Eq.(2.16b) of Ref.\cite{mp}) with respect to the
442: condensation $\phi_0 (= b)$ and setting it to zero,
443: {\it i.e.} $d{\bar V}_G(\phi_0)/{d \phi_0} = 0$.
444: One should note, however, that the GEP is a function of
445: the global field condensate $\phi_0$ not the local field $\phi(x)$.
446: Thus, the GEP is not a part of interaction Hamiltonian describing the
447: field dynamics.
448: In our work, instead of using the GEP,
449: we will directly use the interaction Hamiltonian and the corresponding
450: classical potential (see {\it e.g.} Eq.(\ref{Vclass})).
451:
452: The OR equations (Eqs.(\ref{OR1}) and (\ref{OR2})) have three solutions
453: and one of them is trivial: $t=0$, $b=0$. The two nontrivial mass
454: solutions
455: are shown in Fig.1 as solid and dashed lines.
456: The dashed line solution on Fig.1 becomes similar to the trivial one when
457: the dimensionless coupling becomes large: $M_{G\to\infty}\to m$,
458: $b_{G\to\infty}\to 0$. However, the solid line solution from Fig.1 is what
459: we are interested in because it fulfills the duality. For this case,
460: $M_{G\to\infty}^2\to {\frac {3}{2\pi}}m^2 G\ln G$,
461: $b_{G\to\infty}\to \sqrt{\frac{3}{4\pi}\ln G}$.
462: As one can see in this solution, when the dimensionless coupling $G$ of
463: the initial theory becomes large, the dimensionless couplings of the
464: quasiparticle theory, $\chi^{(3)}=gb/M^2$ and $\chi^{(4)}=g/M^2$,
465: get small\cite{efbook}: $\chi^{(3)}_{G\to\infty}\to
466: \sqrt{\frac{\pi}{3\ln G}}$,
467: $\chi^{(4)}_{G\to\infty}\to \frac{2\pi}{3 G\ln G}$. Thus, the
468: quasiparticle theory becomes perturbative when the initial coupling $G$
469: gets very strong.
470: The physical mass in this regime is close to
471: $M\approx m \sqrt{{\frac{3}{2\pi}}G \ln G}$, and the corrections are in
472: the order of ${\cal O}\bigl((\chi^{(3)})^2\bigr)$. In this way, we can
473: see the duality between the initial nonperturbative theory of unbroken
474: $\phi \rightarrow -\phi$ symmetry with a large dimensionless coupling $G$
475: and the quasiparticle theory of broken $\Phi \rightarrow -\Phi$ symmetry
476: with small couplings $\chi^{(3)}$ and $\chi^{(4)}$.
477: This duality is exact because the canonical transformation leaves
478: the Hamiltonian intact but changes only its
479: representation in terms of the quasiparticle field variable $\Phi_M(x)$.
480:
481: As we discussed earlier, the multiple solutions are not unexpected because
482: the OR equations, Eqs.(\ref{OR1}) and (\ref{OR2}), are algebraically
483: nonlinear. However, it can be shown (see Appendix B) that the two vacua
484: associated with different values of $t$ or $b$ are unitarily inequivalent
485: in the infinite volume limit. Thus, the two nontrivial solutions are not
486: connected to each other.
487: The bifurcation of quasi-particle masses appear above the coupling
488: $G \approx 9.04$.
489: However, the correct critical coupling of the phase transition should be
490: determined from the comparison of the energy density between the trivial
491: and condensed vacua. In Fig.2, we present the energy density of the
492: nontrivial solutions and find that the condensed vacuum energy density
493: of the duality-related solution gets lower than the trivial vacuum energy
494: density only above $G \approx 10.21$. This value $G \approx 10.21$
495: coincides with the critical coupling constant obtained by the GEP method
496: since our energy density given by Eq.(\ref{energy}) is equal to the GEP
497: at the minimum. This value is remarkably close to the
498: critical coupling constant (10.24(3)) obtained by the lattice
499: calculation\cite{lattice}.
500: Nevertheless, we note that our OR result doesn't take into account the
501: the higher order corrections in the quasiparticle effective coupling
502: $g/M^2$ and obtain the first order phase transition
503: instead of the second order phase transition which may be more
504: accurate description of the phase transition in $(\phi^4)_{1+1}$ theory
505: \cite{Chang,simon}. In this work, we thus focus only on the region where
506: the original coupling is much larger than the critical coupling so that the
507: perturbative calculation in terms of quasiparticle degrees of freedom is
508: allowed.
509:
510: Although there is a coincidence of the critical coupling,
511: it is not clear if the values of quasi-particle masses defined in the GEP,
512: HA and OR methods also coincide. In the GEP approach, it seems only
513: natural to define the quasiparticle mass as the second derivative
514: of the GEP\footnote{ See for example Eq.(3.2) of Ref.\cite{gep4}.}
515: , {\it i.e.} $d^2 {\bar V}_G(\phi_0)/d \phi_0^2$.
516: As discussed earlier, in the GEP approach a variational method is taken
517: to minimize the GEP satisfying just Eq.(\ref{OR1}).
518: Confirming the well-known equivalence between the GEP method and
519: the HA for the ground-state energy \cite{CJT,ft7},
520: the GEP result using this definition of
521: quasiparticle mass, {\it i.e.} the second derivative of the GEP, indeed
522: agrees precisely with the Hartree results (squares and circles in
523: Fig.\ref{geff})\footnote{In Chang's paper \cite{Chang} he found two
524: different types of nontrivial solutions using Hartree method ($4\pi c^2
525: >3$ and $4\pi c^2 <3$ in his notations). Nevertheless, the two type of
526: solutions lead in fact to
527: the same physical results as shown in Fig.\ref{geff}.
528: For more details, see the Appendix A.}.
529: However, one should note that the second
530: derivative of the GEP at the minimum $\phi_0 = b$ does not yield the mass
531: $M$ of the quasi-particle precisely but somewhat modified value
532: $M\sqrt{\frac{4\pi M^2 - 6 g}{4\pi M^2 + 3 g}}$; {\it i.e.}
533: with the notations used in Eq.(4.11) of Ref.\cite{gep4}, the second
534: derivative of the GEP, ${\bar V}_G$, is given by
535: \begin{equation}
536: \label{gepdd}
537: \frac{d^2{\bar V}_G}{d\phi_0^2}\bigg|_{\phi_0 = \pm
538: \sqrt{\frac{x}{8 {\hat \lambda_B}}}} = x \Bigl ( \frac{\pi x - 6 {\hat
539: \lambda_B}}{\pi x + 3 {\hat \lambda_B}} \Bigr ) \neq x,
540: \end{equation}
541: %while $\frac{d^2{\bar V}_G}{d \phi_0^2}|_{\phi_0 = 0} = x|_{\phi_0 =0} = 1$,
542: where $x = \frac{M^2}{m^2}$, ${\hat \lambda_B} = \frac{g}{4 m^2}$ and
543: $\phi_0 = b$ in our notations. Removing $t (= \ln x)$ in Eqs.(\ref{OR1})
544: and (\ref{OR2}), one can immediately get $b = \pm \frac{M}{\sqrt{2 g}}$
545: identical to $\phi_0 = \pm \sqrt{\frac{x}{8 {\hat \lambda_B}}}$.
546: This is the reason why the GEP result given by Eq.(\ref{gepdd})
547: shown by dotted line in Fig.\ref{geff}
548: doesn't coincide either of the bifurcated OR solutions (solid and dashed
549: lines). Of course, if both the ${\bar \Omega}$ equation and the GEP
550: minimum condition $d {\bar V}_G(\phi_0)/d \phi_0 = 0$ are solved
551: simultaneously
552: to determine ${\bar \Omega}$, then it is identical to the OR method and
553: the value of ${\bar \Omega}$ coincides with the quasiparticle mass $M$
554: defined in the OR method. However, we note once again that the GEP and HA
555: methods are quite different from the OR procedure because the GEP and HA
556: methods rely on the variational procedure while the OR method is
557: explicitly solving both Eqs.(\ref{OR1}) and (\ref{OR2}) simultaneously.
558:
559: As we have already discussed, the solid line in Fig.\ref{geff}
560: exhibits a duality between the theory with mass $m$ based on the trivial
561: vacuum $b=0$ and the theory with mass $M$ based on the nontrivial vacuum
562: $b \neq 0$. Namely, a strong coupling theory with mass $m$ is identical
563: to a weak coupling theory with mass $M$. A similar observation of duality
564: can be found in the solutions obtained by GEP and HA (see also Appendix
565: A), although the quasiparticle masses defined in the GEP and HA approaches
566: do not coincide with the quasiparticle mass defined in the OR method.
567: We also note that the GEP result (dotted line) approaches to
568: the OR result given by the solid line in the limit $G \to \infty$
569: since the terms of $x$ overwhelm the terms of $\hat \lambda_B$ in
570: Eq.(\ref{gepdd}) for that limit.
571:
572: In Ref. \cite{mp}, it was argued that the method of
573: Bogoliubov-Valatin transformations gives the same results as the GEP.
574: In fact all of the nonperturbative methods that we discuss in this
575: work,{\it i.e.} OR, GEP and HA, can be described formally in terms of
576: Bogoliubov-Valatin transformations because all of them use only the
577: change of particle mass and the field shift to get the new representation
578: of the field variables.
579: However, these transformations contain parameters (mass ratio
580: and the value of the shift) which can not be determined from
581: the transformations themselves. To fix the parameters, some
582: additional requirement is necessary.
583: The authors of Ref.\cite{mp} used essentially the same variational
584: requirement as the GEP
585: and therefore it is very natural that they found the same results as the
586: GEP. On the other hand, in the OR method, we are not using the GEP at all
587: but relying on the form of the Hamiltonian described in Section I.
588:
589: Applying the correspondence principle to the Hamiltonian (\ref{phi4ham}),
590: we now consider the classical potential given by
591: \begin{equation}
592: V^{class}(\Phi_M^{class}=\phi_M^{class}-b)=\frac{M^2}{2}(\phi^{class}_M-b)^2
593: +\frac{g}{4}(\phi^{class}_M-b)^4 + g b(\phi^{class}_M-b)^3,
594: \label{Vclass}
595: \end{equation}
596: which is written here in terms of the classical field
597: corresponding to the initial, unshifted field $\phi_M$.
598: As it can be easily seen, the potential has extremum at $\phi^{class}_M=b$.
599: This is provided by the demand not to have terms in Hamiltonian which
600: are linear in $\Phi_M$.
601: The double derivative in this point is larger than 0 and equal to
602: $M^2$,
603: {\it i.e.} the classical potential has minimum at $\phi^{class}_M=b$ and
604: the mass corresponding to oscillations around the minimum is precisely $M$.
605: That is why the OR method provides self-consistent procedure
606: of quantization of interacting fields and interpretation of the
607: solutions as particle excitations is valid. (In contrast,
608: $\frac{d^2{\bar V}_G}{d \phi_0^2}|_{\phi_0 = b} \neq \frac{M^2}{m^2}$ for
609: the GEP as shown in Eq.(\ref{gepdd}).) In addition, one can easily
610: prove the symmetry of $V^{class}$ under the
611: exchange of $\phi^{class}_M$ and
612: $-\phi^{class}_M$,{\it i.e.}$V^{class}(\phi_M^{class}) =
613: V^{class}(-\phi_M^{class})$,
614: utilizing the OR equations (Eqs.(\ref{OR1}) and (\ref{OR2})).
615:
616: In Fig.\ref{g4new}, the effective potential $V^{class}$ given by
617: Eq.(\ref{Vclass}) is presented.
618: As one can see in Fig.\ref{g4new},
619: the bifurcated solutions for $G>9.04$ indeed generate two
620: effective potentials (I,II) at a given coupling $G$, {\it e.g.} $G =$ 10
621: or
622: 12. Both potentials (I and II) in Fig.\ref{g4new} have clear minima and the
623: second derivative at each minimum generates the corresponding quasi-particle
624: mass $M$ precisely. The growth of bifurcation is also evident from the
625: comparison of the potentials I and II for $G =$ 10 and 12.
626: We note however that only the potential I provides the physically
627: meaningful duality-related solution and the potential II doesn't
628: satisfy the duality that we discuss in this work.
629:
630: An analysis of realistic models using the oscillator representation method
631: is more complicate (see, for example, discussion for 3+1 dimensional
632: $\phi^4$ theory in \cite{efbook}) due to substantial sophistication of
633: the renormalization procedure. Even the simple models however were not
634: fully studied in the framework of OR method.
635: As presented in this Section, the comparison of the OR predictions with
636: the GEP and Hartree results has been done for
637: 1+1 dimensional $\phi^4$ theory \cite{gep,Chang}. In higher
638: dimensions, the variational methods such as the GEP and HA may have a
639: rather fundamental difficulty.
640: The reason of this problem originates from the
641: error estimates of the approximation
642: in the GEP and Hartree methods.
643: For instance, the HA replaces the interaction term $\lambda \phi^n$ for
644: $n>2$ by the kinetic term $\Delta m^2 \phi^2+E_0$. The task of getting
645: the better approximation is to minimize the
646: positively defined measure $||\lambda \phi^n -
647: \Delta m^2\phi^2-E_0||$. Nevertheless, the vacuum
648: expectation value of the operator $\lambda \phi^n - \Delta m^2\phi^2-E_0 $
649: is not positively defined. Thus, it is necessary
650: to calculate the vacuum expectation value for its square that
651: is always positive. In that case, even if the interaction term itself
652: is renormalizable, its square may be not (in particular, for $D>2$).
653: That is why in higher dimensions the GEP and Hartree methods come across
654: a renormalization problem in computing their error measures\cite{Munoz}.
655: More discussion about the difficulty in the GEP method for the higher
656: dimension can be found in Ref.\cite{wudka}.
657: However, this is not the case for the OR method, since it does not use any
658: variational procedure.
659:
660: Therefore, among the three (OR,GEP,Hartree), the OR method seems to be
661: most attractive. The reasons may be summarized as follows:\\
662: i) The problems of renormalization in higher dimensions are technical,
663: but not principal.\\
664: ii) The Hamiltonian is identically rewritten via canonical transformation
665: and consequently the transformed Hamiltonian keeps the identical
666: information as the original Hamiltonian.\\
667: iii) The second derivative of effective potential gives the
668: quasi-particle mass precisely in the OR, while the
669: others give somewhat modified values.
670:
671: \section{Application of the OR method to the $(\phi^6)_{1+1}$ theory}
672: Using the OR method, we now compute the parameters of different phases
673: for the scalar field theory with both $\phi^4$ and $\phi^6$ interactions.
674: The starting classical Hamiltonian density is
675: \begin{equation}
676: {\cal H}_m=\frac{1}{2}\pi^2_m+\frac{1}{2}(\nabla\phi_m)^2
677: +\frac{m^2}{2}\phi^2_m
678: +\frac{g}{4}\phi^4_m + h\phi^6_m.
679: \label{clasphi6}
680: \end{equation}
681: The way to proceed is the same as in the case of $\phi^4$ theory.
682: Using the formulas (\ref{normfree}) and (\ref{normexp}) and shifting the
683: quantization point, we get:
684: \begin{equation}
685: N_m({\cal H}_m)=N_M \bigl ({\cal H}_M^{right} + {\cal H}_1 \bigr ) +
686: \varepsilon_M,
687: \label{hamilt6}
688: \end{equation}
689: where the ``right'' Hamiltonian density ${\cal H}_M^{right}$ is given by
690: \begin{eqnarray}
691: \label{rightH}
692: {\cal H}_M^{right} &=& \frac{1}{2}\pi^2_M+\frac{1}{2}(\nabla\Phi_M)^2+
693: \frac{M^2}{2}\Phi_M^2 + \frac{g}{4}\Phi_M^4 +h\Phi_M^6 + gb\Phi_M^3
694: \\ \nonumber &+& 6hb\Phi_M^5+15hb^2\Phi_M^4
695: + 20hb^3\Phi_M^3-\frac{15ht}{4\pi}\Phi_M^4-\frac{15hbt}{\pi}\Phi_M^3,
696: \end{eqnarray}
697: the remainder ${\cal H}_1$ is given by
698: \begin{eqnarray}
699: {\cal H}_1 &=& \Bigl ( \frac{1}{2}(m^2-M^2)+\frac{3gb^2}{2}-\frac{3gt}{8\pi}
700: +15b^4h-\frac{45b^2ht}{2\pi}+\frac{45ht^2}{16\pi^2}\Bigr )\Phi_M^2
701: \nonumber \\
702: &+&\Bigl ( m^2 b+gb^3-\frac{3gbt}{4\pi}+6hb^5-\frac{15hb^3t}{\pi}+
703: \frac{45bht^2}{8\pi^2} \Bigr )\Phi_M,
704: \end{eqnarray}
705: and the energy density $\varepsilon_M$ of the vacuum is given by
706: \begin{eqnarray}
707: \varepsilon_M
708: &=&\frac{m^2b^2}{2}+\frac{gb^4}{4}-\frac{m^2t}{8\pi}-\frac{3gb^2t}
709: {8\pi}+\frac{3gt^2}{64\pi^2} \nonumber \\
710: &+&\frac{1}{8\pi}(M^2-m^2)+hb^6-\frac{15hb^4t}{4\pi}
711: +\frac{45hb^2t^2}{16\pi^2}-\frac{15ht^3}{64\pi^3}.
712: \end{eqnarray}
713: As in the last section, here we use the notation $\Phi_M = \phi_M - b$.
714:
715: The requirement to have the Hamiltonian in the right form means that
716: the remainder ${\cal H}_1$ is equal to zero. This gives the following two
717: equations for $t$ and $B$ ($B=b^2$):
718: \begin{equation}
719: 1-e^t+3GB-\frac{3Gt}{4\pi}+30HB^2-\frac{45HBt}{\pi}+\frac{45Ht^2}{8\pi^2}=0,
720: \label{equ1}
721: \end{equation}
722: \begin{equation}
723: 1+GB-\frac{3Gt}{4\pi}+6HB^2-\frac{15HBt}{\pi}+\frac{45Ht^2}{8\pi^2}=0,
724: \label{equ2}
725: \end{equation}
726: where
727: $G=g/m^2$ and $H=h/m^2$ are the dimensionless couplings.
728: The conditions (\ref{equ1}) and (\ref{equ2}) again provide the minimum of
729: the classical potential at $\phi^{class}_M=b$ and the right value of the
730: double derivative around it.
731:
732: Solving Eqs. (\ref{equ1}) and (\ref{equ2}) numerically, we found
733: five nontrivial mass solutions shown in
734: Figs. \ref{phasesnew1}, \ref{phasesnew2} and \ref{phasesnew3}. The
735: corresponding vacuum energy densities are presented in Figs. \ref{vac6sym},
736: \ref{vac6brok} and \ref{vac6spur}.
737: In Fig.\ref{vac6sym}, two identical plots are shown in two different
738: angles of view to make clear presentation of the result.
739: Among the five, two solutions correspond to the symmetric phase,
740: {\it i.e.} $b=0$. They are represented by the upper and lower branches
741: (or sheets) of the manifold shown in Fig.\ref{phasesnew1}. The vacuum
742: energy densities for these solutions are shown in Fig.\ref{vac6sym}.
743: The upper and lower branches in Fig.\ref{phasesnew1} correspond to the
744: lower and upper branches in Fig.\ref{vac6sym}, respectively.
745: The upper branch in Fig.\ref{phasesnew1} is the physically meaningful
746: duality-related solution while the lower branch in Fig.\ref{phasesnew1}
747: does not provide the duality that we are interested in this work. The
748: duality-related solution
749: has the lower energy density as shown in Fig.\ref{vac6sym}.
750: The dimensionless couplings which correspond
751: to the $\Phi^6$ and $\Phi^4$ interactions of the duality-related
752: quasiparticle theory are given by
753: \begin{equation}
754: \chi_s^{(6)} = \frac{h}{M_s^2},~~~\chi_s^{(4)}=\frac{g}{4M_s^2}-
755: \frac{15ht}{4\pi M_s^2}.
756: \label{pertunbr}
757: \end{equation}
758: Note here that $\chi_s^{(4)}$ includes both couplings of $g$ and $h$. This is
759: due to the self-organization of fields via the quantum fluctuations in the
760: vacuum, {\it i.e.} the original $\phi^6$ interaction induces the effective
761: $\Phi^4$ interaction with the negative effective coupling in this case.
762: Their dependencies on $H$ are presented in Fig.\ref{coup1},
763: where the original $\phi^4$ interaction coupling $g$ is put to zero for
764: simplicity. It can be seen from Fig.\ref{coup1}, both $\chi_s^{(4)}$
765: and $\chi_s^{(6)}$ become small as $H$ grows.
766: Thus, the qusiparticle theory in the domain of large $H$ can be solved by
767: a standard perturbation method.
768:
769: Among the rest three solutions, the two solutions corresponding to the
770: broken-symmetry phase with the nonzero real condensation ($b=\sqrt{B}$)
771: are shown as the two branches in Fig.\ref{phasesnew2}. The vacuum energy
772: densities for these solutions are shown in Fig.\ref{vac6brok}.
773: Again the upper (lower) branch of Fig.\ref{phasesnew2} corresponds to
774: the lower (upper) branch of Fig.\ref{vac6brok}.
775: Similar to the case of
776: symmetric phase (Fig.\ref{phasesnew1}), the lower branch in
777: Fig.\ref{phasesnew2} does not provide the duality that we are interested
778: in this work. However, the upper branch in Fig.\ref{phasesnew2} corresponds
779: to the duality-related solution and gives the lower value of the vacuum
780: energy density (the lower branch in Fig.\ref{vac6brok}).
781: For this broken-symmetry phase, the dimensionless couplings of the
782: quasiparticle interactions are given by
783: \begin{eqnarray}
784: \chi_{bs}^{(6)}=\frac{h}{M_{bs}^2},~~~
785: \chi_{bs}^{(5)}=\frac{6hb}{M_{bs}^2},~~~
786: \chi_{bs}^{(4)}=\frac{g}{4M_{bs}^2}+
787: \frac{15hb^2}{M_{bs}^2}-\frac{15ht}{4\pi M_{bs}^2},
788: \nonumber \\
789: \chi_{bs}^{(3)}=\frac{gb}{M_{bs}^2}+\frac{20hb^3}{M_{bs}^2}-
790: \frac{15hb t}{\pi M_{bs}^2}.
791: \label{pertbr}
792: \end{eqnarray}
793: The $H$-dependence of these couplings is shown in Fig.\ref{coup2}
794: (again, we put $g=0$ to simplify the presentation).
795: At sufficiently large $H$ the couplings get small and the quasiparticle
796: theory becomes perturbative. When the dimensionless
797: couplings of the symmetric solution go to zero,
798: $\chi^{(6)}_s,\chi^{(4)}_s \to 0$,
799: the dimensionless couplings of the broken-symmetry theory also go to zero
800: as shown in Fig.\ref{couplings}.
801:
802: The last solution shown in Fig.\ref{phasesnew3} has a pure imaginary value
803: for the condensate $b$ ({\it i.e.} negative $B$). In Fig.\ref{vac6spur},
804: we show the corresponding vacuum energy density. Because the Hamiltonian
805: in this case becomes non-hermitian, we call these solutions
806: spurious. The spurious solutions appear only after the inclusion of
807: $\phi^6$ interaction. Thus, they are consequences of the higher nonlinearity
808: in the OR equations. Since they don't seem to bear any interesting duality
809: property that we discuss in this work, we don't present their results in
810: the main text but just summarize those in the Appendix C with some
811: discussion.
812:
813: The most interesting feature of the $\phi^6$ theory is the appearence of
814: the two apparently different quasiparticle perturbation theories
815: (see the upper branches of Figs. \ref{phasesnew1} and \ref{phasesnew2})
816: which represent the initial
817: nonperturbative theory. The first one corresponds to the symmetric
818: phase with the dimensionless quasiparticle couplings given by
819: Eq.(\ref{pertunbr}) and the second one corresponds to the broken-symmetry
820: phase with the dimensionless couplings given by Eq.(\ref{pertbr}).
821: The subscripts ``s'' and ``bs'' of the coupling $\chi$ and the mass $M$
822: represent the correspondence to the symmetric and broken-symmetry phases,
823: respectively. The two quasiparticle theories ({\it i.e.} $\chi_s$'s and
824: $\chi_{bs}$'s) differ from each other significantly in the sence that
825: one of them ($\chi_s$'s) preserves the
826: $\phi\to -\phi$ symmetry of the initial Hamiltonian (\ref{clasphi6})
827: upon changing $\Phi\to -\Phi$, while
828: the other ($\chi_{bs}$'s) does not respect the $\Phi\to -\Phi$ symmetry.
829: Physically, this means that in the theory of symmetric phase with
830: the $\chi_s$ couplings the processes
831: involving odd number of particles are not allowed while they are
832: allowed in the broken-symmetry theory with $\chi_{bs}$ couplings. This is
833: because in the theory of symmetric phase,
834: the Hamiltonian depends on the even powers of $\Phi$
835: (to provide $\Phi \to -\Phi$ symmetry) and thus the S-matrix contains only
836: the even number of creation and annihilation operators.
837:
838: However, one may wonder if these two apparently different quasiparticle
839: theories are in fact equivalent to each other describing the same physics
840: with just different degrees of freedom. In this respect,
841: we notice that the two different quasiparticle theories give
842: the same effective potential in the limit of very large $G$ and
843: $H$. In Figs.\ref{effpot1},\ref{effpot2} and \ref{effpot3}
844: we show the classical potentials which
845: correspond to the two quasiparticle theories at different values of
846: the coupling $\chi_s^{(6)}$.
847: As it can be seen, when the quasiparticle coupling $\chi_s^{(6)}$
848: becomes small ({\it i.e.} $H$ becomes very large), the relative difference
849: between the two quasiparticle
850: potentials diminishes substantially. We verified that the two classical
851: potentials are indeed identical to each other in the $\chi_s^{(6)}\to 0$
852: limit.
853:
854: In the GEP approach \cite{gep6}, it was also recognized that there are three
855: solutions of symmetric phase (one of them is the trivial solution and two
856: others are the nontrivial solutions corresponding to the ones that we
857: showed in Fig.\ref{phasesnew1}). As we consider only one
858: of the two nontrivial solutions (the upper branch in Fig.\ref{phasesnew1})
859: as the physically relevant (duality-related) solution, the GEP approach
860: also dealt with only one (physically relevant)
861: solution among the two nontrivial solutions.
862: In Ref.\cite{gep6} it was shown that the form of $\bar \Omega$ equation
863: of the GEP approach is invariant under the exchange of $m$ and $M_s$
864: with an appropriate redefinition of effective couplings
865: $G'=(G-\frac{15H}{\pi}t_0)e^{-t_0}$ and $H'=He^{-t_0}$, as well as
866: $t'=t-t_0$, where $t_0=\ln\frac{M_s^2}{m^2}$ satisfies the
867: constraint equation given by Eq.(\ref{equ2}): $1-e^{-t_0}=
868: \frac{3Gt_0}{4\pi}-\frac{45Ht_0^2}{8\pi^2}$ (Note $B=0$ for the symmetric
869: phase). Indeed, we can show the form
870: invariance of both OR Eqs. (\ref{equ1}) and (\ref{equ2}) under
871: the same transformations.
872:
873: Observing such form invariance, it was argued in the GEP approach \cite{gep6}
874: that the nontrivial symmetric phase solution is just a duplication of the
875: trivial solution and thus should be avoided. However, as we discussed above,
876: the duality between the trivial and nontrivial symmetric phase
877: solutions provides the physical meaning to the form invariance of
878: the OR equations. Because of the duality, we see the utility of
879: both trivial and nontrivial solutions in the symmetric phase.
880: While the trivial
881: solution can be utilized for the perturbation theory at small $G$ and $H$,
882: the nontrivial symmetric phase solution can be used for the construction
883: of quasiparticle perturbation theory in the regime of very large $G$ and $H$.
884:
885: Furthemore, since there exists also a duality between the trivial
886: solution and the nontrivial broken-symmetry solution we can rather easily
887: understand the equivalence of the classical effective potentials
888: between the symmetric phase and the
889: broken-symmetry phase for the very large couplings of the original theory
890: (see Fig.\ref{effpot3}).
891: In the limit of $\chi_s^{(6)}\to 0$ (which means also $\chi_{bs}$'s$\to 0$,
892: as shown in Fig.\ref{couplings}),
893: the higher order corrections get negligible and thus the equivalence of
894: the classical potentials between the symmetric and broken-symmetry phases
895: is revealed manifestly as shown in Fig.\ref{effpot3}.
896:
897:
898: \section{Summary and Conclusions}
899: The formation of quasiparticle in RQFT may be a good example of
900: self-organizing nature in the nonperturbative strong interactions. This
901: nature is realized
902: by altering the vacuum structure and lowering the vacuum energy density
903: as the strength of the interactions gets more enhanced. Our main goal in this
904: paper was to provide the explicit examples that reveal such nature of
905: self-organizing RQFT and show the utility of duality generated by the
906: canonical transformation in solving the nonperturbative strong interaction
907: problem. For the explicit but simple examples, we analized
908: of 1+1 dimensional $\phi^4$ and $\phi^6$ theories. Our purpose was
909: fulfilled by applying the OR method which is a generalization of the
910: canonical quantization scheme.
911:
912: The OR method uses canonical commutation relations for
913: the free field operators and employs
914: the canonical transformation of quantum fields represented by the change
915: of the field mass and the condensation of the field.
916: The idea to demand the right form of the Hamiltonian,
917: {\it i.e.} excluding the terms linear in $\Phi$ and
918: taking the coefficient of $\Phi^2$ as $M^2/2$
919: guarantees that the classical potential has the minimum at the
920: quantization point $\phi^{class}_M=b$ and that the double derivative of the
921: classical potential at this point is given by $M^2$. This assures that the OR
922: method gives a self-consistent procedure of quantization and a
923: valid description of the field excitations around the minimum point in
924: terms of quasiparticles.
925:
926: In contrast to the GEP and HA methods, the OR method is not based
927: on the variational procedure so that it does not have a bottleneck
928: of renormalization problem in higher dimensions as discussed in Section II.
929: However, the OR method has not been utilized as extensively as the GEP and HA
930: methods. Even in the 1+1 dimensional scalar field theory, not much of the
931: OR results were contrasted from the GEP and HA results although the OR
932: results for the $(\phi^4)_{1+1}$ theory were presented in
933: Ref.\cite{efbook}. We attempted to fill such gap in
934: this work.
935:
936: For the fixed classical
937: Hamiltonian and renormalization procedure (fixed by the counter-terms),
938: it is natural to expect that the OR method yields several different
939: solutions for the quasiparticle mass and condensation because the OR
940: equations are in general nonlinear. However, the unitary inequivalence
941: arising from these different OR vacuum solutions prohibits the
942: transitions among the quasiparticles with different mass and/or condensation.
943: Although some of the solutions correspond
944: to the heavier quasiparticles, they do not decay into the light ones
945: due to the unitary inequivalence;
946: {\it i.e.} each sort of the quasiparticles is stable within the OR.
947: The unitary inequivalence between the OR solutions is shown in Appendix B.
948: A recent application of the unitary inequivalence to the flavor
949: mixing phenomena
950: can be found in Ref.\cite{ji-mish}. In this work, we selected only
951: the duality-related OR solutions because they can provide the immediate
952: physical interpretation. The reason why the physical interpretation
953: is immediate for the duality-related solutions is because the strong
954: coupling particle theory becomes solvable
955: once it is transformed into an exactly equivalent quasiparticle theory
956: with weak couplings.
957:
958: Using the $(\phi^4)_{1+1}$ theory, we compared the OR results
959: with the results obtained by GEP and HA. In Ref.\cite{efbook}, it was
960: discussed that the value of the critical coupling $G\approx 10.21$
961: obtained by the OR method
962: coincides with the one obtained by the GEP method.
963: We also note that this value of the critical coupling agrees
964: remarkably well with the lattice result \cite{lattice}.
965: Due to the equivalence
966: between the HA and GEP methods for the ground state energy \cite{CJT,ft7}
967: the agreement of the critical couplings between the OR and HA methods follows
968: as well. Although the critical values of coupling constant
969: $G$ turned out to be same for all three methods (OR, GEP, HA),
970: the OR method gives
971: the quantitatively different dependence of the nontrivial field mass on the
972: effective coupling constant, compare to the GEP and HA results
973: (see Fig.\ref{geff} and the discussions in Section II).
974:
975: We have extended the application to the 1+1 dimensional
976: $\phi^6$ theory and
977: compared the results of OR method with what was already obtained
978: by the GEP. We found that the different solutions for the mass of
979: quasi-particles between the OR and the GEP are sustained.
980: In the OR analysis of $(\phi^6)_{1+1}$ theory we found two physically
981: meaningful duality-related solutions (upper sheets of Figs.\ref{phasesnew1}
982: and \ref{phasesnew2}), one in the symmetric phase and the other in the
983: broken-symmetry phase. Although these two solutions have apparently
984: different symmetry property, we find that their effective potentials do
985: agree in the limit of very large couplings $G$ and $H$ of the initial theory,
986: {\it i.e.} very small effective couplings $\chi_s$'s and $\chi_{bs}$'s of
987: the quasiparticle theories. This indicates an existence of a unique
988: effective
989: potential to the quasiparticles regardless of their symmetry properties.
990: The actual perturbation theory using the quasiparticles degrees of
991: freedom with the systematic higher order corrections may deserve further
992: investigation.
993:
994: \begin{center}
995: {\large\bf Acknowledgments}
996: \end{center}
997: CRJ thanks Yuriy Mishchenko for many useful discussions.
998: This work is supported by the Brain Korea 21 project in year 2001, the Korea
999: KRF 2001-015-DP0085, the Korea KOSEF 1999-2-111-005-5, the
1000: RFBR-01-02-16431, the RFBR-03-02-17291, the INTAS-2000-366 and the US DOE
1001: under grant No. DE-FG02-96ER40947. The North Carolina Supercomputing
1002: Center and the National Energy Research
1003: Scientific Computer Center are also acknowledged for the grant of Cray time.
1004:
1005: \setcounter{equation}{0}
1006: %\setcounter{figure}{0}
1007: \renewcommand{\theequation}{\mbox{A\arabic{equation}}}
1008: \begin{center}
1009: {\bf APPENDIX A: Hartree Approximations in the $(\phi^4)_{1+1}$ Theory}
1010: \end{center}
1011: The essential idea of the Hartree approximation\cite{Chang} is to linearize
1012: the non-linear field equations using the mean fields,
1013: {\it e.g. $\phi^3 \rightarrow 3 \langle \phi^2 \rangle \phi -2
1014: \langle \phi \rangle ^3$, etc.},
1015: splitting the fields into classical mean fields
1016: ($\phi_c = \langle \phi \rangle$)
1017: plus the quantum fields($\phi_q$), {\it i.e. $\phi = \phi_c + \phi_q$}.
1018: In the classical limit, we know that the ground state is given by
1019: $\phi^2=c^2$ if $c^2$ is defined by $c^2 = -\frac{m^2}{g}$ and positive
1020: ({\it i.e. $c^2>0$}).
1021: The quantum fluctuations $\langle \phi_q^2(x) \rangle$ are found to be
1022: infinite.
1023: However, one can renormalize
1024: the theory using a mass counter term $\frac{1}{2} B \phi^2$, where
1025: the constant $B$ is determined by requiring that $\phi_c = c$ is a static
1026: solution (corresponding to the so-called ``abnormal vacuum state" or a
1027: sort of vacuum condensate)
1028: of a Hartree equation and this leads to
1029: $B= -3 g \langle \phi_q^2 \rangle_{\phi_c =c}$.
1030: After renormalization, the effective potential $V(\phi_c)$ is obtained in
1031: terms of the subtracted quantum fluctuation, {\it i.e.
1032: $\Delta \langle \phi_q^2 \rangle = \langle \phi_q^2 \rangle_{\phi_c}
1033: - \langle \phi_q^2 \rangle_{\phi_c = c}$},
1034: which satisfies the iterative (mass-gap type) equation\cite{Chang};
1035: \begin{equation}
1036: \Delta\langle\phi_q^2\rangle = \frac{1}{4\pi} \ln\frac{2c^2}
1037: {3\phi_c^2 -c^2 + 3\Delta\langle\phi_q^2\rangle}.
1038: \label{gapeq}
1039: \end{equation}
1040: Using the solutions of Eq.(\ref{gapeq}), one can find that $\phi_c = c$
1041: is a local minimum (maximum) of $V(\phi_c)$ for $4\pi c^2 >3 (0<4\pi c^2<3)$
1042: while $\phi_c=0$ is always a local minimum. For $0<4\pi c^2<3$, a nontrivial
1043: local minimum exists at $\phi_c = \sqrt{c^2 -
1044: 3\Delta\langle\phi_q^2\rangle} = c^\prime$.
1045: Expanding $V(\phi_c)$ in each of the local minima, one can introduce the
1046: following mass parameters;
1047: \begin{eqnarray}
1048: m_0^2 = \frac{d^2 V(\phi_c)}{d\phi_c^2}\bigg|_{\phi_c=0}
1049: = g(3\Delta\langle\phi_q^2\rangle|_{\phi_c=0} - c^2), \nonumber \\
1050: m_c^2 = \frac{d^2 V(\phi_c)}{d\phi_c^2}\bigg|_{\phi_c=c}
1051: = 2gc^2 \frac{8\pi c^2 -6}{8\pi c^2 +3}, \nonumber \\
1052: m_{c^\prime}^2 = \frac{d^2 V(\phi_c)}{d\phi_c^2}\bigg|_{\phi_c=c^\prime}
1053: = 4g(c^2-3\Delta\langle\phi_q^2\rangle)
1054: \frac{4\pi(c^2-3\Delta\langle\phi_q^2\rangle)-3}
1055: {8\pi(c^2-3\Delta\langle\phi_q^2\rangle)+3}.
1056: \label{masses}
1057: \end{eqnarray}
1058: Then, the intrinsic strength measured in terms of these mass parameters
1059: are given by the dimensionless couplings; {\it i.e.}
1060: \begin{equation}
1061: g_0 = \frac{g}{m_0^2}, \mbox{\hskip 5mm} g_c = \frac{g}{m_c^2},
1062: \mbox{\hskip 5mm} g_{c^\prime} = \frac{g}{m_{c^\prime}^2}.
1063: \label{dimlesscoupling}
1064: \end{equation}
1065: For the entire $c^2>0$ region, it turns out that the nontrivial vacuum
1066: solutions exist only for $g_0 \geq g_{crit} \approx 9.045$ (note that it
1067: is not the same as the critical coupling of the phase transition $G=10.21$
1068: as discussed in Section II),
1069: where the critical coupling $g_0 = g_{crit}$ occurs at $4\pi c^2 =3$.
1070: Thus, it appears that one can find the two nontrivial vacuum solutions;
1071: one for $4 \pi c^2 > 3$ and the other for $0< 4 \pi c^2 < 3$. However,
1072: we find that the two solutions are indeed identical in the sense that
1073: $g_c$ and $g_{c^\prime}$ coincide when we plot them (see
1074: Fig.~\protect\ref{app}) as a function of $g_0$,
1075: {\it i.e.} $g_c(g_0) = g_{c^\prime}(g_0)$ for all $g_0 \geq g_{crit}$.
1076:
1077: We also notice a duality between the theory with $g_0$ based on the
1078: trivial vacuum $\phi_c =0$ and the theory with $g_c$ (or $g_{c^\prime}$)
1079: based on the nontrivial vacuum $\phi_c = c$ for $4 \pi c^2 > 3$
1080: (or $\phi_c = c^\prime$ for $0 < 4 \pi c^2 < 3$).
1081: Namely, a strong coupling $g_0$ theory is identical to a weak coupling
1082: $g_c$ (or $g_{c^\prime}$) theory.
1083:
1084: \setcounter{equation}{0}
1085: %\setcounter{figure}{0}
1086: \renewcommand{\theequation}{\mbox{B\arabic{equation}}}
1087: \begin{center}
1088: {\bf APPENDIX B: Unitary Inequivalence between the Nontrivial OR
1089: Solutions}
1090: \end{center}
1091:
1092: In this Appendix, we show the unitary inequivalence between a vacuum
1093: with $b_1$ and $t_1$ and another vacuum $b_2$ and $t_2$, where $t_1\neq
1094: t_2$ or $b_1 \neq b_2$.
1095:
1096: The canonical transformation taking a trivial vacuum ({\it i.e.}
1097: $b=0$ and $t=1$) into a nontrivial one ({\it i.e.} $b \neq 0$ and
1098: $t \neq 1$) has the following form ({\it e.g.} for the bosons)
1099: \begin{equation}
1100: \label{eq01z}
1101: \begin{array}{cc}
1102: 1. & a_k \rightarrow a_k-2\pi mb\delta (k); \\
1103: 2. &
1104: \begin{array}{c}
1105: a_k \rightarrow a_k\cosh (\lambda )-a_{-k}^{\dagger}\sinh (\lambda ), \\
1106: a_k^{\dagger} \rightarrow a_k^{\dagger}\cosh (\lambda )-a_{-k}\sinh
1107: (\lambda ).
1108: \end{array}
1109: \end{array}
1110: \end{equation}
1111: Their corresponding operators have the form
1112: \begin{equation}
1113: \label{eq02z}
1114: \begin{array}{c}
1115: U_1=\exp \left\{ -2\pi mb(a_0-a_0^{\dagger })\right\}; \\
1116: U_2=\exp \left\{ \frac 12\int d\vec k\lambda (k)(a_{-k}a_k-a_k^{\dagger
1117: }a_{-k}^{\dagger })\right\}
1118: \end{array}
1119: \end{equation}
1120: with $\lambda (k)=\frac 12\ln \frac{\omega (k)}{\omega (k,t)}$ and $\omega
1121: (k)=\sqrt{k^2+t^2m^2}$.
1122:
1123: To connect the two vacua, one with $b_1$ and $t_1$ and another with
1124: $b_2$ and $t_2$, we can use the transformation given by Eq.(\ref{eq02z})
1125: with $b=b_2-b_1$ and $\lambda (k)=\frac 12\ln
1126: \frac{\omega _1(k)}{\omega _2(k)}$ where $\omega _i(k)=\sqrt{k^2+t_i^2m^2}$.
1127: To show that, let us define $V_i:\left\langle 0,1\right\rangle
1128: \rightarrow
1129: \left\langle b_i,t_i\right\rangle $ by the transformation given by
1130: Eq.(\ref{eq02z}). Then
1131: $$
1132: \left\langle b_1,t_1\right\rangle \rightarrow \left\langle
1133: b_2,t_2\right\rangle =\left\langle b_1,t_1\right\rangle \rightarrow
1134: \left\langle 0,1\right\rangle \rightarrow \left\langle b_2,t_2\right\rangle
1135: $$
1136: and therefore $V=V_2V_1^{-1}:\left\langle b_1,t_1\right\rangle \rightarrow
1137: \left\langle b_2,t_2\right\rangle $. Straightforward algebraic
1138: manipulation with Eq.($\ref{eq02z}$) then results in the unitary
1139: inequivalence between
1140: the vacuum with $\left\langle b_1,t_1\right\rangle$ and the vacuum
1141: with $\left\langle b_2,t_2\right\rangle$.
1142:
1143: In Ref.\cite{efbook}, it was mentioned that%
1144: $$
1145: \begin{array}{c}
1146: U_1=e^{-\frac 12(2\pi mb)^2[a_0,a_0^{\dagger }]}:U_1:, \\
1147: U_2=e^{-\frac V{(2\pi )^3}\int d\vec k\ln (\cosh [\lambda (k)])}:U_2:.
1148: \end{array}
1149: $$
1150: where $[a_0,a_0^{\dagger }]=\delta ^{(3)}(0)$. Here, we can equate
1151: $\delta^{(3)}(0)$ to $\frac V{(2\pi )^3}$. Note also
1152: that transformation $U_1$ acts only on the $k=0$ field operators, therefore%
1153: $$
1154: \begin{array}{c}
1155: U_2\circ U_1=\left( \prod\limits_{k\neq 0}U_2(k)\right) \left( U_2(0)\circ
1156: U_1(0)\right) , \\
1157: :U_2\circ U_1:=\left( \prod\limits_{k\neq 0}:U_2(k):\right) :U_2(0)\circ
1158: U_1(0):.
1159: \end{array}
1160: $$
1161: Hence%
1162: $$
1163: U_2\circ U_1=e^{-\frac V{(2\pi )^3}\int\limits_{k\neq 0}d\vec k\ln (\cosh
1164: [\lambda (k)])}\prod\limits_{k\neq 0}:U_2(k):\left( U_2(0)\circ
1165: U_1(0)\right) .
1166: $$
1167: Then for the inner product of the vacuum $\left| 0\right\rangle _1$
1168: associated with $\left\langle b_1,t_1\right\rangle $ and $\left|
1169: 0\right\rangle _2$ associated with $\left\langle b_2,t_2\right\rangle $ we
1170: have%
1171: $$
1172: _1\left\langle 0|0\right\rangle _2=\prod {}_{1,k}\left\langle
1173: 0|0\right\rangle _{2,k}\leq \prod_{k\neq 0}{}_{1,k}\left\langle
1174: 0|0\right\rangle _{2,k}=e^{-\frac V{(2\pi )^3}\int\limits_{k\neq 0}d\vec
1175: k\ln (\cosh [\lambda (k)])}\rightarrow 0,
1176: $$
1177: as $$\text{ }V\rightarrow \infty \text{.}$$
1178: The two vacua therefore are unitary inequivalent if their $t$ parameters
1179: are
1180: different. If $t_1=t_2$ but $b_1\neq b_2$ then $\lambda (k)\equiv 0$ and $%
1181: V=U_1(b_2-b_1).$ Then%
1182: $$
1183: _1\left\langle 0|0\right\rangle _2=\prod {}_{1,k}\left\langle
1184: 0|0\right\rangle _{2,k}={}_{1,k=0}\left\langle 0|0\right\rangle
1185: _{2,k=0}=e^{-\frac 12(2\pi mb)^2\delta (0)}=0\text{,}
1186: $$
1187: in the infinite volume limit.
1188:
1189: Therefore $_1\left\langle 0|0\right\rangle _2=0$ in the infinite volume
1190: limit whenever $t_1\neq t_2$ or $b_1\neq b_2$.
1191:
1192:
1193: \setcounter{equation}{0}
1194: %\setcounter{figure}{0}
1195: \renewcommand{\theequation}{\mbox{B\arabic{equation}}}
1196: \begin{center}
1197: {\bf APPENDIX C: Spurious Solutions in the $(\phi^6)_{1+1}$ Theory}
1198: \end{center}
1199: Among the solutions of OR Eqs.(\ref{equ1}) and (\ref{equ2}) for 1+1
1200: dimensional $\phi^6$ theory there is one with negative $B$ that gives
1201: pure imaginary condensate $b=i\sqrt{|B|}$. This solution is shown
1202: in Fig.\ref{phasesnew3} and the corresponding vacuum energy density is shown
1203: in Fig.\ref{vac6spur}.
1204: In this case the Hamiltonian becomes non-hermitian because
1205: it contains anti-hermitian terms in odd powers of $\Phi_M$.
1206: When one calculates average values
1207: of such operators in any state with definite number
1208: of particles, these terms drop out, because odd number of creation
1209: and annihilation operators can not leave the number of particles
1210: unchanged. However, the physical states are complicate and
1211: contain indefinite number of particles
1212: so that the antihermitian terms would contribute to the physical
1213: observables and give them imaginary expectation values.
1214: Moreover, our starting Hamiltonian density
1215: given by Eq.(\ref{clasphi6}) involves only the real (not complex as in
1216: the O(2) symmetric theory) scalar field $\phi_m$ and even after shifting
1217: the quantization point the field $\Phi_M$ is required to remain as real
1218: not complex. For these reasons, we call the imaginary $b$ solution
1219: spurious. In principle, we can throw away the spurious solution and
1220: do not consider the imaginary-$b$ phase in the simple one-component model.
1221:
1222: Nevertheless, in the history of many-body analysis sometimes the occurance
1223: of such spurious solution (forbidden in principle) led to interesting
1224: physics. For example, the occurance of imaginary solution has been
1225: previously observed in the analysis of many-body system when the method
1226: gets more accurate but the number of degrees of freedom is not sufficient.
1227: Such example may be illustrated in the random phase approximation (RPA)
1228: calculation of a fermion system\cite{Kadanoff} where the imaginary energy
1229: poles are found if the fermion interaction is attractive and weak below a
1230: certain critical temperature. These imaginary poles were
1231: interpreted\cite{Kadanoff} as the mathematical manifestation
1232: of the instability which leads to the superconducting state. Such
1233: imaginary poles were however not found in the less accurate method such as
1234: the Tamm-Dancoff approximation (TDA).
1235: Thus, if there is any possibility that the imaginary solution could
1236: actually mean something interesting in the many-body analysis,
1237: we should not overlook such a possibility.
1238:
1239: Due to such a tantalizing possibility that the imaginary solution could
1240: occur indicating an existence of a new phase, we note a few more
1241: particular features of the spurious solutions
1242: instead of simply throwing those solutions away.
1243: In particular, it may be interesting to note that the corresponding
1244: quasi-particle mass values become quite small ($M<<m$)
1245: when the imaginary $b$ solutions have the least vacuum energy density
1246: as shown in Figs.\ref{phasesnew3} and \ref{vac6spur}.
1247: We do not know yet, however, whether these spurious solutions are anything
1248: to do with the condensation of (zero-mass) Goldstone-bosons if the degrees
1249: of freedom are added in such a way that the fields become complex and a
1250: continuous symmetry (see however the discussion for O(2) symmetry near the
1251: end of this Appendix) is restored.
1252: With the limited degrees of freedom at hands, the theory we
1253: consider in the present work does not lead to an appearance
1254: of some effective charge in the theory. For the spurious solutions,
1255: the point of quantization is fixed as pure imaginary and can not be
1256: rotated by any arbitrary phase.
1257: Also, in the effective Hamiltonian
1258: given by Eq.(\ref{rightH}), $b$ can be thought as a part of effective
1259: coupling. For example, consider the term $g b \Phi^3_M$ in
1260: Eq.(\ref{rightH}) and build up an effective potential between two $\Phi_M$
1261: particles intermediating another $\Phi_M$ as an exchange-boson. Then, the
1262: coefficient $g b$ of $\Phi^3_M$ is an effective coupling of the triple
1263: vertices and the square of the coefficient would determine an overall sign
1264: of effective potential. The detailed form of the potential is of course
1265: determined only after taking into account the propagator of
1266: exchange-boson. However, the overall sign of effective potential
1267: determines whether the force between the two $\Phi_M$ particles is
1268: attractive or repulsive. Due to the sign change of the potential, if some
1269: bound states may not be formed (although they could be formed for the
1270: real $b$),
1271: then the corresponding energy eigenvalues would appear as complex values
1272: indicating the instability of those states. In some sense, the imaginary
1273: $b$ resembles a sort of optical potential which has been used rather
1274: frequently in the phenomenology of many-body nuclear physics.
1275: A recent work showing an example of how the imaginary coupling can be
1276: used in practice was presented in Ref.\cite{neuberg}.
1277:
1278: To show an overall picture of the effective potential in $(\phi^6)_{1+1}$
1279: theory, we present the dimensionless effective potential $V(\phi) = \Bigl
1280: [V^{class}(\phi^{class}_M) + \varepsilon_M
1281: \Bigr ] /m^2$ for the broken-symmetry solution and the spurious solution
1282: with $H/G = -0.209$($\alpha=-\beta$ in notations of
1283: Ref.\protect\cite{gep6}) in Fig.\ref{overallV6} for several
1284: different $G$ values.
1285: For simplicity, we didn't plot $V(\phi)$ for the symmetric phase $(b=0)$
1286: solution.
1287: In Fig.\ref{overallV6}, $\phi$ denotes
1288: $\phi^{class}_M$ and the symmetry $V^{class}(\phi^{class}_M) =
1289: V^{class}(-\phi^{class}_M)$ is manifest for the OR solutions
1290: with $B = b^2 > 0$, {\it i.e.} $Im(\phi) = 0$.
1291: Indeed, no matter what the sign of $B$ is, we can in general prove the
1292: symmetry $V(\phi) = V(-\phi)$ using the OR equations given by
1293: Eqs.(\ref{OR1}) and (\ref{OR2}).
1294: Thus, for the spurious solutions drawn off from $Im(\phi) = 0$ plane,
1295: one can realize the symmetry $V(\phi) = V(-\phi)$ by considering
1296: both planes of $Im(\phi) = \pm |b|$ although only the $Im(\phi) = +|b|$
1297: solutions are shown in Fig.\ref{overallV6}.
1298: The solid, long-dashed, short-dashed, dotted and
1299: dot-dashed lines stand for $G$=-0.838,-1.68,-2.51,-3.35 and -4.19
1300: ($\alpha$=-0.1,-0.2,-0.3,-0.4 and -0.5 in
1301: notations of Ref.\protect\cite{gep6}), respectively.
1302: For the coupling constants $G = $ -2.51 (short-dashed), -3.35 (dotted)
1303: and -4.19 (dot-dashed), the effective potentials corresponding to the
1304: three nontrivial solutions including
1305: a spurious solution with the imaginary $b$
1306: are shown in Fig.\ref{overallV6}:
1307: {\it e.g.} three dot-dashed lines (the spurious one is on the plane
1308: $Im(\phi) \neq 0$ and the other two real broken-symmetry solutions are on
1309: the plane
1310: $Im(\phi) = 0$) are shown for the single value of $G= -4.19$.
1311: For the comparison with the trivial solutions, we added the solid line
1312: for $G= -0.838$ and the long-dashed line for $G= -1.68$ which have the
1313: minimum only at $Re(\phi) = Im(\phi) =0$.
1314:
1315: In Figs.\ref{realV6III}-\ref{largeV6III}, the effective potentials
1316: for the spurious solutions are shown. Since the spurious solutions
1317: have $Im(\phi) \neq 0$, the corresponding effective potentials $V(\phi)$
1318: are complex. The real parts of effective potentials for the
1319: spurious solutions of $G=$-0.838(solid line), -1.68(long-dashed line),
1320: -2.51(short-dashed line), -3.35(dotted line) and -4.19(dot-dashed line)
1321: are shown in Figs.\ref{realV6III} and \ref{largeV6III} while the imaginary
1322: parts for the same coupling constants are shown in Fig.\ref{imV6III}.
1323: Also, the horizontal axes of Figs.\ref{realV6III}-\ref{largeV6III} are
1324: given by the real field $\Phi^{class}_M = \phi^{class}_M -b$ rather than
1325: the complex field $\phi^{class}_M$ due to the imaginary $b$.
1326: However, the effective potentials for $G=$ -2.51, -3.35 and -4.19 in
1327: Figs.\ref{realV6III}-\ref{largeV6III} correspond to the ones
1328: drawn off from the plane $Im(\phi)=0$ in Fig.\ref{overallV6}.
1329: As shown in Fig.\ref{vac6spur}, the energy density of the spurious
1330: solution gets the lowest among all the OR solutions
1331: for the small coupling such as $G=$ -0.838 and -1.68. Especially,
1332: the quasi-particle mass gets close to zero as $G$ becomes close to zero.
1333: This behavior may be noticed from Fig.\ref{realV6III} where the second
1334: derivative of $Re(V^{class})$ near $\phi^{class}_M = b$ is the smallest
1335: for $G=$ -0.838. The corresponding behaviors of $Im(V^{class})$ are
1336: shown in Fig.\ref{imV6III}. It is manifest from Fig.\ref{imV6III} that
1337: $Im(V^{class})$ contains only the odd powers of the real field
1338: $\Phi^{class}_M (= \phi^{class}_M -b)$, which can also be explicitly shown
1339: from our effective Hamiltonian given by Eq.(\ref{rightH}).
1340: The large field behaviors of $Re(V^{class})$ are shown in
1341: Fig.\ref{largeV6III}. We can see that the minima of $Re(V^{class})$
1342: appear in the large real field $\Phi^{class}_M$ for the spurious
1343: solutions, which one can also notice from the effective Hamiltonian
1344: (Eq.(\ref{rightH})) by keeping only the even powers of $\Phi_M$.
1345:
1346: In Fig.\ref{rabbit}, we show the domain of nontrivial solutions
1347: that have the vacuum energy density less than zero
1348: in the coupling space of $G$ and $H$.
1349: Without the spurious solutions, the domain of nontrivial OR solutions
1350: coincides with the GEP result (Fig.1 in Ref.\cite{gep6})
1351: \footnote{In Ref.\cite{gep6}, Stevenson
1352: and Roditi used somewhat different
1353: notations for couplings. Their couplings $\alpha$ and $\beta$ are related to
1354: our $G$ and $H$ as
1355: $\alpha =\frac{3G}{8\pi}$ and $\beta=\frac{45H}{8\pi^2}$, respectively.
1356: The boundary of allowed $\alpha$ and $\beta$ regions for the GEP
1357: were found as
1358: a curve (see Fig.1 of Ref.\cite{gep6}) that can be drawn by the
1359: following
1360: formulas: $\alpha = (2z+3+(z-3)e^z)/z^2, \beta = 3(z+2+(z-2)e^z)/z^3$ with
1361: $0<z<\infty$.} denoted by
1362: the solid line of boundary. Taking into account the spurious solutions,
1363: however, the domain of trivial solutions reduces to the
1364: shaded region of the phase diagram shown in Fig.\ref{rabbit}.
1365: The appearance of the spurious phase ($B = b^2 < 0$) in the domain where
1366: couplings are small may be interesting because it is quite opposite to the
1367: appearance of the normal nontrivial phase ($B > 0$), for which one would
1368: expect that the interaction term that leads to the
1369: formation of vacuum condensate is dominant at the larger coupling and
1370: therefore the nontrivial phases appear in the domain of large couplings.
1371: The presence of the nontrivial solution for any values of the couplings is
1372: also interesting in connection with the
1373: Haag's theorem \cite{haag}, according to which, if the vacuum
1374: state of a Lorentz invariant theory is unique, then all of its
1375: observables are identical with those of the free theory.
1376: Even very small admixture of $\phi^6$ interaction
1377: to the $\phi^4$ theory allows to have a set of vacuum states for any
1378: value of the $\phi^4$ coupling and thus to overcome the conditions of the
1379: theorem.
1380:
1381: In order to explore whether the spurious solution can be considered as
1382: an evidence of some real solution by adding more degrees of freedom,
1383: we applied the OR method
1384: to the two-component $(\phi^6)_{1+1}$ theory.
1385: In this model, the imaginary condensation of one component could be
1386: represeneted by a real condensation of the other component.
1387: However, we found that the imaginary-$b$ solution of the one-component
1388: model corresponds to imaginary-$b$ solution of two-component O(2) model,
1389: but not to the real solution.
1390: The matching was done by solving an arbitrary two-component model
1391: \begin{eqnarray}
1392: \label{hamo2}
1393: {\cal H}=N_m\Bigl ( \frac{1}{2}(\pi^2_1+\pi_2^2)+\frac{1}{2}((\nabla\phi_1)^2
1394: +(\nabla\phi_2)^2)
1395: +\frac{m^2}{2}(\phi_1^2+\phi_2^2)+ \\
1396: \frac{g}{4}(\phi^4_1+\phi^4_2) + h(\phi^6_1+\phi^6_2) +
1397: \beta(\frac{g}{2}\phi_1^2\phi_2^2 + 3h\phi_1^4\phi_2^2 + 3h\phi_1^2\phi_2^4)
1398: \Bigr ). \nonumber
1399: \end{eqnarray}
1400: For $\beta=0$ and 1, Eq.(\ref{hamo2}) corresponds to O(1)$\times$O(1) and
1401: O(2)
1402: models,
1403: respectively. The evolution of $b^2$ for a ``typical''
1404: value of $G$ and $H$ (we took $G=1, H=3$) is shown on Fig.\ref{bevol}.
1405: As it can be seen, the negative $b^2$ solution in O(1)$\times$O(1) never
1406: becomes positive as $\beta$ gets increased and even more spurious
1407: solutions appear in the limit of O(2). However, there is one-to-one
1408: correspondence between positive $b^2$ solutions in O(1)$\times$O(1) and
1409: O(2). We found that the same is true for any other $G$ and $H$.
1410: Thus, our most immediate exploration didn't realize the anticipated
1411: possibility that the negative $b^2$ solution in O(1)$\times$O(1) theory
1412: corresponds to a positive $b^2$ solution in O(2) theory. This may be a
1413: symptom of nonexistence of Goldstone
1414: mode in $1+1$ dimension known as the Coleman's theorem\cite{coleman1}.
1415: Thus, further exploration including the higher dimension may be necessary
1416: for the better understanding of the spurious solution.
1417:
1418: \begin{thebibliography}{99}
1419: \bibitem{CJT}
1420: J.M.Cornwall, R.Jackiw and E.Tomboulis, Phys. Rev. D{\bf 10}, 2428 (1974).
1421:
1422: \bibitem{efimov}
1423: G.~V.~Efimov,
1424: %``On A Phase Structure Of A Two-Dimensional N Component Phi**4 Field Theory,''
1425: Int.\ J.\ Mod.\ Phys.\ A {\bf 4}, 4977 (1989).
1426: %%CITATION = IMPAE,A4,4977;%%
1427:
1428: \bibitem{chang2}
1429: S.~J.~Chang,
1430: %``The Existence Of A Second Order Phase Transition In The Two-Dimensional Phi**4 Field Theory,''
1431: Phys.\ Rev.\ D {\bf 13}, 2778 (1976)
1432: [Erratum-ibid.\ D {\bf 16}, 1979 (1976)].
1433: %%CITATION = PHRVA,D13,2778;%%
1434:
1435: \bibitem{magruder}
1436: S.~F.~Magruder,
1437: %``The Existence Of Phase Transition In The (Phi**4) In Three-Dimensions Quantum Field Theory,''
1438: Phys.\ Rev.\ D {\bf 14}, 1602 (1976).
1439: %%CITATION = PHRVA,D14,1602;%%
1440:
1441: \bibitem{efbook}
1442: M.~Dineykhan, G.~V.~Efimov, G.~Ganbold and S.~N.~Nedelko,
1443: %``Oscillator representation in quantum physics,''
1444: Lect.\ Notes Phys.\ {\bf M26}, 1 (1995).
1445: %%CITATION = LNPHA,M26,1;%%
1446:
1447: \bibitem{ft7}
1448: P.M. Stevenson, Phys. Rev. D {\bf 30}, 1712 (1984).
1449:
1450: \bibitem{gep}
1451: P.M. Stevenson, Phys. Rev. D {\bf 35}, 2407 (1987).
1452:
1453: \bibitem{gep4}
1454: P.M. Stevenson, Phys.\ Rev.\ D {\bf 32}, 1389 (1985).
1455:
1456: \bibitem{gep6}
1457: P.~M.~Stevenson and I.~Roditi,
1458: %``The Gaussian Effective Potential. Iii. Phi**6 Theory And Bound States,''
1459: Phys.\ Rev.\ D {\bf 33}, 2305 (1986).
1460: %%CITATION = PHRVA,D33,2305;%%
1461:
1462: \bibitem{mp}
1463: H. Mishra and A. R. Panda, J. Phys. G : Nucl. Part. Phys. {\bf 18}, 1301 (1992).
1464: %``The Gaussian Effective Potential. 2. Lambda Phi**4 Field Theory,''
1465: %%CITATION = PHRVA,D32,1389;%%
1466:
1467: \bibitem{Chang}
1468: S.~J.~Chang, Phys.\ Rev.\ D {\bf 12}, 1071 (1975).
1469: %``Quantum Fluctuations In A Phi**4 Field Theory. 1. The Stability Of
1470: %The Vacuum.
1471:
1472: \bibitem{hartree}
1473: S.J.Chang and J.A.Wright, Phys.
1474: Rev. D {\bf 12}, 1595 (1975); S.J.Chang and T.-M.Yan, Phys. Rev.
1475: D {\bf 12}, 3225 (1975).
1476: %%CITATION = PHRVA,D12,1071;%%
1477:
1478: \bibitem{coleman}
1479: S.~R.~Coleman,
1480: %``Quantum Sine-Gordon Equation As The Massive Thirring Model,''
1481: Phys.\ Rev.\ D {\bf 11}, 2088 (1975).
1482: %%CITATION = PHRVA,D11,2088;%%
1483:
1484: \bibitem{lattice}
1485: W.~Loinaz and R.~S.~Willey,
1486: %``Monte Carlo simulation calculation of critical coupling constant for
1487: %continuum phi**4(2),''
1488: Phys.\ Rev.\ D {\bf 58}, 076003 (1998)
1489: [arXiv:hep-lat/9712008].
1490: %%CITATION = HEP-LAT 9712008;%%
1491:
1492: \bibitem{simon}
1493: B.Simon, {\it The $P(\phi)_2$ Euclidean (Quantum) Field Theory},
1494: Princeton University Press, New Jersey, 1974.
1495:
1496: \bibitem{Munoz}
1497: R.~Munoz-Tapia, J.~Taron and R.~Tarrach,
1498: %``The Uncertainty Of The Gaussian Effective Potential,''
1499: Int.\ J.\ Mod.\ Phys.\ A {\bf 3}, 2143 (1988).
1500: %%CITATION = IMPAE,A3,2143;%%
1501:
1502: \bibitem{wudka}
1503: U.~Ritschel,
1504: %``Variance Method In Quantum Field Theory And Exact Strong Coupling Limit Of Gaussian Effective Potential,''
1505: Z.\ Phys.\ C {\bf 54}, 297 (1992), \\
1506: %%CITATION = ZEPYA,C54,297;%%
1507: J.~Wudka,
1508: %``Variational Calculations And Renormalization,''
1509: Phys.\ Rev.\ D {\bf 37}, 1464 (1988).
1510: %%CITATION = PHRVA,D37,1464;%%
1511:
1512: \bibitem{ji-mish} C.-R.Ji and Y.Mishchenko, Phys. Rev. D {\bf 65}, 096015
1513: (2002); Phys. Rev. D {\bf 64}, 076004 (2001).
1514:
1515: \bibitem{Kadanoff}
1516: L.P.Kadanoff and P.C.Martin, Phys. Rev. {\bf 124}, 670 (1961);
1517: D.J.Thouless, ``The Quantum Mechanics of Many-Body Systems", Second Edition,
1518: Academic Press, New York and London, 1972.
1519:
1520: \bibitem{neuberg}
1521: N.E.Lighterink and B.L.G.Bakker, hep-ph/0010167.
1522:
1523: \bibitem{haag}
1524: R.~Haag, Dan.\ Mat.\ Fys.\ Medd. {\bf 29}, No.\ 12, 1 (1955), \\
1525: A.~S.~Wightman and S.~S.~Schweber, Phys.\ Rev. {\bf 98}, 812 (1955).
1526:
1527: \bibitem{coleman1}
1528: S.~R.~Coleman,
1529: %``There Are No Goldstone Bosons In Two-Dimensions,''
1530: Commun.\ Math.\ Phys.\ {\bf 31}, 259 (1973).
1531: %%CITATION = CMPHA,31,259;%%
1532: \end{thebibliography}
1533:
1534: \newpage
1535:
1536: \begin{figure}[htb]
1537: \centering
1538: \begin{minipage}[c]{0.45\hsize}
1539: \epsfig{figure=phi4.eps, width=\hsize}
1540: \caption{The nontrivial mass solutions for OR (solid and dashed
1541: lines), GEP (dotted line) and Hartree methods (squares correspond to the
1542: 4$\pi c^2<3$ solution, circles to the 4$\pi c^2>3$ one). Here, $G=g/m^2$ and
1543: for $G < 9.04$ there are no nontrivial solutions.
1544: \label{geff}}
1545: \end{minipage}
1546: \hspace*{0.5cm}
1547: \begin{minipage}[c]{0.45\hsize}
1548: \vspace*{5mm}
1549: \epsfig{figure=e_h0.eps, width=\hsize}
1550: \caption{Vacuum energy densities of the nontrivial OR solutions
1551: (solid and dashed
1552: lines), GEP (dotted line) and Hartree methods (squares correspond to the
1553: 4$\pi c^2<3$ solution, circles to the 4$\pi c^2>3$ one). Here, $G=g/m^2$ and
1554: for $G < 9.04$ there are no nontrivial solutions.
1555: \label{energh0}}
1556: \end{minipage}
1557: \end{figure}
1558:
1559:
1560: \begin{figure}[htb]
1561: \centering
1562: \epsfig{figure=g4new.eps,width=0.45\hsize}
1563: \caption{The effective classical potential corresponding to
1564: Fig.~\protect\ref{geff} with $G =$ 8, 10 and 12. Although $G = 8$
1565: has only a trivial effective potential denoted by a solid line,
1566: both $G =$ 10 and 12 have two nontrivial effective potentials
1567: denoted by I and II.\label{g4new}}
1568: \end{figure}
1569:
1570:
1571: \begin{figure}[h]
1572: \centering
1573: \begin{minipage}[c]{0.45\hsize}
1574: \epsfig{file=sol_0.eps,width=\hsize}
1575: \caption{Nontrivial solutions of the OR equations for $\phi^4$ and
1576: $\phi^6$ theory, symmetric phase.}
1577: \label{phasesnew1}
1578: \end{minipage}
1579: \hspace*{0.5cm}
1580: \begin{minipage}[c]{0.45\hsize}
1581: \epsfig{file=sol.eps,width=\hsize}
1582: \caption{Nontrivial solutions of the OR equations for $\phi^4$ and
1583: $\phi^6$ theory, broken-symmetry phase.}
1584: \label{phasesnew2}
1585: \end{minipage}
1586: \end{figure}
1587:
1588: \begin{figure}[htb]
1589: \centering
1590: \epsfig{file=sol_fake.eps,width=0.45\hsize}
1591: \caption{Nontrivial solutions of the OR equations for $\phi^4$ and
1592: $\phi^6$ theory, imaginary $b$ phase.}
1593: \label{phasesnew3}
1594: \end{figure}
1595:
1596: \begin{figure}[htb]
1597: \centering
1598: \begin{minipage}[c]{0.45\hsize}
1599: \epsfig{file=e_0.eps,width=\hsize}
1600: \end{minipage}
1601: \hspace*{0.5cm}
1602: \begin{minipage}[c]{0.45\hsize}
1603: \epsfig{file=e_0a.eps,width=\hsize}
1604: \end{minipage}
1605: \caption{Vacuum energy density of nontrivial solutions
1606: of the OR equations for $\phi^4$ and $\phi^6$ theory, symmetric phase.
1607: The two plots (right and left) are identical but presented in two different
1608: angles of view. The $90^{\circ}$ clockwise rotation of the left plot
1609: around the axis of $\varepsilon_M/m^2$ is identical to the right plot.}
1610: \label{vac6sym}
1611: \end{figure}
1612:
1613: \begin{figure}[h]
1614: \centering
1615: \begin{minipage}[c]{0.45\hsize}
1616: \epsfig{file=e_b.eps,width=\hsize}
1617: \caption{Vacuum energy density of nontrivial solutions
1618: of the OR equations for $\phi^4$ and
1619: $\phi^6$ theory, broken-symmetry phase.}
1620: \label{vac6brok}
1621: \end{minipage}
1622: \hspace*{0.5cm}
1623: \begin{minipage}[c]{0.45\hsize}
1624: \epsfig{file=e_b_im.eps,width=\hsize}
1625: \caption{Vacuum energy density of nontrivial solutions
1626: of the OR equations for $\phi^4$ and
1627: $\phi^6$ theory, imaginary $b$ phase.}
1628: \label{vac6spur}
1629: \end{minipage}
1630: \end{figure}
1631:
1632: \begin{figure}[h]
1633: \centering
1634: \begin{minipage}[c]{0.45\hsize}
1635: \epsfig{file=chi1.eps,width=\hsize}
1636: \caption{Dimensionless couplings of the symmetric perturbative
1637: solution for $G=0$.}
1638: \label{coup1}
1639: \end{minipage}
1640: \hspace*{0.5cm}
1641: \begin{minipage}[c]{0.45\hsize}
1642: \epsfig{file=chi2.eps,width=\hsize}
1643: \caption{Dimensionless couplings of the broken-symmetry perturbative
1644: solution for $G=0$.}
1645: \label{coup2}
1646: \end{minipage}
1647: \end{figure}
1648:
1649: \begin{figure}[h]
1650: \centering
1651: \begin{minipage}[c]{0.45\hsize}
1652: \epsfig{file=couplings.eps,width=\hsize}
1653: \caption{Dimensionless couplings of the broken-symmetry perturbative
1654: solution versus the dimensionless coupling of the symmetric
1655: solution ($G=0$).}
1656: \label{couplings}
1657: \end{minipage}
1658: \hspace*{0.5cm}
1659: \begin{minipage}[c]{0.45\hsize}
1660: \epsfig{file=eff_pot_sbs0.eps,width=\hsize}
1661: \caption{Classical potential of broken-symmetry (BS) and symmetric (S)
1662: theories at $\chi^{(6)}_s=5.85\cdot 10^{-3}$ ($H=100, G=0$).}
1663: \label{effpot1}
1664: \end{minipage}
1665: \end{figure}
1666:
1667: \begin{figure}[h]
1668: \centering
1669: \begin{minipage}[c]{0.45\hsize}
1670: \epsfig{file=eff_pot_sbs1.eps,width=\hsize}
1671: \caption{Classical potential of broken-symmetry (BS) and symmetric (S)
1672: theories at $\chi^{(6)}_s=1.33\cdot 10^{-3}$ ($H=10^6, G=0$).}
1673: \label{effpot2}
1674: \end{minipage}
1675: \hspace*{0.5cm}
1676: \begin{minipage}[c]{0.45\hsize}
1677: \epsfig{file=eff_pot_sbs15.eps,width=\hsize}
1678: \caption{Classical potential of broken-symmetry (BS) and symmetric (S)
1679: theories at $\chi^{(6)}_s=1.02\cdot 10^{-3}$ ($H=10^{15}, G=0$).}
1680: \label{effpot3}
1681: \end{minipage}
1682: \end{figure}
1683:
1684: \begin{figure}[htb]
1685: \centering
1686: \epsfig{file=g0gcphi4.eps, width=0.8\hsize, angle=270}
1687: \caption{(a) $x$ axis is $g_0$ and $y$ axis is $g_c$ (solid line)
1688: and $g_{c'}$ (dot). (b) $x$ axis is $g_0$ and $y$ axis is
1689: $\frac{g_0}{g_c}$ (solid line) and $\frac{g_0}{g_{c'}}$ (dot).}
1690: \label{app}
1691: \end{figure}
1692:
1693: \begin{figure}[hbt]
1694: \centering
1695: \epsfig{file=potall.eps,width=0.7\hsize}
1696: \caption{ The dimensionless effective potential $V(\phi) =
1697: \Bigl[V^{class}+\varepsilon_M \Bigr]/m^2$ for $H/G =
1698: -0.209$($\alpha=-\beta$ in
1699: notations of Ref.\protect\cite{gep6}), where $\phi$ means
1700: $\phi^{class}_M$. $x$ axis is $Re(\phi)$,
1701: $y$ axis is $Im(\phi)$ and $z$ axis is $V(\phi)$. In general, for
1702: the complex $\phi$ the potential $V(\phi)$ is complex.
1703: For $G=$-0.838(solid line) and -1.68(long-dashed line),
1704: we show only the trivial solutions ($t=0$ and $b=0$) for the comparison with
1705: the
1706: nontrivial solutions for $G=$ -2.51(short-dashed line),-3.35(dotted line)
1707: and -4.19(dot-dashed line).
1708: For each coupling constant, $G=$ -2.51(short-dashed line), -3.35(dotted
1709: line) and -4.19(dot-dashed line), there are two real $b$
1710: broken-symmetry solutions
1711: and one pure imaginary $b$ solution. The two real $b$
1712: broken-symmetry solutions for each
1713: coupling constant($G=$-2.51,-3.35 and -4.19) are shown on the $Im(\phi)=0$
1714: plane. For the pure imaginary $b$ solutions,
1715: the potential graphs are drawn
1716: on the planes off from the $Im(\phi)=0$ plane with the different
1717: imaginary $b$ values depending on the $G$ values.}
1718: \label{overallV6}
1719: \end{figure}
1720:
1721: \begin{figure}[hbt]
1722: \centering
1723: \begin{minipage}[c]{0.45\hsize}
1724: \epsfig{file=fig7f.eps,width=\hsize}
1725: \caption{ Real part of effective potentials for the spurious solutions
1726: of G=-0.838(solid line), -1.68(long-dashed line), -2.51(short-dashed
1727: line), -3.35(dotted line) and -4.19(dot-dashed line).}
1728: \label{realV6III}
1729: \end{minipage}
1730: \hspace*{0.5cm}
1731: \begin{minipage}[c]{0.45\hsize}
1732: \epsfig{file=fig7h.eps,width=\hsize}
1733: \caption{ Imaginary part of effective potentials for the spurious solutions
1734: of G=-0.838(solid line), -1.68(long-dashed line), -2.51(short-dashed
1735: line), -3.35(dotted line) and -4.19(dot-dashed line).}
1736: \label{imV6III}
1737: \end{minipage}
1738: \end{figure}
1739:
1740: \begin{figure}[hbt]
1741: \centering
1742: \epsfig{file=fig7z.eps,width=0.45\hsize}
1743: \caption{ Real part of effective potentials for the spurious solutions
1744: of G=-0.838(solid line), -1.68(long-dashed line), -2.51(short-dashed
1745: line), -3.35(dotted line) and -4.19(dot-dashed line) in the larger scale
1746: of classical fields. The blow-up of the results in a tiny region of
1747: $\phi^{class}_M -b < 0.025 $ is shown in Fig.\protect\ref{realV6III}.}
1748: \label{largeV6III}
1749: \end{figure}
1750:
1751:
1752: \begin{figure}[htb]
1753: \centering
1754: \begin{minipage}[c]{0.45\hsize}
1755: \epsfig{file=diagr.eps,width=\hsize}
1756: \caption{The domain of trivial solutions in $G$, $H$ parameter space
1757: (shadowed area). If the spurious solution shown in
1758: Fig.\protect\ref{phasesnew3} is not taken into account, then the domain
1759: enlarges to the inside of the
1760: entire area bounded by the external solid line, which is equivalent to
1761: the GEP result shown in Fig.1 of Ref.\protect\cite{gep6}.}
1762: \label{rabbit}
1763: \end{minipage}
1764: \hspace*{0.5cm}
1765: \begin{minipage}[c]{0.45\hsize}
1766: \epsfig{file=bet.eps,width=\hsize}
1767: \caption{Correspondence between solutions of OR equations for O(1)$\times $O(1)
1768: and O(2) $(\phi^6)_{1+1}$ theories at $G=1.0$, $H=3.0$.}
1769: \label{bevol}
1770: \end{minipage}
1771: \end{figure}
1772:
1773:
1774: \end{document}
1775: