1: %% put new commands here
2: \chapter{Higher Derivative Field Theories on the Lattice}
3: \label{ch:BOA}
4:
5: \section{The Naive Lattice Action and Phase Diagram}
6:
7: The need of a lattice for the higher derivative scalar field theory
8: presented in the previous chapter is not
9: for the purpose of regularization, but rather,
10: to make the degree of freedom finite so that a nonperturbative study
11: of the model can be performed in computer
12: simulations. The lattice spacing $a$
13: introduces a new short distance energy scale with the associated
14: momentum cutoff $\Lambda=\pi/a$. In order to recover the higher
15: derivative field theory in the continuum, we would have to work
16: towards the $\Lambda/M \rightarrow \infty$ limit with a fixed ratio
17: of $M/m_H$. The lattice action we choose to study \cite{hhiggs4,dallas4} is
18: \be
19: {\cal L}_E=-\kappa \phi(x)(-\Box-{\Box^3 \over M^4})\phi(x)
20: +(1-8\kappa)\phi(x)^2 -\lambda(\phi(x)^2-1)^2 ,
21: \ee
22: where the $\Box$ is the lattice Laplace operator. The phase structure
23: of this lattice model is quite similar to the conventional $O(N)$ scalar
24: field theory. It has two phases as shown in
25: Figure~(\ref{fig:ch4.phase}). The $O(N)$ symmetric phase
26: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Phase Diagram
27: \begin{figure}[htb]
28: \vspace{5mm}
29: \centerline{ \epsfysize=3.0cm
30: \epsfxsize=5.0cm \epsfbox{Phase.ps}}
31: \vspace{17mm}
32: \caption{ The phase diagram of the lattice model at infinite bare
33: coupling. Data points are obtained from
34: Monte Carlo simulations.
35: The dotted line is calculated in the large-N expansion.
36: The solid line displays a fixed $M_R/m_H$ ratio towards the continuum
37: limit of the higher derivative theory.}
38: \label{fig:ch4.phase}
39: \end{figure}
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% End of Phase Diagram
41: is separated from the broken phase with residual $O(N-1)$ symmetry
42: by a second order phase transition line for every value of the
43: lattice coupling constant $\lambda$ in the $(\kappa,M)$ plane.
44: Near the critical line, we expect to recover the continuum theory
45: without the lattice artifacts. However, the critical behavior of
46: our model is more complicated than the conventional $O(N)$ model. It can
47: represent different universal continuum theories along different
48: paths towards the critical line. Tuning the value of $\kappa$ towards
49: the critical line for any fixed value of $M$ corresponds to the
50: trivial field theory in the continuum. In this limit, the operator
51: $\phi\Box^3\phi$ becomes irrelevant in the critical region.
52: However, if we tune the value of $\kappa$ towards the critical line
53: in such a way that the ratio $M_R/m_H$ remains fixed, we will recover
54: the continuum higher derivative field theory with the corresponding
55: ratio of the ghost mass parameter and the Higgs mass. In this limit
56: the operator $\phi\Box^3\phi$ cannot be viewed as an irrelevant
57: operator \cite{neub4} in the Lagrangian. Thus,
58: it becomes clear, from the
59: discussion above, that if we want to study the higher derivative
60: field theory, we have to work towards the second limit.
61:
62: In the practical application, however, this limit is not very
63: easy to arrange. One reason is that if we want our results
64: to represent the continuum results, we have to keep the ghost mass
65: parameter $M$ reasonably small in lattice units
66: in order to get rid of the lattice
67: effects associated with it. On the other hand, we need to put the
68: Higgs mass below the ghost mass parameter. Therefore, we are
69: very restricted in the parameter space. On the one hand, making
70: the Higgs mass smaller will lead to huge finite size effects for
71: the practical lattice sizes; on the other hand, making the Higgs mass
72: larger will push up the ghost mass and will result in large lattice
73: effects. So we have a rather narrow range in the Pauli-Villars
74: correlation length $M/m_H$. Typical values we took in the beginning
75: of our simulation were: $M=0.8\sim1.0$, $m=0.3\sim0.4$.
76: This, of course, was unsatisfactory because the ratio
77: $M/m_H=2\sim 3$ is too narrow of a range. If we
78: view this theory as a Pauli-villars regulated theory, for example, we
79: would hope to see the conventional scaling behavior in the
80: large $M/m_H$ limit. It turns out that the scaling form
81: may apply only for rather large $M/m_H$ values which is
82: impossible for us to investigate using this naive lattice action.
83: Also, due to this restricted range, it was also impossible for
84: us to study the scattering phase shift profile of the model.
85: This type of analysis offers us a very good way of extracting
86: the mass value for an unstable particle in the finite box ( see
87: Chapter~(\ref{ch:LUSC}) for full discussion ).
88: This restriction in the parameters is purely due to the
89: introduction of the underlying lattice structure. When
90: we were able to eliminate most of the lattice effects, we were then
91: able to enlarge our parameter space quite substantially.
92: Therefore, the need for an improved lattice action
93: becomes quite obvious.
94:
95: \section{ The Improved Lattice Action }
96:
97: Improving the lattice action so that it has better
98: Euclidean invariance was studied long ago \cite{syman4}. Our choice
99: of the improvement corresponds to modifying the lattice
100: Laplacian so that it resembles the continuum
101: Laplacian. Therefore, we take
102: \be
103: p_{I}^2= \hat{p}^2 + {a_1 } \sum_{\mu} \hat{p}_{\mu}^4
104: + {a_2 } \sum_{\mu} \hat{p}_{\mu}^6
105: + {a_3 } \sum_{\mu} \hat{p}_{\mu}^8
106: + {a_4 } \sum_{\mu} \hat{p}_{\mu}^{10}
107: + {a_5 } \sum_{\mu} \hat{p}_{\mu}^{12}
108: + {a_6 } \sum_{\mu} \hat{p}_{\mu}^{14} ,
109: \ee
110: where the coefficients are given by the following table
111: \begin{table}[htb]
112: \begin{tabular}{|cccccccccc|}
113: \hline
114: $a_1$
115: &$a_2$
116: &$a_3$
117: &$a_4$
118: &$a_5$
119: &$a_6$
120: &$a_7$
121: &$a_8$
122: &$a_9$
123: &$a_{10}$ \\
124: \hline
125: ${1\over 12}$
126: &${1\over 90}$
127: &${1\over 560}$
128: &${1\over 3150}$
129: &${1\over 16632}$
130: &${1\over 84084}$
131: &${1\over 411840}$
132: &${1\over 1969110}$
133: &${1\over 9237800}$
134: &${1\over 42678636}$\\
135: \hline
136: \end{tabular}
137: \end{table}
138:
139: In fact, we calculated the renormalized coupling constant in the
140: large $N$ limit and we found that this improved action significantly
141: decreased the lattice effects. With this improved lattice action, even at
142: $M=2.0$, there was negligible lattice effects on the large $N$
143: results. The phase diagram of the improved action is similar to
144: the naive action.
145:
146: The improved action offers us another power: the possibility
147: of performing a phase shift simulation on the higher derivative
148: $O(N)$ model. This is the subject in Chapter~(\ref{ch:LUSC}).
149: As we will demonstrate, without the improved action, we are in the
150: parameter range that is impossible for this type of simulation because
151: we would need unrealisticly large lattices to extract the
152: phase shift. With the improved action, this type of calculation
153: becomes possible.
154:
155: \section{ The Rotator States and Born Oppenheimer Approximation }
156:
157: Studying the higher derivative $O(N)$ model in the broken phase and
158: the corresponding Higgs mass problem requires a better understanding
159: of the symmetry breaking mechanism in the finite volume. In fact this
160: is already an important issue in the conventional $O(N)$ theory
161: without the higher derivative terms added.
162: The symmetry breaking mechanism has been understood very well
163: in the infinite volume limit. However, it has not been answered
164: satisfactorily in the $O(N)$ model in a finite volume.
165:
166: There are several complications. First of all, the notion of
167: symmetry breaking in the infinite volume cannot be applied to
168: a system in a finite volume. Strictly speaking, in a finite volume,
169: the symmetry is never broken.
170: Secondly, it turns out that the dynamics of the zeromode
171: are crucial for the understanding, and the zeromode
172: is coupled to other modes in a complicated
173: way. For the one component $\phi^4$ theory, Hartree type of
174: approximation will give us a very good description of the
175: symmetry breaking. For the $O(N)$ model, extra care must
176: be paid to the motion of the zeromode and new approximation
177: schemes are needed for the understanding of the problem.
178: This section consists of several parts.
179: In the first part, we will review what is known to the
180: symmetry breaking in a ordinary one-component $\phi^4$
181: theory in the broken phase. It turns out that this is
182: a very instructive model to study. In the second part,
183: the conventional $O(N)$ model is studied in the broken phase.
184: Here we introduce the Born-Oppenheimer Approximation (or Adiabatic
185: Approximation) and fully investigate
186: the dynamics of the zeromode.
187: In the third part, we consider some important
188: applications of the Born-Oppenheimer Approximation.
189: The machinery is applied to the ground state
190: and higher energy excited states.
191: The rotator correction to the energy of these states
192: is calculated. This will serve as a theoretical guide line to
193: the analysis of the simulation results in Chapter~(\ref{ch:SIMU}).
194: Then,
195: the higher derivative $O(N)$ theory is presented in the
196: next section.
197:
198: \subsection{ Symmetry Breaking of the One-component $\phi^4$ Model }
199:
200: Consider the one component $\phi^4$ theory in a cubic box. The
201: Hamiltonian of the theory is given by
202: \be
203: {\cal H} = {1 \over 2} \pi^2 +{1 \over 2} (\nabla \phi)^2 -
204: {1 \over 2} \mu_0^2 \phi^2 +{\lambda_0 } \phi^4 .
205: \ee
206: This Hamiltonian is obviously invariant under the
207: change $\phi \rightarrow -\phi$. That is to say
208: that one can construct a parity operator $P$, which
209: flips the sign of the $\phi$ field and it
210: commutes with the Hamiltonian. Therefore,
211: all the eigenstates of the Hamiltonian can be
212: chosen to have a definite parity.
213:
214: We can build up two approximate ground states of the
215: Hamiltonian $|\pm\rangle$, which are Gaussian wavefunctions centered at
216: $\pm v$ respectively.
217: This picture is very well illustrated
218: by the so-called ``Hartree approximation''. We start with a trial
219: wave functional which is a Gaussian
220: \be
221: \Psi(\phi) = N \exp \left( -{1 \over 2}
222: (\phi(x)-v)G^{-1}(x,y)(\phi(y)-v) \right)
223: \ee
224: where $v$,$G^{-1}(x,y)$ are variational parameters and the summation
225: over $x$ and $y$ is implied. The approximate ground state of the
226: system can be found by using the minimization condition of the
227: energy. This condition in the broken phase will give us two
228: solutions for the parameter $v$, namely,
229: $v=\pm \sqrt{\mu^2_0/4\lambda_0}$ and the propagator $G^{-1}(x,y)$
230: is given by
231: \be
232: G(x,y)={1 \over 2} \int {d^3k \over (2\pi)^3}
233: {e^{i \bk \cdot \bx} \over \sqrt{\bk^2 + m^2_R} }
234: \ee
235: If we denote these two states as $| \pm \rangle$ then we see
236: they satisfy the following properties
237: \be
238: P |\pm \rangle =|\mp \rangle, \;\; \langle +|-\rangle=e^{-v^2m_RL^3} .
239: \ee
240: Note that the
241: states $| \pm \rangle$ are not orthogonal to
242: each other in the finite volume.
243: The true ground state and the first excited state are given by the
244: symmetric and antisymmetric linear combination of these two
245: states
246: \be
247: |0\rangle = {1 \over \sqrt{2}} (|+\rangle + |-\rangle ) ,
248: \;\;\;\;
249: |1\rangle = {1 \over \sqrt{2}} (|+\rangle - |-\rangle ) .
250: \ee
251: The true ground state is a parity even
252: state, while the first excited state is a parity odd state.
253: The energy difference between the two is exponentially
254: small when the volume is large.
255: This means that if the
256: system were started at one of the minimum, after a long enough
257: time, there is a finite
258: probability of finding the system
259: tunneled to the other minimum. The typical time
260: scale for this is $1/\Delta E$, where $\Delta E$ is the
261: energy difference between the ground and the first excited state.
262:
263: If we use a infinite volume, the state
264: $|+\rangle$ would be exactly orthogonal to the state
265: $|- \rangle$; then the system starting from one
266: particular minimum of the potential as the true vacuum
267: will stay there, without knowing
268: the other one and the $\phi \rightarrow -\phi$
269: symmetry is broken. However, in a
270: finite but large volume, the system
271: will stay around one minimum for such a
272: long enough time that we may say
273: the symmetry is ``almost broken''.
274:
275: In the one component model, the symmetry is a discrete symmetry
276: and the Hartree approximation gives us a very good understanding
277: of the symmetry breaking mechanism in the finite volume.
278: Nonperturbative works have also been done to measure the
279: energy gap between the ground state and the first excited
280: state, which is related to the surface tension of the
281: system.
282:
283: \subsection{$O(N)$ Model: General Setup}
284:
285: The situation is much more complicated when we try to do
286: a similar analysis for the $O(N)$ model. The main reason
287: is that the symmetry is a continuous symmetry, therefore,
288: the dynamics of the zeromode is much more complicated.
289:
290: We could first try out the Hartree approximation, but it
291: will not give us the right energy spectrum of the theory.
292: This is because the Hartree approximation
293: treats every mode of the system equally. In
294: the one component model this is valid, but it is not valid for the
295: $O(N)$ model. In the $O(N)$ model there exists one special mode,
296: that is, the direction of the zero Fourier mode which
297: can be characterized by an $O(N)$ unit vector. In a large but
298: finite volume, this mode
299: is a slow varying mode when compared with the other
300: modes. It is the counter part of the parity operator
301: in the one component model. The only difference is that,
302: in the one component
303: model, the parity only takes discrete values and is not
304: dynamical. In the O(N) model, however, this unit vector lives on
305: a $(N-1)$-sphere and has its own dynamics.
306: Therefore, we expect that the Born-Oppenheimer
307: Approximation (BOA), also known as the
308: Adiabatic Approximation, will give us a very good description
309: of the zeromode dynamics.
310:
311: The Born-Oppenheimer Approximation was first introduced
312: in the study of diatomic molecules. In the molecular
313: problem, there are two types of degrees of freedom. The motion
314: of the electron is called ``fast'', and the motion of
315: the nucleus is called ``slow''. Therefore, when solving the
316: energy eigenvalues of the system, one should first nail down
317: the slow variable, namely the configuration of the
318: nucleus, and solve the fast variable spectrum. In this step,
319: the configuration of the slow variable is treated as an
320: external field. The eigenvalues and eigenstates that
321: come out will, in general, depend on the prescribed
322: configuration of the slow variable. These eigenvalues
323: are then taken back into the Schr\"odinger equation for
324: the slow variables as the effective potential, which
325: reflects the feedback of the fast variable to the
326: slow variable. Finally, the Schr\"odinger equation for
327: the slow variable is solved to get the spectrum of the
328: molecule.
329:
330: The spectrum of the molecule has a three fold hierarchy:
331: electron energy, oscillation energy and rotational energy.
332: These energy gaps are characterized by different powers of
333: a small parameter, which is the ratio $m_e /m_N$, where
334: $m_e$ is the mass of the electron and $m_N$ is the mass
335: of the nucleus. Born-Oppenheimer is a very good
336: approximation for the molecule, since this ratio is so small.
337: It is not hard to imagine that a Hartree approximation to the
338: molecular problem would be a poor choice, since it
339: treats the electron (fast variable) and the nucleus (slow variable)
340: equally, while ignoring the enormous difference in the mass of the two.
341: Similarly, in our application, Hartree is a poor approximation
342: for the same reason.
343: To get the right picture, one has to separate the special
344: zeromode and use the Born-Oppenheimer type of approximation.
345:
346: In our model, we will treat the direction of the
347: zeromode as the only slow variable. We will use the
348: same Born-Oppenheimer type of spirit to solve for
349: the energy levels of our model.
350:
351: We begin with the Hamiltonian
352: \be
353: H = \sum_{\bx}{1 \over 2} \pi^a\pi^a +{1 \over 2} \nabla \phi^a
354: \nabla \phi^a -
355: {1 \over 2} \mu_0^2 \phi^a\phi^a
356: +{\lambda_0 } (\phi^a \phi^a)^2 ,
357: \ee
358: where for convenience we have discretize the system on a
359: cubic lattice. The operator $\pi^a(\bx)$ is just the
360: derivative operator $(-i)\partial/\partial \phi^a(\bx)$ in the
361: field variable diagonal representation.
362: Two classes of symmetry operators that commute with this Hamiltonian are
363: very important. First, there are
364: global $O(N)$ symmetry generators $Q^{ab}$ given by
365: \be
366: Q^{ab}=\sum_{\bx} \phi^{a}(\bx)\pi^{b}(\bx)
367: -\phi^{b}(\bx)\pi^{a}(\bx)
368: \ee
369: We also have the $3$-momentum operators $P_i$
370: \be
371: P_i = \sum_x [\phi^a(\bx+\bei)-\phi^{a}(\bx)]\pi^{a}(\bx) ,
372: \ee
373: where $\bei$ is the unit vector in the $i$ direction. It is
374: trivial to verify that these operators commute with the full
375: Hamiltonian hence are symmetries of the theory.
376:
377: We now introduce the Fourier modes of the field variable,
378: \ba
379: \phi^a(\bx) &=& \bar{\phi}^a +
380: {1 \over \sqrt{V}} \sum_{\bk > 0}
381: \phi^a_{\bk} e^{i\bk \cdot \bx}
382: +\phi^{a*}_{\bk} e^{-i\bk \cdot \bx} ,
383: \nonumber \\
384: \pi^a(\bx) &=& {(-i) \over V}{\partial \over \partial \bar{\phi}^a}
385: +{(-i) \over \sqrt{V}} \sum_{\bk > 0}
386: e^{-i\bk \cdot \bx}{\partial \over \partial \phi^a_{\bk}}
387: +e^{+i\bk \cdot \bx}{\partial \over \partial \phi^{a*}_{\bk}} ,
388: \ea
389: where $V=L^3$ is the $3$-volume of the box. As we mentioned
390: above, the zero mode $\bar{\phi}^a$ plays a very important role
391: in the broken phase. Therefore we have singled out this mode from the
392: nonzero momentum modes. Let us define:
393: \ba
394: \bar{\phi}^a &=& (v+\sigma) n^a , \;\;\; n^a n^a=1 ,
395: \nonumber \\
396: P_{L}^{ab}&=&n^a n^b , \;\;\;\; P_{T}^{ab}=\delta^{ab} -n^a n^b ,
397: \ea
398: where $v=\mu_0^2/4\lambda_0$ is the
399: vev of the theory.
400: In the Born-Oppenheimer type of
401: approach, we will treat the direction of $\bar{\phi}^a$, namely
402: $n^a$, as the only
403: slow varying variable and treat the rest as fast variables.
404: The justification of this will be seen shortly.
405: Then the Hamiltonian can be expressed in terms of
406: these Fourier modes.
407:
408: We use the radial variables for the mode
409: $\bar{\phi}^a$. Thus we write the wavefunctional of
410: the system as $\Psi=\rho^{-(N-1)/2}\psi$ and the
411: effective Hamiltonian for $\psi$ will contain only the
412: second derivative with respect to $\rho$. For the nonzero
413: Fourier modes, let us introduce the creation and
414: annihilation operators as
415: \ba
416: \label{eq:ch4.creatl}
417: L^a_{\bk} &=& {1 \over \sqrt{2}}P_{L}^{ab}
418: (\sqrt{\Omega_{\bk}} \phi^a_{\bk}
419: + {1 \over \sqrt{\Omega_{\bk}}}
420: {\partial \over \partial\phi^{b*}_{\bk} } ) ,
421: \nonumber \\
422: L^{a\dagger}_{\bk} &=& {1 \over \sqrt{2}}P_{L}^{ab}
423: (\sqrt{\Omega_{\bk}} \phi^{a*}_{\bk}
424: - {1 \over \sqrt{\Omega_{\bk}}}
425: {\partial \over \partial\phi^{b}_{\bk} } ) ,
426: \nonumber \\
427: L^{a}_{-\bk} &=& {1 \over \sqrt{2}}P_{L}^{ab}
428: (\sqrt{\Omega_{\bk}} \phi^{a*}_{\bk}
429: + {1 \over \sqrt{\Omega_{\bk}}}
430: {\partial \over \partial\phi^{b}_{\bk} } ) ,
431: \nonumber \\
432: L^{a\dagger}_{-\bk} &=& {1 \over \sqrt{2}}P_{L}^{ab}
433: (\sqrt{\Omega_{\bk}} \phi^a_{\bk}
434: + {1 \over \sqrt{\Omega_{\bk}}}
435: {\partial \over \partial\phi^{b*}_{\bk} } ) ,
436: \ea
437: where $\Omega_{\bk}=\sqrt{m_0^2+\bk^2}$ is the higgs excitation.
438: We can define the Higgs creation and annihilation operators as
439: \be
440: h_{\bk}=n^a L^a_{\bk}, \;\;\;\;\;h^{\dagger}_{\bk}=n^a L^{a\dagger}_{\bk},
441: \;\;\;\;\;
442: \sigma= {1 \over \sqrt{2Vm_0}} (h_0 + h^{\dagger}_0) .
443: \ee
444: Similarly, we can define the transverse Goldstone creation and
445: annihilation operators as
446: \ba
447: \label{eq:ch4.creatt}
448: T^a_{\bk} &=& {1 \over \sqrt{2}}P_{T}^{ab}
449: (\sqrt{\omega_{\bk}} \phi^a_{\bk}
450: + {1 \over \sqrt{\omega_{\bk}}}
451: {\partial \over \partial\phi^{b*}_{\bk} } ) ,
452: \nonumber \\
453: T^{a\dagger}_{\bk} &=& {1 \over \sqrt{2}}P_{T}^{ab}
454: (\sqrt{\omega_{\bk}} \phi^{a*}_{\bk}
455: - {1 \over \sqrt{\omega_{\bk}}}
456: {\partial \over \partial\phi^{b}_{\bk} } ) ,
457: \nonumber \\
458: T^{a}_{-\bk} &=& {1 \over \sqrt{2}}P_{T}^{ab}
459: (\sqrt{\omega_{\bk}} \phi^{a*}_{\bk}
460: + {1 \over \sqrt{\omega_{\bk}}}
461: {\partial \over \partial\phi^{b}_{\bk} } ) ,
462: \nonumber \\
463: T^{a\dagger}_{-\bk} &=& {1 \over \sqrt{2}}P_{T}^{ab}
464: (\sqrt{\omega_{\bk}} \phi^a_{\bk}
465: + {1 \over \sqrt{\omega_{\bk}}}
466: {\partial \over \partial\phi^{b*}_{\bk} } ) ,
467: \ea
468: where $\omega_{\bk} = |{\bk}|$ is the Goldstone energy.
469: In terms of these operators, we can rewrite the effective
470: Hamiltonian in the following form
471: \ba
472: H&=& \sum_{\bk} \Omega_{\bk} h^{\dagger}_{\bk} h_{\bk}
473: +\sum_{\bk \neq 0} \omega_{\bk} T^{a\dagger}_{\bk} T^{a}_{\bk}
474: +H_{\rm int} +{ L^2 + \Delta_N \over 2 V (v+\sigma)^2} ,
475: \nonumber \\
476: H_{\rm int}&=& \sum_{\bx} {4 \lambda_0 v }
477: h(h^2 + \tilde{\phi^a_T}\tilde{\phi^a_T} )
478: + {\lambda_0 }(h^2 + \tilde{\phi^a_T}\tilde{\phi^a_T})^2 ,
479: \ea
480: where the fields $h(\bx)$ and $\tilde{\phi^a_T}$ are given by
481: \ba
482: h(\bx)&=& \sum_{\bk} {1 \over \sqrt{2V\Omega_{\bk}}}
483: ( h_{\bk}e^{i \bk \cdot \bx}
484: +h^{\dagger}_{\bk}e^{-i \bk \cdot \bx} ) ,
485: \nonumber \\
486: \tilde{\phi^a_T}(\bx)&=& \sum_{\bk \neq 0} {1 \over \sqrt{2V\omega_{\bk}}}
487: ( T^{a}_{\bk}e^{i \bk \cdot \bx}
488: +T^{a\dagger}_{\bk}e^{-i \bk \cdot \bx} ) ,
489: \ea
490: and the constant $\Delta_N = (N-3)(N-1)/4$. The operator
491: $L^2 \equiv L^{ab}_0 L^{ab}_0 /2$ is the $O(N)$ Casimir of
492: the zeromode variable.
493: The above creation and annihilation operators satisfy the following
494: commutation relations
495: \be
496: [T^a_{\bk}, T^{b\dagger}_{\bp}]=P^{ab}_T \delta_{\bk \bp},
497: \;\;\;
498: [L^a_{\bk}, L^{b\dagger}_{\bp}]=P^{ab}_L \delta_{\bk \bp} .
499: \ee
500: It is also very convenient to introduce the following decomposition
501: for the fields. For a given $O(N)$ unit vector $n^a$, we can find
502: additional $N-1$ unit vectors which, together with $n^a$, form a
503: complete set in the $O(N)$ space. We therefore define
504: \ba
505: \label{eq:ch4.compht}
506: n^a_0 &\equiv& n^a , \;\; n^a_{\alpha} n^a_{\beta}
507: =\delta_{\alpha \beta} , \;\; \alpha, \beta=0,1,...N-1 ,
508: \nonumber \\
509: T^a_{\bk}&=& n^a_i T_{i \bk}, \;\;
510: T_{i \bk}= n^a_i T^a_{\bk} ,
511: \nonumber \\
512: L^a_{\bk}&=& n^a_0 h_{\bk}, \;\;
513: h_{\bk}= n^a_0 L^a_{\bk} .
514: \ea
515: It is readily checked that these operators satisfy the standard
516: commutation relations
517: \be
518: [h_{\bk},h^{\dagger}_{\bp}]=\delta_{\bk \bp},
519: \;\;\;
520: [T_{i \bk},T^{\dagger}_{j \bp}]=\delta_{ij}\delta_{\bk \bp}.
521: \ee
522: Note that due to the leftover $O(N-1)$ symmetry, the determination of
523: the unit vectors $n^a_i$ is not unique. However, the physical quantities
524: will not depend on this ambiguity. Moreover we can calculate the
525: commutator of the operator $L^{ab}_0$ with the unit vectors,
526: \be
527: [L^{ab}_0, n^c_i] =(-i) n^{[a}_{\alpha} \delta^{b]c},
528: \;\;\;\;\; \alpha=0,1,...N-1 .
529: \ee
530: Now we can set up a basis in our Hilbert space from the eigenstate
531: of the free Hamiltonian. We will also choose the angular momentum
532: eigenstate of the $n^a$ variable, namely
533: \be
534: |n,\{ n^L_{\bk},n^T_{\bk} \},lm\rangle = |n\rangle \otimes
535: \prod_{\bk \neq 0} |n^L_{\bk} \rangle \otimes |n^T_{\bk} \rangle
536: \otimes |lm\rangle ,
537: \ee
538: where the state $|lm\rangle$ is the eigenstate of $L^2$ with eigenvalue
539: $l(l+N-1)$. This is just a symbolic notation
540: of the state. Strictly speaking, for
541: $O(N)$ model, we need more quantum numbers to specify
542: the state. Note that the oscillator part of the state actually
543: depends on the unit vector $n^a$ via the definition of the
544: longitudinal and transverse projection, although
545: the eigenvalue does not. Therefore, if we were to
546: act the operator $L^2$ on the states above, it would not only
547: act on the state $|lm\rangle$, but also act on the rest of the components
548: ( except $|n\rangle$, of course, since it is the radial zero
549: momentum mode ). As we will see below, the Born-Oppenheimer
550: Approximation will first neglect the effect of $L^2$ on the
551: fast modes and only consider the slow mode part of the
552: state, i.e. $|lm \rangle$. The next order correction has
553: to take this into account and the BOA is valid when
554: the correction is small.
555:
556: The global $O(N)$ generator $Q^{ab}$ now can be expressed in terms
557: of the creation and annihilation operators defined in
558: Equation~(\ref{eq:ch4.compht})
559: %\ba
560: %Q_{ab}&=&L^{ab}_0-i\sum_{\bk \neq 0}
561: % L^{[a\dagger}_{\bk}L^{b]}_{\bk}+
562: % T^{[a\dagger}_{\bk}T^{b]}_{\bk}
563: %\nonumber \\
564: %&+&g_{\bk}(T^{[a}_{\bk}L^{b]}_{-\bk}
565: % +L^{[a\dagger}_{\bk}T^{b\dagger]}_{-\bk})
566: %+f_{\bk}(L^{[b}_{-\bk}T^{a\dagger]}_{-\bk}
567: % +T^{[b}_{\bk}L^{a\dagger]}_{\bk})
568: %\ea
569: \be
570: Q^{ab}=L^{ab}_0-i\sum_{\bk \neq 0}
571: T^{\dagger}_{\alpha \bk}T_{\beta \bk} n^{[a}_{\alpha} n^{b]}_{\beta} ,
572: \ee
573: where the index $\alpha$ and $\beta$ run from $0$ to $(N-1)$.
574: The operators $T_{0 \bk}$ and the functions $f_{\bk}$, $g_{\bk}$ are
575: defined by the following:
576: \ba
577: T^{\dagger}_{0 \bk}=f_{\bk}h^{\dagger}_{\bk}-g_{\bk}h_{-\bk} ,
578: &&
579: T_{0 \bk}=f_{\bk}h_{\bk}-g_{\bk}h^{\dagger}_{-\bk} ,
580: \nonumber \\
581: f_{\bk}={1 \over 2}( \sqrt{{\Omega_{\bk} \over \omega_{\bk}}}
582: +\sqrt{{\omega_{\bk} \over \Omega_{\bk}}} ) ,
583: &&
584: g_{\bk}={1 \over 2}( \sqrt{{\Omega_{\bk} \over \omega_{\bk}}}
585: -\sqrt{{\omega_{\bk} \over \Omega_{\bk}}} ) .
586: \ea
587: We will need the result of $L^2$ acting on the oscillator
588: states. Let us consider the object
589: $L^2 |0_{\bk \neq 0}\rangle$.
590: This can be easily obtained by noticing that the state is
591: annihilated by the global $O(N)$ generators $Q^{ab}$.
592: Therefore we would induce that
593: \be
594: \label{eq:ch4.lab0}
595: L^{ab}_0|0_{\bk \neq 0} \rangle =\sum_{\bk \neq 0}
596: i g_{\bk} n^{[a}_{0} n^{b]}_{i}
597: h^{\dagger}_{-\bk}T^{\dagger}_{i \bk}
598: |0_{\bk \neq 0}\rangle .
599: \ee
600: We will also need the commutation relations between
601: $L^{ab}_0$ and the annihilation operators
602: \ba
603: \label{eq:ch5.commu}
604: [ L^{ab}_0 , L^c_{\bk} ]&=& (-i) ( n^{[a}\delta^{b]c}
605: h_{\bk}
606: +n^c ( f_{\bk} n^{[a} T^{b]}_{\bk}
607: +g_{\bk} n^{[a} T^{b\dagger]}_{-\bk}) ) ,
608: \nonumber \\
609: {[ L}^{ab}_0 , T^{c}_{\bk} ]&=& (-i) ( n^{[a} \delta^{b]c}
610: (f_{\bk}h_{\bk}-g_{\bk}h^{\dagger}_{-\bk})
611: + n^c n^{[a} T^{b]}_{\bk} ) ,
612: \ea
613: or equivalently, in terms of operators $T_{i \bk}$ and $h_{\bk}$,
614: we have
615: \ba
616: \label{eq:ch5.commucompo}
617: {[ L}^{ab}_0 ,h_{\bk} ] &=& (-i) n^{[a}_0 n^{b]}_i
618: (f_{\bk} T_{i \bk} + g_{\bk} T^{\dagger}_{i -\bk}) ,
619: \nonumber \\
620: {[ L}^{ab}_0 ,T_{i \bk} ] &=&(-i)n^{[a}_i n^{b]}_j T_{j\bk}+
621: (-i) n^{[a}_0 n^{b]}_i
622: (f_{\bk} h_{\bk} - g_{\bk} h^{\dagger}_{- \bk} ) .
623: \ea
624: The corresponding commutators with the creation operators
625: can be obtained from hermitian conjugation of the
626: above equations. We now have all the tools to study the
627: rotator energy spectrum.
628:
629: \subsection{$O(N)$ Model: The Ground States}
630:
631: The full ground state of the free Hamiltonian consists of two
632: parts. One is the ground state of oscillator states. The other part
633: is the rotator states.
634: \be
635: | G,lm \rangle =| 0_{osc} \rangle \otimes | lm \rangle .
636: \ee
637: Note that, before the rotator energy contributions are taken into
638: account, the degeneracy of the ground state is infinite, since any
639: rotator state $| lm \rangle$ will belong to the same energy.
640: It is very easy to check that all these states are eigenstates of
641: the momentum operator with eigenvalue of $0$. They can also be taken
642: as the eigenstates of the appropriate $O(N)$ operators. To see this,
643: notice that when the operator $Q^{ab}$ is applied to the states, it
644: generally has two contributions. One is from $Q^{ab}$ acting on the
645: oscillator state $| 0_{osc} \rangle$, which is zero in this case; the
646: other is from $Q^{ab}$ acting on the rotator states $|lm \rangle$, which
647: is equivalent to $ L^{ab}_0 |lm \rangle$. Therefore, we have
648: \be
649: Q^{ab} |G,lm \rangle = |0_{osc} \rangle \otimes L^{ab}_0 |lm \rangle .
650: \ee
651: By taking the states $|lm \rangle$ to be the eigenstates of the
652: zeromode Casimir, we also make the ground states to have the
653: appropriate $O(N)$ charges. Basically, the oscillator ground state
654: has $O(N)$ charge $0$, and all the $O(N)$ charges comes from the
655: rotator states.
656:
657: As mentioned above, the leading order ground states are infinitely
658: degenerate due to the rotator states. This degeneracy is lifted
659: once the first nonvanishing rotator correction is taken into account.
660: To do this, let us evaluate the matrix element of the rotator
661: energy operator $(1/2)L^{ab}_0L^{ab}_0\omega_r$ among the ground states
662: \ba
663: \!\!\!\!L^{ab}_0 |0_{osc}\rangle \otimes |lm \rangle &=&
664: (L^{ab}_0 |0_{osc}\rangle) \otimes |lm \rangle +
665: |0_{osc}\rangle \otimes (L^{ab}_0 |lm \rangle )
666: \nonumber \\
667: \!\!\!\!\!\! &=&\sum_{\bk \neq 0}
668: i g_{\bk} n^{[a}_{0} n^{b]}_{i}
669: h^{\dagger}_{-\bk}T^{\dagger]}_{i \bk}
670: |0_{osc}\rangle \otimes |lm \rangle
671: + |0_{osc} \rangle \otimes (L^{ab}_0 |lm \rangle ) ,
672: \ea
673: where we have used Equation~(\ref{eq:ch4.lab0}). Therefore we get
674: \be
675: \!\!\!\! \langle G,l'm'|1/2L^{ab}_0L^{ab}_0\omega_r| G,lm \rangle =
676: \left[l(l+N-2)\omega_r + (N-1)\omega_r \sum_{\bk \neq 0} g^2_{\bk}
677: \right] \delta_{ll'}\delta_{mm'} .
678: \ee
679: As expected, the degeneracy is lifted, and the ground states now
680: have a degeneracy of $l^2$, for a given $l$ value. The second term
681: in the above equation is an $l$ independent constant and can be absorbed
682: into the definition of the ground state energy. The first term is
683: $l$ dependent and is known as the rotator energy spectrum. This
684: formula was derived before by Leutwyler using
685: the rigid rotator approximation in the chiral Lagrangian formalism.
686: The significance of this energy spectrum in the Monte Carlo simulation
687: was also discussed \cite{leut4}.
688: The low energy excitations of the model exhibit a three hierarchy
689: of energy gaps. The largest energy gap is the mass gap of the Higgs
690: particle, whose energy is independent of the $3$-volume. The second
691: largest gap is the Goldstone particle, whose energy gap is typically
692: of order ${\cal O}(1/L)$, where $L$ is the size of the $3$ dimensional
693: cubic box. The smallest gap is the rotator energy differences between
694: different $l$ values, whose energy is of order
695: ${\cal O}(1/L^3)$. In a practical simulation, the size of the Higgs gap
696: and the Goldstone energy gap are of the same order, since the size of
697: the box is not large enough. However, the energy gap of the rotator
698: is usually much smaller compared with the Higgs and Goldstone. Therefore
699: we expect the Born-Oppenheimer picture should be a very good description
700: of the theory in the finite box.
701:
702: It is also possible to evaluate the second order correction to
703: the ground state energy. Let us first look at the following
704: quantity
705: \ba
706: L^{ab}_0L^{ab}_0|G,lm\rangle&=&
707: (L^{ab}_0L^{ab}_0|0_{osc}\rangle )\otimes |lm\rangle
708: \nonumber \\
709: &+&2(L^{ab}_0|0_{osc}\rangle )\otimes(L^{ab}_0 |lm\rangle)+
710: |0_{osc}\rangle \otimes(L^{ab}_0L^{ab}_0 |lm\rangle) .
711: \ea
712: The first term in the above equation will contribute at the second order
713: but it is a term independent of $l$. If we only focus on the
714: $l$-dependent terms, we can forget about this term. The last term
715: is diagonal, it will not contribute to the second order correction of
716: the ground state energy. Therefore, only the second term will give
717: us $l$-dependent contribution to the ground state energy.
718: The state left over is
719: \be
720: 2(L^{ab}_0|0_{osc}\rangle )\otimes(L^{ab}_0 |lm\rangle)=
721: 2\sum_{\bk \neq 0,a,b,i} i g_{\bk}n^{[a}_0n^{b]}_i
722: h^{\dagger}_{\bk}T^{\dagger}_{i,-\bk} |0_{osc} \rangle
723: \otimes(L^{ab}_0 |lm\rangle) .
724: \ee
725: Therefore, it will contribute a second order energy correction that
726: looks like
727: \be
728: E^{(2)}_{0l}=-\sum_{\bk \neq 0,i}
729: { \omega_r g^2_{\bk} \over \omega_{\bk}+\Omega_{\bk}}
730: \langle lm| L^{cd}_0n^{[c}_{0} n^{[d}_{i} n^{[a}_{0} n^{b]}_{0}L^{ab}_0
731: | lm \rangle .
732: \ee
733: The matrix element that appears in the above equation can be simplified
734: by noticing
735: \be
736: L^{cd}_0n^{[c}_{0} n^{[d}_{i} n^{[a}_{0} n^{b]}_{0}L^{ab}_0=
737: 4 L^{ab}_0 n^b_0 n^c_0 L^{ac}_0 .
738: \ee
739: We can pick our $O(N)$ axis such that the unit vector is in the
740: direction $(0,0,\cdots,1)$. Then the above operator simplifies to
741: the difference of two Casimirs: the Casimir of the $O(N)$ and the
742: Casimir of the unbroken $O(N-1)$
743: \be
744: E^{(2)}_{0l}=-\sum_{\bk \neq 0}
745: { 2\omega^2_r g^2_{\bk} \over \omega_{\bk}+\Omega_{\bk}}
746: \left( l(l+N-2) -\langle lm|L^2_{O(N-1)}|lm\rangle \right) .
747: \ee
748: This correction is usually very small for practical simulation
749: parameters. However, there is another second order correction of the
750: rotator Hamiltonian. Recall that we can expand the rotator Hamiltonian
751: into the form
752: \be
753: {L^2+\Delta_N \over 2 V (v+\sigma)^2} =
754: {L^2+\Delta_N \over 2 V v^2} (1 - {2\sigma \over v}
755: +{3\sigma^2 \over v^2}+\cdots ) .
756: \ee
757: Note that the operator
758: $\sigma=(h_0+h^{\dagger}_0)/\sqrt{2Vm_0}$. therefore we
759: have a systematic expansion in inverse powers of
760: the $3$-volume. If we introduce the rotator energy
761: $\omega_r \equiv 1/2Vv^2$, then we have a expansion
762: in terms of the small quantity $\omega_r/m$.
763: Therefore, there is another contribution from the
764: operator $L^2 \omega_r(\sigma/v)^2$ which is also
765: of the order of $\omega^2_r$. So we have
766: \ba
767: E^{(2)}_{0l}&=&[l(l+N-2)+\Delta_N]\omega_r{ 3 \omega_r \over m}
768: \nonumber \\
769: &-&\sum_{\bk \neq 0}
770: { 2\omega^2_r g^2_{\bk} \over \omega_{\bk}+\Omega_{\bk}}
771: \left( l(l+N-2) -\langle lm|L^2_{O(N-1)}|lm\rangle \right) .
772: \ea
773: The first term basically takes into account the nonrigid effects
774: of the rotator.
775:
776: \subsection{$O(N)$ Model: The Zero Momentum Higgs States}
777:
778: Let us consider the energy corrections to the state
779: $|n,0_{\bk \neq 0},lm \rangle$.
780: Since the zero momentum Higgs excitation is just the
781: radial excitation which commute with the angular
782: variables. Therefore the energy corrections to the state
783: are very much like the corrections for the ground states.
784: \ba
785: E^{(0)}_{nl}&=& (n+1/2) m ,
786: \nonumber \\
787: E^{(1)}_{nl}&=& \left[l(l+N-2)+\Delta_N \right]\omega_r ,
788: \nonumber \\
789: E^{(2)}_{nl}&=&(l(l+N-2)+\Delta_N)\omega_r(n+1/2){6\omega_r \over m}
790: \nonumber \\
791: &-&\sum_{\bk \neq 0}
792: { 2\omega^2_r g^2_{\bk} \over \omega_{\bk}+\Omega_{\bk}}
793: \left( l(l+N-2) -\langle lm|L^2_{O(N-1)}|lm\rangle \right) .
794: \ea
795: The second order correction is down by an extra factor of $\omega_r/m$
796: compared with the first order correction. However, due
797: to the large numerical factor in front, the effect of the first term is
798: still quite significant.
799: In a practical simulation, the correlation function
800: $\langle n^a(0)n^a(\tau)\rangle$
801: is measured and used as a way of extracting the
802: vacuum expectation value $v$. This correlation
803: function will pick up the
804: energy difference of $\Delta l=1$ states.
805: In most of the old simulations on $O(4)$, the
806: ratio $\omega_r/m$ is very small and
807: the rigid approximation of the rotator
808: energy gives very reliable results. In our
809: recent simulations on the higher derivative
810: theories, this ratio is of the order of
811: $10$ percent and the correction is noticeable.
812: We would find the wrong $v$ value if we did not
813: include this correction.
814:
815: \subsection{$O(N)$ Model: Two Pion States}
816:
817: We can perform the similar calculation for the two pion states. Let us
818: take the isospin zero channel states
819: $(N-1)^{-1/2} T^{\dagger}_{i,\bk}T^{\dagger}_{i,-\bk} |0\rangle
820: \otimes |lm \rangle$. We have
821: \ba
822: \!\!\!\!\!\!&&\langle l'm'| \otimes \langle 0|
823: T_{i,\bk} T_{i,-\bk} L^2 \omega_r
824: T^{\dagger}_{i,\bk} T^{\dagger}_{i,-\bk}
825: | 0 \rangle \otimes |l m \rangle/(N-1)
826: \nonumber \\
827: \!\!\!\!\!\!&=&\left( l(l+N-2) + (N-1) \sum_{\bp \neq 0} g^2_{\bp}
828: +2f^2_{\bk}+2g^2_{\bk} \right) \omega_r
829: \delta_{ll'} \delta_{mm'} ,
830: \ea
831: which implies that relative to the ground state the finite volume
832: correction is
833: \be
834: \Delta(2\omega_{\bk}) = 2(f^2_{\bk}+g^2_{\bk})\omega_r .
835: \ee
836: This correction is also very small when we extract the two pion
837: energy. For the simulation points where we extract the two
838: pion energies, this correction is below $1$ percent and is
839: therefore hidden in the statistical errors.
840:
841: \section{ Symmetry Breaking of the Higher Derivative $O(N)$ Model}
842:
843: The similar analysis can be done with the higher derivative
844: $O(N)$ model. Having discussed the ordinary $O(N)$ theory
845: we will be very brief and only point out the differences.
846: Many steps are also similar to the quantization of the
847: higher derivative theory which was discussed in detail in
848: Chapter~(\ref{ch:QUAN}).
849:
850: One starts with the general higher derivative Lagrangian which
851: has a global $O(N)$ symmetry
852: \be
853: {\cal L} = {1 \over 2} \phi^a (-\rho_1 \Box
854: -\rho_2 \Box^2-\rho_3 \Box^3 ) \phi^a
855: +{1 \over 2} \mu_0^2 \phi^a\phi^a
856: -{\lambda_0 } (\phi^a \phi^a)^2 ,
857: \ee
858: where $\Box=\partial^2_t -\nabla^2$
859: is the Minkowski space d'Alambert operator and
860: the coefficients are parametrized as
861: \be
862: \rho_1=1+{m^2_0 \over {\cal M}^2}+{m^2_0 \over \bar{{\cal M}}^2},
863: \;\;\;\;
864: \rho_2={1 \over {\cal M}^2}+{1 \over \bar{{\cal M}}^2}
865: +{m^2_0 \over {\cal M}^2\bar{{\cal M}}^2},
866: \;\;\;\;
867: \rho_3={1 \over {\cal M}^2\bar{{\cal M}}^2} .
868: \ee
869: After the usual steps of indefinite metric quantization, and
870: introduction of the Fourier modes, the
871: Hamiltonian has the form
872: ( see Equation~(\ref{eq:ch2.hambro}) to
873: Equation~(\ref{eq:ch2.hambrof}) for detail )
874: %\ba
875: %{\cal H}&=& i \pi^a_1\pi^a_2 + {1 \over 2\rho_3}\pi^a_3\pi^a_3
876: % +{1 \over 2} \pi^a_2 (\rho_1 -2\rho_2\nabla^2
877: % +3 \rho_3 \nabla^4) \pi^a_2
878: %\nonumber \\
879: %\!\!\!\! &+&{1 \over 2} \phi^a_1 (-\rho_1\nabla^2 -\rho_2\nabla^4
880: % -\rho_3 \nabla^6) \phi^a_1
881: % +{1 \over 2} \phi^a_3 (\rho_2
882: % -3\rho_3 \nabla^2) \phi^a_3
883: % +i \phi^a_2\phi^a_3
884: %\nonumber \\
885: % &-&{1 \over 2} \mu_0^2 \phi^a_1\phi^a_1
886: % +{\lambda_0 } (\phi^a_1 \phi^a_1)^2
887: %\ea
888: %The corresponding $O(N)$ generator is given by
889: %\be
890: %Q_{ab}=\sum_{i,\bx} \phi^{a}_i(\bx)\pi^{b}_i(\bx)
891: % -\phi^{b}_i(\bx)\pi^{a}_i(\bx)
892: %\ee
893: %which obviously commutes with the Hamiltonian.
894: %
895: %Next, the Fourier modes are introduced for each variable
896: %\ba
897: %\phi^a_i(\bx) &=& \bar{\phi}^a_i +
898: % {1 \over \sqrt{V}} \sum_{\bk > 0}
899: % \phi^a_{i,\bk} e^{i\bk \cdot \bx}
900: % +\phi^{a*}_{i,\bk} e^{-i\bk \cdot \bx}
901: %\nonumber \\
902: %\pi^a_i(\bx) &=& {(-i) \over V}{\partial \over \partial \bar{\phi}^a_i}
903: % +{(-i) \over \sqrt{V}} \sum_{\bk > 0}
904: % e^{-i\bk \cdot \bx}{\partial \over \partial \phi^a_{i,\bk}}
905: % +e^{+i\bk \cdot \bx}{\partial \over \partial \phi^{a*}_{i,\bk}}
906: %\ea
907: %and the Hamiltonian is brought into the following form
908: \ba
909: H &=& {1 \over V}(i \pi^a_{10}\pi^a_{20}
910: +{1 \over 2 \rho_3} \pi^a_{30}\pi^a_{30}
911: +{\rho_1 \over 2} \pi^a_{20}\pi^a_{20} )
912: +V ( {\rho_2 \over 2} \bar{\phi}^a_3\bar{\phi}^a_3
913: + i \bar{\phi}^a_2\bar{\phi}^a_3 )
914: \nonumber \\
915: && +
916: \sum_{\bk >0} i \pi^a_{1\bk}\pi^{a*}_{2\bk}
917: +i \pi^{a*}_{1\bk}\pi^{a}_{2\bk}
918: +{1 \over \rho_3} \pi^{a}_{3\bk}\pi^{a*}_{3\bk}
919: +(\rho_1+2\rho_2\bk^2+3\rho_3\bk^4)
920: \pi^{a}_{2\bk}\pi^{a*}_{2\bk}
921: \nonumber \\
922: &&
923: + (\rho_1\bk^2 + \rho_2\bk^4 + \rho_3\bk^6 )
924: \phi^{a}_{1\bk}\phi^{a*}_{1\bk}
925: + ( \rho_2\bk^4 +3\rho_3\bk^2 )
926: \phi^{a}_{3\bk}\phi^{a*}_{3\bk}
927: + i \phi^a_{2\bk}\phi^{a*}_{3\bk}
928: +i \phi^{a*}_{2\bk}\phi^{a}_{3\bk}
929: \nonumber \\
930: &&
931: -\sum_{\bx}{1 \over 2} \mu_0^2 \phi^a_1\phi^a_1
932: +\sum_{\bx}{\lambda_0 } (\phi^a_1 \phi^a_1)^2 .
933: \ea
934: Because we are now treating the system in a finite volume, we can
935: no longer neglect the motion of the zeromode. Instead, following
936: the idea of Born-Oppenheimer Approximation ( or Adiabatic
937: Approximation ),
938: we will single out the direction of the $\bar{\phi_1}^a$ variable
939: and make it the slow variable in our Born-Oppenheimer approximation.
940: We can then decompose
941: \be
942: \phi^a_1 =v n^a + h(\bx) n^a +\tilde{\phi}^a_{1T}(\bx) ,
943: \ee
944: and similarly for the $\phi_2$ and $\phi_3$ variables.
945: The Hamiltonian is then written as sum of three types of
946: terms
947: \ba
948: H &=& H_0+H_{\bk \neq 0}+H_{\rm int} ,
949: \nonumber \\
950: H_0&=&
951: {1 \over V}(i \pi^a_{10}\pi^a_{20}
952: +{1 \over 2 \rho_3} \pi^a_{30}\pi^a_{30}
953: +{\rho_1 \over 2} \pi^a_{20}\pi^a_{20} )
954: +V ( {\rho_2 \over 2} \bar{\phi}^a_3\bar{\phi}^a_3
955: + i \bar{\phi}^a_2\bar{\phi}^a_3+{m^2_0 \over 2}\sigma^2 ) ,
956: \nonumber \\
957: H_{\bk \neq 0}&=&
958: \sum_{\bk >0} i \pi^a_{1\bk}\pi^{a*}_{2\bk}
959: +i \pi^{a*}_{1\bk}\pi^{a}_{2\bk}
960: +{1 \over \rho_3} \pi^{a}_{3\bk}\pi^{a*}_{3\bk}
961: +(\rho_1+2\rho_2\bk^2+3\rho_3\bk^4)
962: \pi^{a}_{2\bk}\pi^{a*}_{2\bk}
963: \nonumber \\
964: &+&(\rho_1\bk^2+\rho_2\bk^4+\rho_3\bk^6+m^2_0)
965: \phi^{a}_{1\bk L}\phi^{a*}_{1\bk L}
966: + (\rho_1\bk^2+\rho_2\bk^4+\rho_3\bk^6)
967: \phi^{a}_{1\bk T}\phi^{a*}_{1\bk T}
968: \nonumber \\
969: &+&(\rho_2+3\rho_3\bk^2)
970: \phi^{a}_{3\bk}\phi^{a*}_{3\bk}
971: +i\phi^{a}_{2\bk}\phi^{a*}_{3\bk}
972: +i\phi^{a*}_{2\bk}\phi^{a}_{3\bk} ,
973: \nonumber \\
974: H_{\rm int}&=& \sum_{\bx}
975: {4 \lambda_0 v }h(h^2+
976: \tilde{\phi}^{a}_{1T}\tilde{\phi}^{a}_{1T})
977: +{\lambda_0 }(h^2+
978: \tilde{\phi}^{a}_{1T}\tilde{\phi}^{a}_{1T})^2 .
979: \ea
980: This Hamiltonian is identical to what we had in
981: Chapter~(\ref{ch:QUAN}), except for the $H_0$ piece
982: ( see Equation~(\ref{eq:ch2.hampiece}) ).
983: For example,
984: the $\bk \neq 0$ piece can be diagonalized in the same way
985: as in Chapter~(\ref{ch:QUAN}). The interaction piece is also
986: expressed as the creation and annihilation operators through
987: the field variables. The $H_0$ piece can be decomposed as follows
988: in the finite volume. For convenience we use the rescaled
989: variables given by
990: \ba
991: p^a_1&=&(\rho_1 V)^{-1/2}\pi^a_{10},
992: \;\;\;\;\;
993: p^a_2=\sqrt{{\rho_1 \over V}}\pi^a_{20},
994: \;\;\;\;\;
995: p^a_3=(\rho_3 V)^{-1/2}\pi^a_{30} ,
996: \nonumber \\
997: q^a_1&=&(\rho_1 V)^{1/2}\bar{\phi}^a_{1},
998: \;\;\;\;\;
999: q^a_2=\sqrt{{V \over \rho_1}}\bar{\phi}^a_{2},
1000: \;\;\;\;\;
1001: q^a_3=(\rho_3 V)^{1/2}\bar{\phi}^a_{3} ,
1002: \ea
1003: and use the radial variables for $q^a_1$
1004: \be
1005: q^a_1=\sqrt{\rho_1 V}(v+\sigma) n^a= \rho n^a .
1006: \ee
1007: The derivatives for the $q^a_1$ are now substituted by
1008: \be
1009: {\partial \over \partial q^a_1}=n^a
1010: {\partial \over \partial \rho}
1011: +(\delta^{a\alpha}-n^an^{\alpha})
1012: {\partial \over \partial n^{\alpha}} ,
1013: \ee
1014: where the index $a$ runs from $1$ to $N$ while the index
1015: $\alpha$ only runs from $1$ to $N-1$.
1016: The main difference lies in the derivative term with respect
1017: to the rotator variable $n^a$. In Chapter~(\ref{ch:QUAN}),
1018: this was neglected because we were in the infinite volume.
1019: This term practically serves as the kinetic energy of the
1020: zeromode variable $n^a$.
1021: One can establish
1022: the following identity
1023: \be
1024: i p^a_1p^a_2 = i p_{2L}p_{1 \rho} -i p^a_{2T}
1025: {n^b \over \rho} L^{ab}_0 ,
1026: \ee
1027: where $L^{ab}_0$ is the generator of the variable
1028: $q^a_1$ only, i.e.,
1029: \be
1030: L^{ab}_0= (-i)(q^a_1 {\partial \over \partial q^b_1}
1031: -q^b_1 {\partial \over \partial q^a_1} ) .
1032: \ee
1033: With these transformations, $H_0$ is further decomposed into
1034: three parts
1035: \ba
1036: H_0&=& H_{0L}+H_{0T}+H_{0LT}
1037: \nonumber \\
1038: H_{0L}&=& i p_{2L}p_{y} +{1 \over 2}p^2_{2L}
1039: +{1 \over 2}p^2_{3L}+{\rho_2 \over 2\rho_3}q^2_{3L}
1040: +i \sqrt{{\rho_1 \over \rho_3}}q_{2L}q_{3L}
1041: +{m^2_0 \over 2\rho_1} y^2 ,
1042: \nonumber \\
1043: H_{0T}&=& {1 \over 2}p^a_{2T}p^a_{2T}
1044: +{1 \over 2}p^a_{3T}p^a_{3T}
1045: +{\rho_2 \over 2\rho_3}q^a_{3T}q^a_{3T}
1046: +i \sqrt{{\rho_1 \over \rho_3}}q^a_{2T}q^a_{3T} ,
1047: \nonumber \\
1048: H_{0LT}&=&
1049: (-i)p^a_{2T}
1050: {n^b \over \rho} L^{ab}_0 .
1051: \ea
1052: The longitudinal part has the same form as the simple oscillator
1053: and can be diagonalized easily. The transverse part can also be
1054: diagonalized as shown in Chapter~(\ref{ch:QUAN})
1055: \be
1056: H_{0T}=\sum_{i \neq 0,a}a^{(+)a}_{i0T}a^{(-)a}_{i0T} \omega_{i0T}
1057: \ee
1058: where the summation of $a$ is from $1$ to $N$ and the
1059: energy gap is $\omega_{10T}=\cm_g$ and
1060: $\omega_{20T}=\cmb_g$.
1061: In terms of these operators we can write out the explicit
1062: form of $p^a_{2T}$
1063: \be
1064: p^a_{2T}=\sum_{i\neq 0} \sqrt{\omega_{i0T} \over 2} \epsilon_{i}
1065: (a^{(-)a}_{i0T}-a^{(+)a}_{i0T}) ,
1066: \ee
1067: where the polarization factor $\epsilon_i$ is given by
1068: $\epsilon_1=\epsilon^{*}_2=i/(1-e^{-4i\theta_g})^{1/2}$.
1069:
1070: To summarize, in the finite volume we would have the following
1071: Hamiltonian
1072: \ba
1073: H&=& H_0 + H_{\rm int} +{(-i)\over2}
1074: \left( p^a_{2T}{n^b \over \rho}L^{ab}_0
1075: +{n^b \over \rho}L^{ab}_0 p^a_{2T} \right) ,
1076: \nonumber \\
1077: H_0&=& \sum_{i,\bk,\lambda} a^{(+)a}_{i\bk\lambda}a^{(-)a}_{i\bk\lambda}
1078: \omega_{i\bk\lambda} ,
1079: \nonumber \\
1080: H_{\rm int}&=& \sum_{\bx}
1081: {4 \lambda_0 v }h(h^2+
1082: \tilde{\phi}^{a}_{1T}\tilde{\phi}^{a}_{1T})
1083: +{\lambda_0 }(h^2+
1084: \tilde{\phi}^{a}_{1T}\tilde{\phi}^{a}_{1T})^2 .
1085: \ea
1086: In this expression, the first two terms are just the Hamiltonian of
1087: the model in the broken phase in the infinite volume. The third
1088: term is a purely finite volume correction which describes the coupling
1089: between the zeromode and the rest of degrees of freedom. The
1090: index $\lambda$ takes the value $L$ and $T$ respectively. All the
1091: operators can be expressed in terms of the creation and annihilation
1092: operators as
1093: \ba
1094: h(\bx)&=& \sum_{i\bk} {c_{iL} \over \sqrt{2\omega_{i\bk L}V}}
1095: \left( n^a a^{(-)a}_{i\bk L} e^{i\bk\cdot\bx}
1096: +n^a a^{(+)a}_{i\bk L} e^{-i\bk\cdot\bx} \right) ,
1097: \nonumber \\
1098: \tilde{\phi}^a_{T}(\bx)&=& \sum_{i\bk \neq 0}
1099: {c_{iT} \over \sqrt{2\omega_{i\bk T}V}}
1100: \left( a^{(-)a}_{i\bk T} e^{i\bk\cdot\bx}
1101: + a^{(+)a}_{i\bk T} e^{-i\bk\cdot\bx} \right) ,
1102: \\
1103: \rho&=&\sqrt{\rho_1 V}(v+\sigma)=\sqrt{\rho_1 V}\left( v
1104: +\sum_{i}{c_{iL} \over \sqrt{2\omega_{i0L}V}}
1105: (a^{(-)}_{i0L}+a^{(+)}_{i0L}) \right) ,
1106: \nonumber
1107: \ea
1108: where the form factors $c_{i\lambda}$ are given by the following
1109: table
1110: \ba
1111: c_{0L}&=& \sqrt{ \cm^2 \cmb^2 \over (m^2_0-\cm^2)(m^2_0-\cmb^2)},
1112: \;\;\;c_{1L}=c^{*}_{2L}
1113: =\sqrt{ \cm^2 \cmb^2 \over (\cm^2-m^2_0)(\cm^2-\cmb^2)},
1114: \nonumber \\
1115: c_{0L}&=& 1,
1116: \;\;\;c_{1T}=c^{*}_{2T}=\sqrt{\cmb^2_g \over (\cm^2_g-\cmb^2_g)} .
1117: \ea
1118: The creation and annihilation operators enjoy the following
1119: commutation realtions
1120: \be
1121: [ a^{(-)a}_{i\bk\lambda} , a^{(+)b}_{j\bp\lambda^{'}} ]
1122: = \delta_{ij}\delta_{\bk \bp}\delta_{\lambda \lambda^{'}}P^{ab}_{\lambda}
1123: .
1124: \ee
1125: In the higher derivative model we have the similar relation
1126: for the $O(N)$ generators acting on the ground state
1127: \be
1128: L^{ab}_0 |0\rangle = i \sum_{i \bk \neq 0}
1129: g_{i \bk} a^{(+)[a}_{i\bk L}a^{(+)b]}_{i\bk T} |0 \rangle .
1130: \ee
1131: With these relations we can now calculate the rotator contribution
1132: to the energy of the state. Due to the selection rule for the
1133: operator $p^a_{2T}$, the first order correction vanishes. The lowest
1134: order correction comes in at the second order in the perturbation
1135: Hamiltonian. Using the representations of the operators in terms
1136: of the creation and annihilation operators, it is easy to show
1137: that the first correction is simply the rotator energy,
1138: \be
1139: E^{(1)}_{0l} = [ l(l+N-2) + \Delta_N ] \omega_r .
1140: \ee
1141: Therefore, just like in the conventional $O(N)$ model in the
1142: broken phase, the rotator energy spectrum is the most densely
1143: spaced excitation and dominates the invariant correlation
1144: functions.
1145:
1146:
1147:
1148:
1149: \begin{thebibliography}{9}
1150: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1151: %%%%%%%%%%% Standard Bibliography File %%%%%%%%%%%%%%%%%%
1152: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1153: %%%%%%
1154: %%%%%%%%%%%%%% The Standard Model in General
1155: %%%%%%
1156: %\bibitem{wein} S.~Weinberg, Phys. Rev. Lett. 19, (1967) 1264.
1157: %\bibitem{salam} A.~Salam, Elementary Particle Theory, Ed., N.~Svartholm,
1158: % Almquist and Wiksell, 1968.
1159: %\bibitem{glash} S.~L.~Glashow, Nucl. Phys. 22, (1961) 579.
1160: %\bibitem{quigg} B.~W.~Lee, C.~Quigg and H.~B.~Thacker, Phys. Rev. D16,
1161: %(1977) 1519.
1162: %%%%%%
1163: %%%%%%%%%%%%%% The Higher Derivative Theory in General
1164: %%%%%%
1165: %\bibitem{ostro} M.~Ostrogradski, Mem. Ac. St. Petersbourg 4 (1850) 385.
1166: %\bibitem{podo} B.~Podolski, Phys. Rev. 62 (1942) 68;
1167: % B.~Podolski and P.~Schwed, Rev. Mod. Phys. 20 (1948) 40.
1168: %\bibitem{pv} W.~Pauli and F.~Villars, Rev. Mod. Phys. 21 (1949) 434
1169: %\bibitem{pais} A.~Pais and G.~E.~Uhlenbeck, Phys. Rev. 79 (1950) 145
1170: %\bibitem{lee} T.~D.~Lee and G.~C.~Wick, Nucl. Phys. B 9 (1969) 209; Phys.
1171: %Rev. D 2 (1970) 1033.
1172: %\bibitem{polk} R.~E.~Cutkosky, P.~V.~Landshoff, D.~I.~Olive and
1173: %J.~C.~Polkinghorne, Nucl. Phys. B12 (1969) 281.
1174: %\bibitem{ghost} K.~Jansen, J.~Kuti, C.~Liu Phys. Lett. B 309 (1993) 119.
1175: %\bibitem{gross} D.~G.~Boulware and D.~J.~Gross, Nucl. Phys. B233 (1983)
1176: %1.
1177: %\bibitem{pauli} W.~Pauli, Rev. Mod. Phys. 15 (1943) 175.
1178: %\bibitem{simon} J.~Z.~Simon, Phys. Rev. D41 (1990) 3720.
1179: %\bibitem{scat} J.~Kuti and C.~Liu, to be published.
1180: %%%%%%
1181: %%%%%%%%%%%% Higher Derivative Gauge Theory and Gravity
1182: %%%%%%
1183: %\bibitem{slav} A.~A.~Slavnov, Nucl.~Phys. B31 (1971) 301.
1184: %\bibitem{hawk} S.~W.~Hawking, Quantum field theory and quantum
1185: %statistics, eds. I.~A.~Batalin et al. (1987) p. 129.
1186: %\bibitem{stelle} K.~S.~Stelle, Phys. Rev. D16 (1977) 953.
1187: %\bibitem{tomb} E.~Tomboulis, Phys. Lett. B97 (1980) 77.
1188: %%%%%%
1189: %%%%%%%%%%%% Algorithms Development
1190: %%%%%%
1191: %\bibitem{kenn} A.~D.~Kennedy and B.~Pendleton, Nucl. Phys. B (Proc.
1192: %Suppl.) 20 (1991) 118.
1193: %\bibitem{parisi} G.~Parisi, Progress in gauge field theory, ed. G.~'t~
1194: %Hooft et al. (Plenum, New York, 1984) 531.
1195: %\bibitem{batrouni} G.~Batrouni, G.~Katz, A.~Kronfeld, G.~P.~Lepage, P.~
1196: %Rossi, B.~Svetitsky and K.~Wilson, Phys. Rev. D32 (1985) 2736.
1197: %\bibitem{kogut} E.~Dagotto and J.~B.~Kogut, Nucl. Phys. B 290 (1987)
1198: %451.
1199: %%%%%%
1200: %%%%%%%%%%%% Triviality in General
1201: %%%%%%
1202: %\bibitem{landau} L.~D.~Landau, A.~A.~Abrikosov and I.~M.~Khalatnikov,
1203: %Doklady Akad. Nauk. USSR 95 (1954) 1177.
1204: %\bibitem{maiani} L.~Maiani, G.~Parisi and R.~Petronzio, Nucl. Phys. B136
1205: %(1978) 115.
1206: %\bibitem{dash} R.~Dashen and H.~Neuberger, Phys. Rev. Lett. 50 (1983)
1207: %1897.
1208: %%%%%%
1209: %%%%%%%%%%%% Higgs Mass Bound
1210: %%%%%%
1211: \bibitem{hhiggs4} K.~Jansen, J.~Kuti, C.~Liu Phys. Lett. B309 (1993) 127.
1212: \bibitem{dallas4} C.~Liu, K.~Jansen and J.~Kuti, Nucl. Phys. B 34 (Proc.
1213: Suppl.), (1994) 635.
1214: %\bibitem{kuti} J.~Kuti, L.~Lin, Y.~Shen, Nucl. Phys. (Proc. Suppl.) B 4
1215: %(1988) 397; Phys. Rev. Lett. 61 (1988) 678.
1216: %\bibitem{lusc} M.~L\"uscher and P.~Weisz, Phys. Lett. B212 (1988) 472.
1217: %\bibitem{hase} A.~Hasenfratz et al., Nucl. Phys. B317 (1989) 81.
1218: \bibitem{syman4} K.~Symanzik, Nucl. Phys. B226 (1983) 187.
1219: %\bibitem{zimm} M.~G\"ockeler, H.~Kastrup, T.~Neuhaus and F.~Zimmermann,
1220: % Nucl. Phys. (Proc. Suppl.) B26 (1992) 516.
1221: \bibitem{neub4}U.~M.~Heller, H.~Neuberger and P.~Vranas, Nucl. Phys. B405
1222: (1993) 557.
1223: %\bibitem{einh} M.~B.~Einhorn, Nucl. Phys. B246 (1984) 75.
1224: %M.~B.~Einhorn and D.~N.~Williams, Phys. Lett. B211 (1988) 4570.
1225: %\bibitem{lang} C.~B.~Lang, Phys. Lett. B229 (1989) 97; Nucl. Phys.
1226: %B (Proc. Suppl.) 17 (1990) 665.
1227: %%%%%%
1228: %%%%%%%%%%%% BOA Approximation
1229: %%%%%%
1230: \bibitem{leut4} A.~Hasenfratz et al., Nucl. Phys. B356 (1991) 332.
1231: %%%%%%
1232: %%%%%%%%%%%% Luescher's formula
1233: %%%%%%
1234: %\bibitem{luscf} M.~L\"uscher, Nucl. Phys. B354 (1991) 531; Nucl.
1235: %Phys. B364 (1991) 237.
1236: %\bibitem{luscwolf} M.~L\"uscher, U.~Wolff, Nucl. Phys. B339 (1990)
1237: % 222.
1238: %\bibitem{zimmph1} F.~Zimmermann, J.~Westphalen, M.~G\"ockeler and
1239: %H.~A.~Kastrup, Nucl. Phys. B (Proc. Suppl.) 30 (1993) 879.
1240: %\bibitem{zimmph2} F.~Zimmermann, J.~Westphalen, M.~G\"ockeler and
1241: %H.~A.~Kastrup, Nucl. Phys. B (Proc. Suppl.) 34 (1994) 566.
1242: %\bibitem{dewitt} B.~S.~DeWitt, Phys. Rev. 103 (1956) 1565.
1243: %
1244: \end{thebibliography}
1245:
1246:
1247: \vfill\eject
1248: