1: % draft 10-19-01
2: % revised 11-26-01
3: % revised 02-18-03
4: % revised 11-08-04
5:
6: % preprint style
7: %\documentstyle[preprint,aps,psfig]{revtex}
8: \documentclass[preprint,aps]{revtex4}
9: % galley style
10: %\documentstyle[aps,epsf,twocolumn]{revtex}
11: % preprint style, small
12: %\documentstyle[draft,aps,epsf]{revtex}
13: %\documentstyle[12pt]{article}
14: \usepackage{graphicx}
15:
16: %useful macros
17: \newcommand {\bea}{\begin{eqnarray}}
18: \newcommand {\eea}{\end{eqnarray}}
19: \newcommand {\be}{\begin{equation}}
20: \newcommand {\ee}{\end{equation}}
21: \newcommand {\qslash}{q\!\!\!/}
22: \newcommand {\muslash}{\mu\!\!\!/}
23: \newcommand {\Dslash}{D\!\!\!/}
24:
25: \begin{document}
26:
27:
28: %\draft
29: %\preprint{SUNY-NTG-01-41}
30:
31: \title{Instantons and Monte Carlo Methods in
32: Quantum Mechanics}
33:
34: \author{T.~Sch\"afer$^{1,2}$}
35:
36: \address{
37: $^1$Department of Physics, North Carolina State University,
38: Raleigh, NC 27695\\
39: $^2$Riken-BNL Research Center, Brookhaven National
40: Laboratory, Upton, NY 11973}
41:
42:
43: \begin{abstract}
44: In these lectures we describe the use of Monte Carlo
45: simulations in understanding the role of tunneling
46: events, instantons, in a quantum mechanical toy model.
47: We study, in particular, a variety of methods that
48: have been used in the QCD context, such as Monte Carlo
49: simulations of the partition function, cooling and
50: heating, the random and interacting instanton
51: liquid model, and numerical simulations of non-Gaussian
52: corrections to the semi-classical approximation.
53:
54: \end{abstract}
55: \maketitle
56:
57: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
58: \section{Introduction}
59: \label{sec_intro}
60: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
61:
62: We consider a non-relativistic particle moving in a potential
63: $V(x)$. The Hamiltonian of the system is given by
64: \be
65: \label{H}
66: H = \frac{p^2}{2m} + \lambda \left(x^2-\eta^2\right)^2 .
67: \ee
68: We can rescale $x$ and $t$ such that $2m=\lambda=1$. We will
69: also use $\hbar=1$. This means that the system is characterized
70: by just one dimensionless parameter, $\eta$. The potential $V(x)$
71: with $\eta=1.4$ is shown in Fig.~\ref{fig_spec}a. The physics of
72: this system is easy to understand. Classically, there are two
73: degenerate minima at $x=\pm \eta$. Quantum mechanically, the two
74: states can mix. If the potential barrier is very high, $\eta\to
75: \infty$, the wave functions of the ground state and the first
76: excited state are approximately given by
77: \be
78: \label{psipm}
79: \psi_{0,1}(x)=\frac{1}{\sqrt{2}}(\psi_-(x)\pm \psi_+(x)),
80: \ee
81: where $\psi_\pm(x)$ are the ground state wave functions in the
82: left and right minimum of the potential. The energy splitting
83: between the ground state and the first excited state is
84: exponentially small. The WKB approximation gives
85: \be
86: \Delta E = E_1-E_0 = \sqrt{\frac{6S_0}{\pi}}\omega
87: \exp(-S_0),
88: \ee
89: where $\omega=4\eta$ and $S_0=m^2\omega^3/(12\lambda)=4\eta^3/3$.
90:
91: Applications of the WKB method and of instantons to the
92: double well potential are discussed in many reviews and
93: text books \cite{Polyakov:1976fu,Vainshtein:1981wh,Coleman:1978ae,Shuryak:1988ck,Kleinert:1995,Zinn-Justin:1993wc,Schafer:1996wv,Forkel:2000sq}.
94: It is not our intention to present another review on the
95: subject in these lecture notes. Instead, we will use
96: the double well potential to illustrate a number of
97: numerical methods that have proven useful in the context
98: of QCD and other gauge theories.
99:
100: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
101: \section{Exact Diagonalization}
102: \label{sec_diag}
103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
104:
105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
106: \begin{figure}
107: \begin{minipage}{8cm}
108: \includegraphics[width=7.5cm,clip=true]{pot.eps}
109: \end{minipage}
110: \begin{minipage}{8cm}
111: \includegraphics[width=7.5cm,clip=true]{spectrum.eps}
112: \end{minipage}
113: \caption{\label{fig_spec}
114: a) Double well potential $V(x)=(x^2-\eta^2)^2$ for $\eta=1.4$.
115: We have indicated the position of the ground state and the
116: first three excited state. b) Spectrum of the double well potential
117: as a function of the parameter $\eta$. In this figure we show
118: the position of the first six states. We clearly observe that
119: positive and negative parity states become degenerate as $\eta\to
120: \infty$.}
121: \end{figure}
122: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
123:
124: The quantum mechanical problem defined by the Hamiltonian
125: equ.~(\ref{H}) can be solved by determining the eigenvalues
126: and eigenvectors of $H$. This can be achieved by choosing a
127: basis and numerically diagonalizing the Hamilton operator in
128: that basis. We have chosen a simple harmonic oscillator basis
129: defined by the eigenstates of
130: \be
131: \label{H_0}
132: H_0 = \frac{p^2}{2m}+\frac{1}{2}m\omega_0^2 x^2.
133: \ee
134: The value of $\omega_0$ is arbitrary, but the truncation
135: error of the eigenvalues computed in a finite basis will
136: depend on the choice of $\omega_0$. In practice, however,
137: this dependence is quite weak. The eigenstates of $H_0$
138: satisfy $H_0|n\rangle = |n\rangle \omega_0(n+1/2)$. The
139: Hamiltonian of the anharmonic oscillator has a very
140: simple structure in this basis. The only non-zero matrix
141: elements are
142: \bea
143: \langle n|H|n\rangle &=&
144: 3A c^4 \left[(n+1)^2+n^2\right]
145: + B c^2 (2n+1)
146: + \omega_0 (n+1/2) + c ,\\
147: \langle n|H|n+2\rangle &=&
148: A c^4 (4n+6)\sqrt{(n+1)(n+2)}
149: +B c^2 \sqrt{(n+1)(n+2)}, \\
150: \langle n|H|n+4\rangle &=&
151: c^4 \sqrt{(n+1)(n+2)(n+3)(n+4)},
152: \eea
153: as well as the corresponding hermitian conjugates. We have
154: also defined $A=1$, $B=-2\eta^2-\omega_0^2/4$, $C=\eta^4$
155: and $c=1/\sqrt{\omega_0}$. Note that both $H$ and $H_0$
156: conserve parity. We can decompose the matrix $H_{nm}=\langle
157: n|H|m\rangle$ into even and odd components, $H=H_{even}+H_{odd}$,
158: such that the eigenvectors of $H_{even}$ and $H_{odd}$ have
159: positive and negative parity, respectively.
160:
161: With the choice $\omega_0=4\eta$ even modest basis sizes such
162: as $N=40$ are sufficient in order to determine the first few
163: eigenvectors very accurately. In Fig.~\ref{fig_spec}b we show the
164: first six eigenvalues as a function of the parameter $\eta$.
165: We clearly observe that as $\eta$ increases pairs of eigenvalues
166: corresponding to even and odd eigenfunctions become almost
167: degenerate. In this limit the eigenfunctions are of the form
168: given in equ.~(\ref{psipm}).
169:
170:
171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
172: \section{Quantum Mechanics on a Euclidean Lattice }
173: \label{sec_latt}
174: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
175:
176: An alternative to the Hamiltonian formulation of the
177: problem is the Feynman path integral \cite{Feynman}.
178: The path integral for the anharmonic oscillator is given by
179: \be
180: \label{pathint}
181: \langle x_1| e^{-iHt_1}|x_0\rangle =
182: \int_{x(0)=x_0}^{x(t_1)=x_1} {\cal D}x\, e^{iS},
183: \hspace{1cm}
184: S=\int_0^{t_1} dt\, \left(
185: \frac{1}{4}\dot{x}^4-(x^2-\eta^2)^2 \right).
186: \ee
187: In the following we shall consider the euclidean
188: partition function
189: \be
190: \label{z}
191: Z(T) = \int {\cal D}x\, e^{-S_E}, \hspace{1cm}
192: S_E=\int_0^{\beta} d\tau\, \left(
193: \frac{1}{4}\dot{x}^4+(x^2-\eta^2)^2 \right),
194: \ee
195: where $\beta=1/T$ is the inverse temperature. The partition
196: function can be expressed in terms of the eigenvalues of the
197: Hamiltonian, $Z(T)=\sum_n\exp(-E_n/T)$. In the following we
198: shall study numerical simulations using a discretized euclidean
199: action. For this purpose we discretize the euclidean time coordinate
200: $\tau_i=ia,\,i=1,\ldots n$. The discretized action is given by
201: \be
202: \label{S_disc}
203: S = \sum_{i=1}^{n}\left\{ \frac{1}{4a} (x_i-x_{i-1})^2
204: + a(x_i^2-\eta^2)^2 \right\},
205: \ee
206: where $x_i=x(\tau_i)$. We shall consider periodic boundary
207: conditions $x_0=x_n$. The discretized euclidean path integral
208: is formally equivalent to the partition function of a statistical
209: system of ``spins'' $x_i$ arranged on a one-dimensional lattice.
210: This statistical system can be studied using standard Monte-Carlo
211: sampling methods. In the following we will simply use the Metropolis
212: algorithm \cite{Creutz:1980gp,Shuryak:1984xr,Shuryak:1987tr}.
213: The Metropolis method generates an ensemble of configurations
214: $\{x_i\}^{(k)}$ where $i=1,\ldots, n$ labels the lattice points
215: and $k=1,\ldots,N_{conf}$ labels the configurations. Quantum
216: mechanical averages are computed by averaging observables
217: over many configurations,
218: \be
219: \langle {\cal O} \rangle = \lim_{N_{conf}\to\infty}
220: \frac{1}{N_{conf}}\sum_{k=1}^{N_{conf}}
221: {\cal O}^{(k)}
222: \ee
223: where ${\cal O}^{(k)}$ is the value of the classical observable
224: ${\cal O}$ in the configuration $\{x_i\}^{(k)}$. The configurations
225: are generated using Metropolis updates $\{x_i\}^{(k)}\to \{x_i\}^{(k+1)}$.
226: The update consists of a sweep through the lattice during which a trial
227: update $x_i^{(k+1)}= x_i^{(k)} +\delta x$ is performed for every lattice
228: site. Here, $\delta x$ is a random number. The trial update is accepted
229: with probability
230: \be
231: P\left(x_i^{(k)}\to x_i^{(k+1)}\right)=
232: \min\left\{\exp(-\Delta S),1\right\},
233: \ee
234: where $\Delta S$ is the change in the action equ.~(\ref{S_disc}).
235: This ensures that the configurations $\{x_i\}^{(k)}$ are distributed
236: according the ``Boltzmann'' distribution $\exp(-S)$. The distribution
237: of $\delta x$ is arbitrary as long as the trial update is micro-reversible,
238: i.~e.~is equally likely to change $x_i^{(k)}$ to $x_i^{(k+1)}$ and back.
239: The initial configuration is also arbitrary. In order to study
240: equilibration it is often useful to compare an ordered (cold)
241: start with $\{x_i\}^{(0)}=\{\eta\}$ to a disordered (hot) start $
242: \{x_i\}^{(0)}=\{r_i\}$, where $r_i$ is a random variable.
243:
244: The advantage of the Metropolis algorithm is its simplicity
245: and robustness. The only parameter to adjust is the distribution
246: of $\delta x$. We typically take $\delta x$ to be a Gaussian random
247: number with the width of the distribution adjusted such that the
248: average acceptance rate for the trial updates is around $50\%$.
249: Fluctuations of ${\cal O}$ provide an estimate in the error
250: of $\langle {\cal O}\rangle$. We have
251: \be
252: \Delta \langle {\cal O} \rangle =
253: \sqrt{\frac{\langle {\cal O}^2\rangle -\langle{\cal O}\rangle^2}
254: {N_{conf}}}.
255: \ee
256: This requires some care, because the error estimate is based on
257: the assumption that the configurations are statistically independent.
258: In practice this can be monitored by computing the auto-correlation
259: ``time'' in successive measurements ${\cal O}(\{x_i\}^{(k)})$.
260: The auto-correlation time of different observables can be
261: very different. For example, successive measurements of the
262: total energy decorrelate very quickly, but measurements
263: of the topological charge have a much longer correlation
264: time.
265:
266: The energy eigenvalues and wave functions of the quantum
267: mechanical problem can be obtained from the euclidean correlation
268: functions
269: \be
270: \label{qm_cor}
271: \Pi(\tau)=\langle O(0)O(\tau)\rangle.
272: \ee
273: Here, $O(\tau)$ is an operator that can be constructed from the
274: variables $x(\tau)$, e.g. some integer power $O(\tau)=
275: x(\tau)^n$. The euclidean correlation functions are related
276: to the quantum mechanical states via spectral representations.
277: The spectral representation is obtained by inserting a
278: complete set of states into the expectation value
279: equ.~(\ref{qm_cor}). We find
280: \be
281: \label{spec}
282: \Pi(\tau)= \sum_n |\langle 0|O(0)|n\rangle|^2
283: \exp(-(E_n-E_0)\tau),
284: \ee
285: where $E_n$ is the energy of the state $|n\rangle$ and
286: $|0\rangle$ is the ground state of the system. We can
287: write this as
288: \be
289: \Pi(\tau)= \int dE\, \rho(E) \exp(-(E-E_0)\tau).
290: \ee
291: In the case we are studying here there are only bound
292: states and the spectral function $\rho(E)$ is a sum of
293: delta-functions. Equ.~(\ref{spec}) shows that the euclidean
294: correlation function is easy to construct once the energy
295: eigenvalues and eigenfunctions are known. This fact was
296: used in order to calculate the solid lines shown in
297: Figs.~(\ref{fig_cor}). The inverse problem is well defined
298: in principle, but numerically much harder. In the
299: following we will concentrate on extracting just the
300: first few energy levels. A technique that can be used
301: on order determine the spectral function from euclidean
302: correlation functions is the maximum entropy image
303: reconstruction method, see \cite{Jarrel:1996,Asakawa:2000tr}.
304:
305: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
306: \begin{figure}
307: \begin{center}
308: \leavevmode
309: \includegraphics[width=9cm]{config.eps}
310: \end{center}
311: \caption{\label{fig_path}
312: Typical euclidean path obtained in a Monte Carlo simulation
313: of the discretized euclidean action of the double well
314: potential for $\eta=1.4$. The lattice spacing in the
315: euclidean time direction is $a=0.05$ and the total number
316: of lattice points is $N_\tau=800$. The green curve shows
317: the corresponding smooth path obtained by running 100
318: cooling sweeps on the original path. }
319: \end{figure}
320: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
321:
322: The Monte Carlo method is very useful in calculating
323: expectation values in quantum or statistical mechanics.
324: However, the Monte Carlo method does not directly give
325: the partition function or the free energy. In principle
326: one can reconstruct the free energy from the energy
327: eigenvalues but this is not very practical since, as
328: we just mentioned, it is hard to compute the full spectrum.
329: A very effective method for computing the free energy is
330: the adiabatic switching technique. The idea is to start
331: from a reference system for which the free energy is
332: known and calculate the free energy difference to the
333: real system using Monte Carlo methods.
334:
335: For this purpose we write the action as $S_\alpha=S_0+\alpha
336: \Delta S$ where $S$ is the full action, $S_0$ is the action
337: of the reference system, $\Delta S$ is defined by $\Delta S
338: =S-S_0$, and $\alpha$ is a coupling constant. The action
339: $S_\alpha$ interpolates between the real and the reference
340: system. Integrating the relation $\partial \log Z(\alpha)/
341: (\partial\alpha)=-\langle \Delta S \rangle_\alpha$ we find
342: \be
343: \label{adiab}
344: \log(Z(\alpha=1))=\log(Z(\alpha=0))
345: - \int_0^1 d\alpha'\, \langle \Delta S\rangle_{\alpha'} \;\; ,
346: \ee
347: where $\langle .\rangle_\alpha$ is an expectation value
348: calculated using the action $S_\alpha$. In the present case
349: it is natural to use the harmonic oscillator as a reference
350: system. In that case
351: \be
352: Z(\alpha=0) = \sum_n \exp(-\beta E_n^0)
353: = \frac{\exp(-\beta\omega_0/2)}{1-\exp(-\beta\omega_0)},
354: \ee
355: where $\omega_0$ is the oscillator constant. Note that the
356: free energy of the anharmonic oscillator should be independent
357: of $\omega_0$. The integral over the coupling constant $\alpha$
358: can easily be calculated in Monte Carlo simulations by slowly
359: changing $\alpha$ from 0 to 1 during the simulation. In
360: order to estimate systematic errors due to incomplete equilibration
361: it is useful to repeat the calculation with $\alpha$ changing
362: from 1 to 0 and study possible hysteresis effects.
363:
364: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
365: \section{Numerical Results}
366: \label{sec_num}
367: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
368:
369: Numerical results from Monte Carlo simulations of the euclidean
370: path integral are shown in Figs.~\ref{fig_path}-\ref{fig_cor}. The
371: numerical data were obtained using the program {\tt qm.for} which
372: is described in more detail in the appendix. A typical path that
373: appears in the Monte Carlo simulation is shown in Fig.~\ref{fig_path}.
374: The figure clearly shows that there are two characteristic time scales
375: in the problem. On short time scales the motion is controlled by the
376: oscillation time $\tau_{osc}\sim \omega^{-1}\sim (4\eta)^{-1}$.
377: For large $\tau$ the system is governed by the tunneling time
378: $\tau_{tun} \sim \exp(-4\eta^3/3)$. In order to perform reliable
379: simulations we have to make sure that the lattice spacing $a$ is
380: small compared to $\tau_{osc}$ and that the total length of
381: the lattice $Na$ is much larger than the tunneling time
382: \be
383: a\ll \tau_{osc}, \hspace{0.5cm} \tau_{tun}\ll Na .
384: \ee
385: A typical choice of parameters for the case $\eta=1.4$ is a number
386: of lattice points $N=800$, a lattice spacing $a=0.05$ and a number
387: of Metropolis sweeps $N_{conf}=10^5$.
388:
389: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
390: \begin{figure}
391: \begin{center}
392: \leavevmode
393: \includegraphics[width=9cm]{psi2.eps}
394: \end{center}
395: \caption{\label{fig_psi}
396: Probability distribution $|\psi(x)|^2$ in the double well
397: potential for $\eta=1.4$. The solid line shows the ``exact''
398: numerical result obtained by diagonalizing the Hamiltonian
399: in an oscillator basis whereas the histogram shows the
400: distribution of $x$ for an ensemble of euclidean paths. }
401: \end{figure}
402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
403:
404: Fig.~\ref{fig_psi} shows the distribution of $x_i$ obtained
405: in the Monte Carlo simulation compared to the square of the
406: ground state wave function computed by the diagonalization
407: method discussed in section \ref{sec_diag}. As $\eta$ is
408: increased and the potential barrier becomes larger the tunneling
409: time increases exponentially and the number of configurations
410: needed to reproduce the correct wave function becomes very
411: large.
412:
413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
414: \begin{figure}
415: \begin{center}
416: \leavevmode
417: \includegraphics[width=8cm,clip=true]{qmcor.eps}
418: \includegraphics[width=8cm,clip=true]{dqmcor.eps}
419: \end{center}
420: \caption{\label{fig_cor}
421: Fig.~a shows the correlation functions $\langle {\cal O}(0)
422: {\cal O}(\tau)\rangle$ in the double well potential for
423: $\eta=1.4$ and ${\cal O}=x,x^2,x^3$. The solid lines are ``exact''
424: numerical results obtained by diagonalizing the Hamiltonian in an
425: oscillator basis whereas the data point were obtained from Monte
426: Carlo simulations with $a=0.05$ and $N_\tau=800$. Fig.~b shows
427: the logarithmic derivative of the correlators in Fig.~a. In the
428: case of the $\langle x^2(0)x^2(\tau)\rangle$ we subtracted the
429: constant contribution.}
430: \end{figure}
431: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
432:
433: Fig.~\ref{fig_cor} shows the correlation functions of the
434: operators $x,x^2$ and $x^3$. The solid lines show the result
435: obtained using the spectral representation equ.~(\ref{spec})
436: together with the eigenvalues and eigenfunctions determined
437: by numerical diagonalization of the Hamiltonian. The data
438: points show the results from the Monte Carlo simulation. There
439: is a small systematic disagreement for small $\tau$ which is
440: related to discretization errors but the overall agreement
441: is excellent. Energy levels and matrix elements can be obtained
442: from the logarithmic derivative of the correlation function,
443: \be
444: C(\tau)= -\frac{d\log\Pi(\tau)}{d\tau} =
445: \frac{\sum_n (E_n-E_0) |\langle 0|O(0)|n\rangle|^2
446: \exp(-(E_n-E_0)\tau)}
447: {\sum_n |\langle 0|O(0)|n\rangle|^2 \exp(-(E_n-E_0)\tau)}.
448: \ee
449: In the limit $\tau\to\infty$ the function $C(\tau)$ converges to
450: the energy splitting between the ground state and the first excited
451: state that has a non-vanishing transition amplitude $\langle 0|O(0)
452: |n\rangle$. Because of parity invariance, $O=x^n$ connects the
453: groundstate to parity even/odd levels for $n$ even/odd. Since the
454: first excited state is parity odd we have
455: \be
456: \lim_{\tau\to\infty} \frac{d}{d\tau} \log \langle x(\tau)x(0)\rangle
457: = \lim_{\tau\to\infty}\frac{d}{d\tau} \log \langle x^3(\tau)x^3(0)\rangle
458: = E_1-E_0.
459: \ee
460: For even powers of $x$ the situation is more complicated
461: because the correlator has a constant term $|\langle 0|x^{2n}
462: |0\rangle|^2$. After subtracting the constant part, the logarithmic
463: derivative of the correlation function of even powers of $x$ tends
464: to $(E_2-E_0)$. Numerical results are shown in Fig.~\ref{fig_cor}b.
465: We observe that the logarithmic derivative of $\langle x(\tau) x(0)
466: \rangle$ converges very rapidly to $\Delta E_1=(E_1-E_0)$. The
467: numerical results for $\Delta E_2=(E_2-E_0)$ have large
468: uncertainties. These uncertainties are related to the fact that
469: the correlator $\langle x^2(\tau)x(0)\rangle$ is dominated by
470: the subtraction constant $\langle x^2\rangle^2$. The logarithmic
471: derivative of the $\langle x^3(\tau)x^3(0)\rangle$ correlator
472: also tends to $\Delta E_1$, but receives larger contributions
473: from excited states. This feature can be used in order to extract
474: the energies of higher states. The idea is very simple. From the
475: matrix elements $c_1=\langle 0|x|1\rangle$ and $d_1=\langle 0|
476: x^3|1\rangle$ we can determine a new operator ${\cal O}=x/c_1-
477: x^3/d_1$ that does not couple to the first excited state. This
478: operator predominantly couples to the third excited state.
479: Repeating this procedure we can determine the energies of
480: higher excited states. The problem is that correlation
481: functions of higher powers of $x$ are more and more noisy.
482: As a result, finding the energies of highly excited states
483: is very hard, even in the simple problem considered here.
484:
485: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
486: \begin{figure}
487: \begin{center}
488: \includegraphics[width=9cm,clip=true]{free_energy.eps}
489: \end{center}
490: \caption{\label{fig_F}
491: Free energy $F=-T\log(Z)$ of the anharmonic oscillator
492: as a function of the temperature $T=1/\beta$ with $\beta
493: =na$. The solid line was calculated using the spectrum
494: of the Hamiltonian. The data points were obtained using
495: Monte Carlo calculations and the adiabatic switching
496: method.}
497: \end{figure}
498: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
499:
500: In Fig.~\ref{fig_F} we show Monte Carlo results for the
501: partition function compared to ``exact'' results based on
502: the spectrum of the anharmonic oscillator obtained in
503: Sect.~\ref{sec_diag}. The Monte Carlo results agree with
504: the direct calculations but the Monte Carlo method is
505: effectively limited to a small range of temperatures.
506: If the temperature is very small the partition function
507: is dominated by the ground state contribution. In that case,
508: it is much more efficient to compute the ground state
509: energy directly by measuring the expectation value of the
510: Hamiltonian, $E_0=\langle H\rangle$ , with $H=\dot{x}^2/4
511: +V(x)$. There is one subtlety with this approach: If a
512: naive one-sided discretization of the time derivative is
513: used then the continuum limit of the expectation value
514: of the kinetic energy diverges. This problem can be
515: addressed by using an improved discretization of the kinetic
516: energy \cite{Feynman}, or by using the Virial theorem.
517: The Virial theorem implies that
518: \be
519: \langle H \rangle = \langle T+V \rangle
520: = \langle \frac{x}{2}V'+V \rangle .
521: \ee
522: At high temperature more and more states contribute. The
523: main difficulty with the Monte Carlo approach in this
524: regime is that discretization errors have to be carefully
525: monitored.
526:
527: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
528: \section{Extracting the Instanton Content using Cooling}
529: \label{sec_cool}
530: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
531:
532: From Fig.~\ref{fig_path} we can clearly see that for this particular
533: choice of the parameter $\eta$ a typical path contains two
534: components, one related to quantum fluctuations with frequency
535: $\omega$, and one related to tunneling events, instantons. In
536: the continuum limit the instanton solution can be found from
537: the classical equation of motion
538: \be
539: \frac{\delta}{\delta x(\tau)} S_E = 0
540: \hspace{0.5cm}\Rightarrow\hspace{0.5cm}
541: m\ddot{x}=V'(x).
542: \ee
543: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
544: \begin{figure}
545: \begin{center}
546: \leavevmode
547: \includegraphics[width=8cm,clip=true]{qmcoolcor.eps}
548: \includegraphics[width=8cm,clip=true]{dqmcoolcor.eps}
549: \end{center}
550: \caption{\label{fig_cor_cool}
551: Same as Fig.~\ref{fig_cor} but the correlation functions are
552: evaluated from cooled Monte Carlo configurations. The
553: number of cooling sweeps is $N_{cool}=200$. }
554: \end{figure}
555: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
556: The solution which satisfies the boundary condition
557: $x(\tau\to\pm\infty)=\pm\eta$ is given by
558: \be
559: x_{I}(\tau) = \eta\tanh\left[
560: \frac{\omega}{2}(\tau-\tau_0)\right],
561: \ee
562: where $\omega=4\eta$ and $\tau_0$ is the ``location'' of the
563: instanton. The anti-instanton solution is simply given by
564: $x_{A}(\tau)=-x_I(\tau)$. The classical action of the instanton is
565: \be
566: S_0 = \frac{4\eta^3}{3} .
567: \ee
568: The tunneling rate $n_{I+A}=N_{I+A}/\beta$ is exponentially
569: small, $n_{I+A}\sim\exp(-S_0)$. In order to determine the
570: pre-factor one has to study small fluctuations around the
571: instanton solution. This calculation has been carried out
572: to next-to-leading order in the semi-classical expansion. The
573: result is \cite{Zinn-Justin:1993wc,Kleinert:1995,Wohler:pg}
574: \be
575: \label{idens}
576: n_{I+A}= 8\eta^{5/2} \sqrt{\frac{2}{\pi}}
577: \exp\left(-S_0-\frac{71}{72}\frac{1}{S_0}\right).
578: \ee
579: The tunneling events can be studied in more detail after removing
580: short distance fluctuations. A well known method for doing this
581: is ``cooling'' \cite{Hoek:1985gj,Hoek:1986hq}. In the cooling method
582: we only accept Metropolis updates that lower the action. This will
583: drive the system towards the nearest classical solution. Since instantons
584: are classical solutions, cooling can be used to study the instanton
585: content of a quantum configurations. This is clearly seen
586: in Fig.~\ref{fig_path}. The black line is the original, quantum,
587: configuration. The green line is the same configuration after
588: 200 cooling sweeps. It is easy to check that this configuration
589: is very close to a linear superposition of independent
590: tunneling events. For this purpose we can extract the instanton
591: and anti-instanton locations from the zero crossings and
592: compare the cooled configuration to the simple ``sum ansatz''
593: \be
594: \label{sum}
595: x_{sum}(\tau) = \eta \left\{ \sum_{i} Q_i\tanh\left[
596: \frac{\omega}{2}(\tau-\tau_i)\right] -1 \right\},
597: \ee
598: where $Q_i=\pm 1$ is the topological charge of the instanton.
599: The most important question is to what extent physical observables
600: in the cooled configurations resemble those in the original
601: configurations. This provides a measure of the importance of
602: instantons in the double well potential. In Fig.~\ref{fig_cor_cool}
603: we show correlation functions measured in the cooled configurations.
604: These results should be compared with the full correlation functions
605: shown in Fig.~\ref{fig_cor}. We observe that the correlation functions
606: are quite different. Short distance fluctuations eliminated by cooling
607: obviously play an important role. We observe, however, that the level
608: splitting between the ground state and the first excited
609: state is clearly dominated by semi-classical configurations.
610: The logarithmic derivative of both the $\langle x(0)x(\tau)
611: \rangle$ and $\langle x^3(0)x^3(\tau)\rangle$ correlation
612: functions is very well reproduced in the cooled configurations.
613:
614: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
615: \section{The density of instantons}
616: \label{sec_dens}
617: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
618:
619: The cooling method can also be used in order to get an
620: estimate of the total density of instantons and anti-instantons.
621: While the net topological charge, the number of instantons
622: minus the number of anti-instantons, is unambiguously defined
623: the same is not true for the total number of topological objects.
624: There is no clear distinction between a large quantum fluctuation
625: and a very close instanton-anti-instanton pair. In the cooling
626: method this is reflected by the fact that the number of instantons,
627: extracted from the number of zero crossings in the cooled
628: configuration, depends on the number of cooling sweeps.
629:
630: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
631: \begin{figure}
632: \begin{center}
633: \leavevmode
634: \includegraphics[width=8cm,clip=true]{idens.eps}
635: \includegraphics[width=8cm,clip=true]{sinst_f.eps}
636: \end{center}
637: \caption{\label{fig_n_cool}
638: Instanton density and instanton action as a function of the
639: number of cooling sweeps for different values of the parameter
640: $\eta$. The solid and dashed green lines in Fig.~a shows the
641: semi-classical instanton density at one and two-loop order.
642: The solid line in Fig.~b shows the classical instanton action.}
643: \end{figure}
644: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
645:
646: It is clear, however, that the instanton density should be
647: well defined in the semi-classical limit. In this limit there is an
648: exponentially large separation of scales between the tunneling
649: time $\tau_{tun}$ and the scale of ordinary quantum fluctuations
650: $\tau_{osc}$. This separation of scales can also be exploited
651: in the cooling method. Cooling is a local algorithm which implies
652: that it takes on the order of $\tau/a$ cooling sweeps in order
653: to affect coherent structures that exist at a scale $\tau$. We
654: expect that the number of instantons measured using the cooling
655: method is approximately given by the sum of two exponentials,
656: $N_I(n_{cool})=N_{osc}\exp(-n_{cool}a/\tau_{osc})+N_{tun}\exp(-n_{cool}
657: a/\tau_{tun})$. The first exponential describes the disappearance
658: of quantum fluctuations on a time scale $\tau_{osc}$ and the
659: second exponential reflects instanton-anti-instanton annihilation
660: occurring on a time scale $\tau_{tun}$.
661:
662: Numerical results for $N_I(n_{cool})$ are shown in Fig.~\ref{fig_n_cool}.
663: We observe that the data is consistent with the presence of two
664: distinct time scales and that the description in terms of two
665: exponentials becomes better as the semi-classical limit $\eta\to
666: \infty$ is approached. We also note that after the quantum noise
667: has disappeared the instanton density is close to the semi-classical
668: result equ.~(\ref{idens}). A more detailed comparison is shown
669: in Fig.~\ref{fig_idens}. In this figure we show the instanton
670: density after 10 cooling sweeps, the one and two-loop semi-classical
671: result, as well as the level spacing between the ground state and
672: the first excited state.
673:
674: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
675: \begin{figure}
676: \begin{center}
677: \leavevmode
678: \includegraphics[width=9cm,clip=true]{idens_f_log.eps}
679: \end{center}
680: \caption{\label{fig_idens}
681: Instanton density as a function of the parameter $\eta$.
682: The blue symbols show the instanton density extracted
683: from Monte Carlo configurations after 10 cooling sweeps.
684: The red symbols show the results of a Monte Carlo calculation
685: of non-Gaussian effects. The green lines show the semi-classical
686: instanton density at one and two-loop order. The black line shows
687: $\Delta E/2$ where $\Delta E$ is the splitting between the ground
688: state and the first excited state.}
689: \end{figure}
690: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
691:
692: We observe that for $\eta>1.2$, corresponding to a classical
693: instanton action $S_0>2$, the number of instantons extracted
694: using the cooling method agrees very well with the semi-classical
695: approximation. We also note that the two-loop result is a
696: clear improvement over the one-loop approximation for classical
697: actions as small as $S_0\sim 1$. Finally, we observe that the
698: instanton density is close to the level splitting even in the
699: regime where $S_0$ is less than one.
700:
701: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
702: \begin{figure}
703: \begin{center}
704: \leavevmode
705: \includegraphics[width=9cm,clip=true]{qmidens_conf.eps}
706: \end{center}
707: \caption{\label{fig_ngauss}
708: Quantum mechanical paths which appear in a Monte-Carlo
709: calculation of the one-instanton partition function
710: in the double well potential. The calculation involves
711: adiabatic switching between the Gaussian effective
712: potential and the full potential. The smooth curves
713: are the initial configurations in the zero and one-instanton
714: sector. The Monte Carlo updates in the one-instanton
715: sector involve a constraint which keeps the instanton
716: location fixed.}
717: \end{figure}
718: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
719:
720: In Fig.~\ref{fig_n_cool} we also show a Monte Carlo calculation
721: of the instanton density on a small lattice. The idea is very
722: simple. The one-loop calculation of the tunneling rate is
723: based on expanding the action around the classical path to
724: quadratic order
725: \be
726: \label{s_gauss}
727: S = S_0 + \frac{1}{2} \int d\tau\,
728: \delta x(\tau) \left.\frac{\delta^2 S}{\delta x^2}
729: \right|_{x_I(\tau)} \delta x(\tau) + \ldots ,
730: \ee
731: where $\delta x(\tau)=x(\tau)-x_I(\tau)$. As in Sect.~\ref{sec_latt}
732: we can introduce a new action $S_\alpha$ that interpolates between
733: the full action and the Gaussian approximation, $S_\alpha= S_{gauss}
734: +\alpha \Delta S$ with $\Delta S= S-S_{gauss}$. The exact quantum
735: weight of an instanton can be determined by integrating over the
736: coupling constant $\alpha$. We have
737: \be
738: n = n_{gauss}\exp\left[
739: -\int_0^1 d\alpha \left( \langle \Delta S\rangle_\alpha^{(1)}
740: - \langle \Delta S\rangle_\alpha^{(0)} \right)\right],
741: \ee
742: where $\langle .\rangle_\alpha^{(n)}$ is an expectation value
743: in the $n$-instanton sector at coupling $\alpha$. The method
744: is illustrated in Fig.~\ref{fig_ngauss}. The figure shows typical
745: paths that contribute to $\langle \Delta S\rangle$ in the zero
746: and one-instanton sector. The resulting estimate of the instanton
747: density is also shown in Fig.~\ref{fig_idens}. The Monte Carlo
748: results show that the instanton density is reduced compared
749: to the one-loop estimate. For classical instanton actions
750: $S_0>3$ the result is in agreement with the two-loop estimate
751: and the cooling calculation. It is hard to push the Monte Carlo
752: calculation to instanton actions $S_0<3$ because transitions
753: between the zero and two (four, six, $\ldots$) sector become
754: too frequent.
755:
756: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
757: \section{The instanton liquid model}
758: \label{sec_ilm}
759: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
760:
761: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
762: \begin{figure}
763: \begin{center}
764: \leavevmode
765: \includegraphics[width=8cm,clip=true]{rcor.eps}
766: \includegraphics[width=8cm,clip=true]{drcor.eps}
767: \end{center}
768: \caption{\label{fig_cor_rilm}
769: Same as Fig.~\ref{fig_cor} but the correlation functions are
770: evaluated from a random instanton configuration.}
771: \end{figure}
772: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
773:
774: Given the success of the semi-classical approximation in predicting
775: the splitting between the ground state and the first excited state
776: it seems natural to study the correlation functions in the semi-classical
777: approximation in more detail. We begin by considering the contribution
778: from the classical path only. In this case the partition function is
779: given by
780: \be
781: \label{rilm}
782: Z = \sum_{n_I,n_A}\frac{\delta_{n_I,n_A}}{n_I!n_A!}\left(\prod_{i}
783: \int d\tau_i\right) \exp(-S).
784: \ee
785: Here, $n_I,n_A$ are the number of instanton and anti-instantons,
786: $\tau_i$ are the (anti) instanton positions, and $S_0$ is the
787: classical action. In the next section we will discuss the
788: problem of choosing the correct path for a multi-instanton
789: configuration in more detail. The simplest choice is the sum
790: ansatz given in equ.~(\ref{sum}). The coordinate correlation
791: function is given by
792: \be
793: \Pi_{cl}(\tau) = \langle x_{cl}(0)x_{cl}(\tau) \rangle ,
794: \ee
795: where $\langle . \rangle$ denotes an ensemble average over
796: the collective coordinates $\tau_i$. The distribution of
797: collective coordinates is controlled by the partition function
798: equ.~(\ref{rilm}). The simplest approximation is to ignore
799: the interaction between instantons. In this case the action
800: is $S=(n_I+n_A)S_0$ and the distribution of collective
801: coordinates is random. This is known as the instanton gas
802: model or the random instanton approximation.
803:
804: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
805: \begin{figure}
806: \begin{center}
807: \leavevmode
808: \includegraphics[width=9cm,clip=true]{veff_inst.eps}
809: \end{center}
810: \caption{\label{fig_veff}
811: Gaussian effective potential for small fluctuations
812: around a single instanton path centered at $\tau=0$. }
813: \end{figure}
814: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
815:
816: Correlation functions in a random instanton gas are shown
817: in Fig.~\ref{fig_cor_rilm}. We note that the result is
818: very similar to the cooled correlation functions shown
819: in Fig.~\ref{fig_cor_cool}. This is in agreement with
820: our earlier observation that the cooled configurations
821: are very close to a simple superposition of instantons.
822: Like the the cooling calculation the random instanton
823: gas reproduces the splitting between the ground state
824: and the first excited state, but it does not give a
825: good description of other aspects of the correlation
826: functions.
827:
828: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
829: \begin{figure}
830: \begin{center}
831: \leavevmode
832: \includegraphics[width=9cm,clip=true]{config_gauss.eps}
833: \end{center}
834: \caption{\label{fig_conf_gauss}
835: Typical random instanton configuration and the same
836: configuration with Gaussian fluctuations. The noisy
837: path was generated using 10 heating sweeps in the
838: Gaussian potential around the classical path. This figure
839: should be compared with Fig.~\ref{fig_path}.}
840: \end{figure}
841: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
842:
843: It is clear that the main feature that is missing from the
844: ensemble of classical paths is quantum fluctuations. Quantum
845: fluctuations appear at next order in the semi-classical
846: approximation. We already noted that quantum fluctuations
847: determine the pre-exponential factor in the tunneling
848: rate, see equ.~(\ref{idens}). We can write the path as
849: $x(\tau)=x_{cl}(\tau)+\delta x(\tau)$ where $x_{cl}(\tau)$
850: is the classical path and $\delta x(\tau)$ is the fluctuating
851: part. To second order in $\delta x$ the action is given by
852: equ.~(\ref{s_gauss}). For a single instanton it is possible
853: to determine the propagator $\langle \delta x(0)\delta x(\tau)
854: \rangle$ analytically, see equ.~(39) in \cite{Schafer:1996wv}.
855: For an ensemble of instantons we can approximate the
856: propagator as a sum of contributions due to individual
857: instantons. This is the procedure that is used in the
858: QCD calculations described in
859: \cite{Shuryak:1992jz,Shuryak:1992ke,Schafer:1995pz,Schafer:1995uz}.
860:
861: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
862: \begin{figure}
863: \begin{center}
864: \leavevmode
865: \includegraphics[width=8cm,clip=true]{qmgausscor.eps}
866: \includegraphics[width=8cm,clip=true]{dqmgausscor.eps}
867: \end{center}
868: \caption{\label{fig_cor_gauss}
869: Same as Fig.~\ref{fig_cor} but the correlation functions
870: are evaluated in a random instanton ensemble with Gaussian
871: fluctuations. }
872: \end{figure}
873: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
874:
875: Alternatively, we can determine the correlation function
876: numerically, using the ``heating'' method. As the name
877: suggests, this is essentially the inverse of the cooling
878: method. We begin from a classical path and determine the
879: Gaussian effective potential for small fluctuations around
880: the path. For a single instanton, the action is given by
881: \be
882: S= \int d\tau\, \left(
883: \frac{1}{4} \delta \dot{x}^2(\tau)
884: + 4\eta^2 \left[ 1 -\frac{3}{2\cosh^2(2\eta(\tau-\tau_I))}
885: \right]\delta x^2(\tau) \right),
886: \ee
887: see Fig.~\ref{fig_veff}. This action has one zero mode
888: $\delta x(\tau)=-S_0^{-1/2}dx_{cl}(\tau-\tau_I)/(d\tau_I)$
889: which corresponds to translations of the instanton solution.
890: We can eliminate the corresponding non-Gaussian fluctuations
891: by imposing a constraint on the location of the instanton.
892: Using the simple identity
893: \be
894: 1 = \int d\tau_I\, \delta( x(\tau_I))\,
895: \left|\dot x(\tau_I)\right|
896: \ee
897: we see that the corresponding Jacobian is the velocity
898: $\dot x(\tau_I)$. We can now perform Monte Carlo calculations
899: using the Gaussian action for a multi-instanton configuration.
900: The method is illustrated in Fig.~\ref{fig_conf_gauss}. The
901: black line shows the classical path and the green path is
902: the same path with Gaussian fluctuations included. Clearly,
903: this path looks very similar to the full quantum path
904: shown in Fig.~\ref{fig_path}. There are still some
905: differences, however. We notice, in particular, that the
906: fluctuations around the minima of the potential are
907: not completely symmetric. This is related to non-Gaussian
908: effects. We also observe that the heated random instanton
909: path lacks large excursions from the minima of the potential
910: that do not lead to a tunneling event. These effects are
911: due to a combination of instanton interactions and large
912: non-Gaussian effects.
913:
914: Correlation functions in the random instanton configurations with
915: Gaussian fluctuations included are shown in Fig.~\ref{fig_cor_gauss}.
916: We observe that the correlation functions are in much better
917: agreement with the exact results than the correlators obtained
918: from the classical path only. We also see that the correlators
919: not only describe the splitting between the ground state and
920: the first excited state but also provide a reasonable
921: description of the second excited state.
922:
923: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
924: \section{Instanton interactions}
925: \label{sec_int}
926: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
927:
928: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
929: \begin{figure}
930: \begin{center}
931: \leavevmode
932: \includegraphics[width=9cm,clip=true]{sint_sum.eps}
933: \end{center}
934: \caption{\label{fig_sint}
935: Instanton-anti-instanton interaction in units of $S_0$ as
936: a function of the instanton-anti-instanton separation. The solid
937: line shows the result in the sum ansatz. The triangles show the
938: same data plotted as a function of the zero crossing distance.
939: The streamline interaction is shown as the circles and the
940: squares show the effective interaction extracted from the
941: cooled intanton-anti-instanton distribution shown in
942: Fig.~\ref{fig_zdist}.}
943: \end{figure}
944: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
945:
946: Another feature that is missing from the random instanton
947: ensemble is the correlation between tunneling events due
948: to the interaction between instantons. In QCD instanton
949: interactions, in particular those mediated by fermions, are
950: very important and lead to qualitative changes in the instanton
951: ensemble. In the quantum mechanical model studied here the interaction
952: between instantons is short range and only leads to relatively small
953: effects. These effects can nevertheless be clearly identified in
954: very accurate calculations. We refer to \cite{Schafer:1996wv} for
955: a discussion of the contribution of instanton-anti-instanton
956: pairs to the ground state energy.
957:
958: The simplest method for studying the instanton-anti-instanton
959: interaction is to construct a trial function and compute
960: its action. For the sum ansatz given in equ.~(\ref{sum})
961: the result is shown as the solid line in Fig.~\ref{fig_sint}.
962: Asymptotically, the action is given by $S_{IA}(\tau_{IA})=2S_0
963: (1-6\exp(-\eta\tau_{IA})+\ldots)$ where $\tau_{IA}=|\tau_I-\tau_A|$
964: is the instanton-anti-instanton separation. In the opposite limit,
965: $\tau_{IA}\to 0$, the instanton and anti-instanton annihilate and
966: the action tends to zero. It is clear, however, that in this limit
967: the sum ansatz is at best an approximate solution to the classical
968: equation of motion, and it is not obvious how the path should be
969: chosen.
970:
971: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
972: \begin{figure}
973: \begin{center}
974: \leavevmode
975: \includegraphics[width=8cm,clip=true]{stream_prof.eps}
976: \includegraphics[width=7.8cm,clip=true]{stream_act.eps}
977: \end{center}
978: \caption{\label{fig_stream}
979: Solution of the streamline equation for an instanton-anti-instanton
980: pair. Figure a shows the streamline path and the Fig.~b shows the
981: action density. The paths correspond to $S/S_0=2.0,1.8,\ldots,0.2,
982: 0.1$.}
983: \end{figure}
984: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
985:
986: The best way to deal with this problem is the ``streamline" or
987: ``valley" method \cite{Balitsky:1986qn,Verbaarschot:1991sq}. The
988: method is based on the observation that in the space of all
989: instanton-anti-instanton paths there is one almost flat direction
990: along which the action slowly varies between $2S_0$ and 0. All other
991: directions correspond to perturbative fluctuations. We can force the
992: instanton-anti-instanton path to descend along the almost flat
993: direction by adding a constraint
994: \be
995: S_\xi = \xi(\lambda) \int d\tau\,
996: \left( x(\tau)-x_\lambda(\tau)\right)
997: \frac{dx_\lambda(\tau)}{d\lambda}
998: \ee
999: to the classical action. Here, $\lambda$ labels the different
1000: instanton-anti-instanton paths along the streamline and $\xi(\lambda)$
1001: is a Lagrange multiplier. We find the streamline configuration by
1002: starting from a well separated IA pair and letting the system evolve
1003: using the method of steepest descent. This means that we have to solve
1004: \be
1005: \label{qm_stream}
1006: \xi(\lambda)\frac{dx_\lambda(\tau)}{d\lambda} =
1007: \left.\frac{\delta S}{\delta x (\tau)}\right|_{x=x_\lambda},
1008: \ee
1009: with the boundary condition that $x_{\lambda=0}(\tau)\simeq x_{sum}(\tau)$
1010: corresponds to a well separated instanton-anti-instanton pair. Note that
1011: $\xi(\lambda)$ is an arbitrary function that reflects the reparametrization
1012: invariance of the streamline solution. A sequence of paths obtained by
1013: solving equ.~(\ref{qm_stream}) numerically is shown in Fig.~\ref{fig_stream}.
1014: We also show the action density $s=\dot x^2/4+V(x)$. We can see clearly how
1015: the two localized solutions merge and eventually disappear as the
1016: configuration progresses down the valley.
1017:
1018: There is no unique way to parametrize the streamline path
1019: and extract the instanton-anti-instanton action as a function of the
1020: separation between the tunneling events. The simplest possibility
1021: is to used the distance between the zero crossings $\tau_z$. This
1022: definition has the advantage of being very easy to use, but it
1023: prevents us from exploring the part of the streamline trajectory
1024: where the instanton and anti-instanton are so close that the
1025: path never crosses zero. In Fig.~\ref{fig_sint} we compare results
1026: for $S_{IA}(\tau_z)$ obtained from the sum ansatz and the streamline
1027: solution. We observe that for instanton separations $\tau_z>0.3$ the
1028: results are very similar. We also note, however, that for $\tau_z
1029: <0.6$ the zero crossing distance is quite different from the
1030: parameter $\tau_I-\tau_A$ that appears in the sum ansatz.
1031:
1032: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1033: \begin{figure}
1034: \begin{center}
1035: \leavevmode
1036: \includegraphics[width=9cm,clip=true]{zdist.eps}
1037: \end{center}
1038: \caption{\label{fig_zdist}
1039: Distribution of instanton-anti-instanton separations after
1040: 10 cooling sweeps. The separation was extracted from the
1041: distance of the zero-crossings in the cooled configuration.}
1042: \end{figure}
1043: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1044:
1045: One can show that the ambiguities that arise in trying to
1046: define the instanton-anti-instanton interaction at short
1047: distance correspond to similar ambiguities that arise in the
1048: perturbative expansion because and are related to the factorial
1049: growth of higher order terms in the expansion. Only the sum of
1050: the perturbative and the instanton contribution is well defined
1051: and leads to unique predictions for the groundstate energy.
1052: In these lectures we shall not discuss this problem any
1053: further. Instead, we will study the question whether the
1054: full quantum configurations contain evidence of the correlations
1055: between instantons that the classical instanton-anti-instanton
1056: interaction implies.
1057:
1058: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1059: \begin{figure}
1060: \begin{center}
1061: \leavevmode
1062: \includegraphics[width=9cm,clip=true]{iconf.eps}
1063: \end{center}
1064: \caption{\label{fig_iconf}
1065: Typical instanton configuration in an interacting instanton
1066: calculation. The figure shows the location $x$ of the
1067: first 10 instantons (blue) and anti-instantons (red)
1068: over a period of 3000 configurations.}
1069: \end{figure}
1070: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1071:
1072: In Fig.~\ref{fig_zdist} we show a histogram of the
1073: instanton-anti-instanton separation determined in Monte
1074: Carlo simulations of the quantum mechanical partition function.
1075: The data were obtained by measuring the zero crossing distance
1076: after 10 cooling sweeps. For comparison we also show the $IA$
1077: distribution in the random instanton gas. We observe that there
1078: is an enhancement of close $IA$ pairs which corresponds to an
1079: attractive instanton-anti-instanton interaction. The exact magnitude
1080: of this enhancement depends sensitively on the number of cooling
1081: sweeps. As emphasized in the previous paragraph, this is not necessarily
1082: a shortcoming of the cooling method. We can try to translate the
1083: enhancement in the $IA$ distribution into an effective interaction
1084: using the classical relation $n(\tau_{IA})\sim n_0(\tau_{IA})\exp(-
1085: S_{IA}(\tau_{IA}))$. Here, $n(\tau_{IA})$ is the $IA$ distribution,
1086: $n_0(\tau_{IA})$ is the distribution in the random theory and $S_{IA}
1087: (\tau_{IA})$ is the instanton-anti-instanton interaction. The result
1088: is also shown in Fig.~\ref{fig_sint}. We observe that the interaction
1089: extracted from the $IA$ distribution is significantly weaker than the
1090: classical result. This may imply that the full quantum interaction is
1091: weaker than the classical result, or that too many close pairs are
1092: lost during cooling. This question can be studied in more detail using
1093: the methods discussed in Sect.~\ref{sec_dens}.
1094:
1095: Finally, we address the question how to include correlations
1096: between tunneling events in the instanton calculation. For
1097: this purpose we include the instanton-anti-instanton interaction
1098: in the instanton liquid partition function equ.~(\ref{rilm}).
1099: In this context we again have to address the problem of close
1100: instanton-anti-instanton pairs. The simplest approach is to
1101: add a short range repulsive core which excludes configurations
1102: that are not semi-classical. The hard core interaction can be
1103: adjusted in order to reproduce the $IA$ distribution found in the
1104: cooling calculation. In practice we have chosen $S_{core}(\tau_{IA})=
1105: A_c \exp(-\tau_{IA}/\tau_c)$ with $A_c=3$ and $\tau_c=0.3$. In
1106: Fig.~\ref{fig_iconf} we show a typical set of instanton and
1107: anti-instanton trajectories from an interacting instanton calculation.
1108: Correlations between instantons are clearly visible. In particular,
1109: we observe that a number of close instanton-anti-instanton pairs
1110: are formed. We have also studied correlation functions in the
1111: interacting instanton ensemble. We find that differences as compared
1112: to the random ensemble are rather subtle and a detailed study of
1113: non-Gaussian effects is necessary in order to establish the
1114: importance of instanton interactions.
1115:
1116:
1117: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1118: \section{Summary}
1119: \label{sec_sum}
1120: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1121:
1122: In these lectures we presented Monte Carlo methods for
1123: studying the euclidean path integral in Quantum Mechanics.
1124: We also supply a set of computer codes written in fortran
1125: that were used to generate the data shown in the figures.
1126: We encourage the reader to play around with these programs
1127: in order to get a deeper appreciation of the path integral
1128: and of Monte Carlo methods.
1129:
1130: We should note that Monte Carlo calculations of the
1131: euclidean path integral are an extremely poor way to compute
1132: the spectrum or the correlation functions of the anharmonic
1133: oscillator. The code based on diagonalizing the Hamiltonian
1134: is both much faster and much more accurate than the Monte
1135: Carlo codes. The purpose of the Monte Carlo codes is
1136: entirely pedagogical. However, if we proceed from quantum
1137: mechanics to systems involving many more degrees of freedom,
1138: such as four-dimensional field theories, Hamiltonian methods
1139: become more and more impractical and Monte Carlo calculations
1140: based on the euclidean path integral provide the most efficient
1141: method for computing the spectrum and the correlation functions
1142: known to date.
1143:
1144: We also discussed Monte Carlo methods for studying the
1145: contribution of instantons to the euclidean path integral.
1146: In the case of the double well potential there is a parameter,
1147: $\eta$, which controls the instanton action $S_0=4\eta^3/3$.
1148: If $S_0\gg 1$ then instantons are easily identified but the
1149: tunneling rate is small. If $S_0\sim 1$ then instantons are
1150: very abundant but it is hard to determine the instanton density
1151: precisely. We focused on the case $S_0\sim 3$ which is at
1152: the boundary of the semi-classical regime. Even though the
1153: expansion parameter $1/S_0$ is not very small the instanton
1154: density is still well determined and agrees with the level
1155: splitting. We also noticed, however, that non-Gaussian
1156: effects are important in this regime.
1157:
1158: Ultimately, we are interested in the question to what
1159: extent these results can be generalized to QCD. In QCD
1160: there is no free parameter that controls the applicability
1161: of the semi-classical expansion. Unlike the case of the
1162: double well potential, instantons in QCD can have any
1163: size. Asymptotic freedom implies that the action of small
1164: instantons is big, but the action of instantons with size
1165: $\rho\sim\Lambda_{QCD}^{-1}$ is of order one. Nevertheless,
1166: lattice calculations support the idea that the tunneling density
1167: is sizable, $(N/V)\simeq \Lambda_{QCD}^4 \simeq 1\,{\rm fm}^{-4}$,
1168: and that instantons do not strongly overlap \cite{Chu:1994vi}.
1169: The typical instanton size is found to be $\rho\sim (0.3-0.4)$
1170: fm which implies a typical instanton action $S_0\simeq
1171: (5-10)$. The reason that the density is big even though
1172: the action is significantly larger than one is related
1173: to the fact that QCD instantons have many more collective
1174: coordinates, 12 (4 coordinates, 1 size, 7 color angles)
1175: compared to just one in the case of the double well potential.
1176: As a consequence the pre-exponential factor in the tunneling
1177: rate is numerically large.
1178:
1179: There are some important differences between QCD and the
1180: double well potential. In a typical lattice QCD configuration
1181: quantum fluctuations of the gauge field are much bigger than
1182: the classical gauge fields associated with instantons. This
1183: implies that one cannot ``see'' instantons in the gauge
1184: configurations in the same way that one can immediately
1185: identify tunneling events in the quantum mechanical paths.
1186: Compare, for example, Fig.~\ref{fig_path} with Fig.~1
1187: in \cite{Chu:1994vi}. Only after some amount of cooling do
1188: instantons emerge from the quantum noise. On the other hand,
1189: fermions provide an important diagnostic tool in QCD that is
1190: not available in the simple bosonic model analyzed in these
1191: lectures. Instantons lead to localized chiral zero modes of
1192: the Dirac operators that can easily be identified even in noisy
1193: quantum configurations.
1194:
1195: \begin{thebibliography}{20}
1196:
1197: \bibitem{Polyakov:1976fu}
1198: A.~M.~Polyakov,
1199: %``Quark Confinement And Topology Of Gauge Groups,''
1200: Nucl.\ Phys.\ B {\bf 120}, 429 (1977).
1201: %%CITATION = NUPHA,B120,429;%%
1202:
1203: \bibitem{Vainshtein:1981wh}
1204: A.~I.~Vainshtein, V.~I.~Zakharov, V.~A.~Novikov and M.~A.~Shifman,
1205: {\it ABC Of Instantons},
1206: Sov.\ Phys.\ Usp.\ {\bf 24}, 195 (1982);
1207: also in ITEP Lectures on Particle Physics and Field theory,
1208: World Scientific, Singapore (1999).
1209: %%CITATION = SOPUA,24,195;%%
1210:
1211: \bibitem{Coleman:1978ae}
1212: S.~Coleman,
1213: {\it The Uses Of Instantons},
1214: Lecture delivered at 1977 Int. School of Subnuclear Physics,
1215: Erice, Italy (1977); also in Aspects of Symmetry,
1216: Cambridge University Press (1985).
1217:
1218: \bibitem{Shuryak:1988ck}
1219: E.~V.~Shuryak,
1220: {\it The QCD Vacuum, Hadrons And The Superdense Matter},
1221: World Scientific, Singapore (1988).
1222: %%CITATION = 00327,8,1;%%
1223:
1224: \bibitem{Kleinert:1995}
1225: H.~Kleinert,
1226: {\it Path integrals in quantum mechanics, statistics, and polymer physics},
1227: World Scientific, Singapore (1995).
1228:
1229: \bibitem{Zinn-Justin:1993wc}
1230: J.~Zinn-Justin,
1231: {\it Quantum field theory and critical phenomena},
1232: Oxford University Press (1993).
1233: %%CITATION = IMPHA,85,1;%%
1234:
1235: \bibitem{Schafer:1996wv}
1236: T.~Sch{\"a}fer and E.~V.~Shuryak,
1237: Rev.\ Mod.\ Phys.\ {\bf 70}, 323 (1998),
1238: [hep-ph/9610451].
1239: %%CITATION = HEP-PH 9610451;%%
1240:
1241: \bibitem{Forkel:2000sq}
1242: H.~Forkel,
1243: %``A primer on instantons in QCD,''
1244: preprint, hep-ph/0009136.
1245: %%CITATION = HEP-PH 0009136;%%
1246:
1247: \bibitem{Feynman}
1248: R.~P.~Feynman, A.~R.~Hibbs,
1249: {\it Quantum Mechanics and Path Integrals},
1250: McGraw-Hill, New York (l965).
1251:
1252: \bibitem{Creutz:1980gp}
1253: M.~Creutz and B.~Freedman,
1254: %``A Statistical Approach To Quantum Mechanics,''
1255: Annals Phys.\ {\bf 132}, 427 (1981).
1256: %%CITATION = APNYA,132,427;%%
1257:
1258: \bibitem{Shuryak:1984xr}
1259: E.~V.~Shuryak and O.~V.~Zhirov,
1260: %``Testing Monte Carlo Methods For Path Integrals In Some Quantum Mechanical
1261: %Problems,''
1262: Nucl.\ Phys.\ B {\bf 242}, 393 (1984).
1263: %%CITATION = NUPHA,B242,393;%%
1264:
1265: \bibitem{Shuryak:1987tr}
1266: E.~V.~Shuryak,
1267: %``Toward The Quantitative Theory Of The 'Instanton Liquid' 4. Tunneling In The Double Well Potential,''
1268: Nucl.\ Phys.\ B {\bf 302}, 621 (1988).
1269: %%CITATION = NUPHA,B302,621;%%
1270:
1271: \bibitem{Asakawa:2000tr}
1272: M.~Asakawa, T.~Hatsuda and Y.~Nakahara,
1273: %``Maximum entropy analysis of the spectral functions in lattice QCD,''
1274: Prog.\ Part.\ Nucl.\ Phys.\ {\bf 46}, 459 (2001)
1275: [hep-lat/0011040].
1276: %%CITATION = HEP-LAT 0011040;%%
1277:
1278: \bibitem{Jarrel:1996}
1279: M.~Jarrell, and J.~E.~Gubernatis,
1280: % "Bayesian Inference and the Analytic Continuation .. ",
1281: Phys.\ Rep.\ {\bf 269}, 133 (1996).
1282:
1283: \bibitem{Wohler:pg}
1284: C.~F.~Wohler and E.~V.~Shuryak,
1285: %``Two Loop Corr To The Instanton Density For The Double Well Potential,''
1286: Phys.\ Lett.\ B {\bf 333}, 467 (1994)
1287: [hep-ph/9402287].
1288: %%CITATION = HEP-PH 9402287;%%
1289:
1290: \bibitem{Hoek:1985gj}
1291: J.~Hoek,
1292: %``Cooling Of SU(3) Lattice Gauge Field Configurations And The Eta-Prime
1293: %Mass,''
1294: Phys.\ Lett.\ B {\bf 166}, 199 (1986).
1295: %%CITATION = PHLTA,B166,199;%%
1296:
1297: \bibitem{Hoek:1986hq}
1298: J.~Hoek, M.~Teper and J.~Waterhouse,
1299: %``Topology And The Eta-Prime Mass In SU(3) Lattice Gauge Theory,''
1300: Phys.\ Lett.\ B {\bf 180}, 112 (1986).
1301: %%CITATION = PHLTA,B180,112;%%
1302:
1303: \bibitem{Shuryak:1992jz}
1304: E.~V.~Shuryak and J.~J.~M.~Verbaarschot,
1305: %``Quark propagation in the random instanton vacuum,''
1306: Nucl.\ Phys.\ B {\bf 410}, 37 (1993)
1307: [hep-ph/9302238].
1308: %%CITATION = HEP-PH 9302238;%%
1309:
1310: \bibitem{Shuryak:1992ke}
1311: E.~V.~Shuryak and J.~J.~M.~Verbaarschot,
1312: %``Mesonic correlation functions in the random instanton vacuum,''
1313: Nucl.\ Phys.\ B {\bf 410}, 55 (1993)
1314: [hep-ph/9302239].
1315: %%CITATION = HEP-PH 9302239;%%
1316:
1317: %\cite{Schafer:1995pz}
1318: \bibitem{Schafer:1995pz}
1319: T.~Sch{\"a}fer and E.~V.~Shuryak,
1320: %``The instanton liquid in QCD at zero and finite temperature,''
1321: Phys.\ Rev.\ D {\bf 53}, 6522 (1996)
1322: [hep-ph/9509337].
1323: %%CITATION = HEP-PH 9509337;%%
1324:
1325: \bibitem{Schafer:1995uz}
1326: T.~Sch{\"a}fer and E.~V.~Shuryak,
1327: %``Hadronic Correlation Functions in the Interacting Instanton Liquid,''
1328: Phys.\ Rev.\ D {\bf 54}, 1099 (1996)
1329: [hep-ph/9512384].
1330: %%CITATION = HEP-PH 9512384;%%
1331:
1332: \bibitem{Balitsky:1986qn}
1333: I.~I.~Balitsky and A.~V.~Yung,
1334: %``Collective-Coordinate Method For Quasizero Modes,''
1335: Phys.\ Lett.\ B {\bf 168}, 113 (1986).
1336: %%CITATION = PHLTA,B168,113;%%
1337:
1338: \bibitem{Verbaarschot:1991sq}
1339: J.~J.~M.~Verbaarschot,
1340: %``Streamlines and conformal invariance in Yang-Mills theories,''
1341: Nucl.\ Phys.\ B {\bf 362}, 33 (1991)
1342: [Erratum-ibid.\ B {\bf 386}, 236 (1992)].
1343: %%CITATION = NUPHA,B362,33;%%
1344:
1345: \bibitem{Chu:1994vi}
1346: M.~C.~Chu, J.~M.~Grandy, S.~Huang and J.~W.~Negele,
1347: %``Evidence for the role of instantons in hadron structure from lattice QCD,''
1348: Phys.\ Rev.\ D {\bf 49}, 6039 (1994)
1349: [hep-lat/9312071].
1350: %%CITATION = HEP-LAT 9312071;%%
1351:
1352:
1353: \end{thebibliography}
1354:
1355: \newpage
1356:
1357: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1358: \appendix
1359: \section{Computer Codes}
1360: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1361:
1362: \noindent
1363: All programs are written in standard fortran 77, have extensive
1364: comments and do not require any libraries. Some of the programs
1365: contain subroutine for generating random numbers or for diagonalizing
1366: matrices that were taken from Numerical Recipes (Numerical Recipes
1367: in Fortran, W.~H.~Press, S.~A.~Teukolsky, W.~T.~Vetterling and
1368: B.~D.~Flannery, Cambridge University Press).
1369: \vspace*{0.5cm}
1370:
1371: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1372: \noindent
1373: {\bf 1. qmdiag.for}
1374: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1375:
1376: \noindent
1377: This programs computes the spectrum and the eigenfunctions
1378: of the anharmonic oscillator. The results are used in order
1379: to compute euclidean correlation functions.
1380:
1381: \noindent
1382: Input: {\tt fort.05}
1383:
1384: \begin{tabular}{p{2.2cm}p{14cm}}
1385: $f$ & minimum of anharmonic oscillator potential $V(x)=(x^2-f^2)^2$ \\
1386: $N$ & dimension of basis used for diagonalizing $H$ (choose $N\geq 40$)\\
1387: $\omega_0$ & unperturbed oscillator frequency (choose $\omega_0\sim 4f$)
1388: \end{tabular}
1389: \vspace*{0.3cm}
1390:
1391: \noindent
1392: Output: {\tt qmdiag.dat}
1393:
1394: \begin{tabular}{p{2.2cm}p{13cm}}
1395: $E_n$ & eigenvalue of Hamiltonian \\
1396: $c_n$ & dipole matrix element $c_n^2=|\langle 0|x|n\rangle|^2$\\
1397: $d_n$ & quadrupole matrix element $d_n^2=|\langle 0|x^2|n\rangle|^2$\\
1398: $e_n$ & quadrupole matrix element $d_n^2=|\langle 0|x^3|n\rangle|^2$\\
1399: $\psi(x)$ & ground state wave function \\
1400: $\langle x(0)x(\tau)\rangle$ & euclidean correlation function,
1401: also for $x^2$ and $x^3$\\
1402: $d\log\Pi/(d\tau)$ & log derivative of $\Pi(\tau)= \langle x(0)x(\tau)
1403: \rangle$ \\
1404: $Z(\beta)$ & partition function
1405: \end{tabular}
1406:
1407: \newpage
1408: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1409: \noindent
1410: {\bf 2. qm.for}
1411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1412:
1413: \noindent
1414: This programs computes correlation functions of the anharmonic
1415: oscillator using Monte Carlo simulations on a euclidean
1416: lattice.
1417:
1418: \noindent
1419: Input: {\tt fort.05}
1420:
1421: \begin{tabular}{p{2.0cm}p{13cm}}
1422: $f$ & minimum of anharmonic oscillator potential $V(x)=(x^2-f^2)^2$ \\
1423: $n$ & number of lattice points in the euclidean time direction
1424: ($n\sim 800$) \\
1425: $a$ & lattice spacing ($a\sim 0.05$) \\
1426: $ih$& $ih=0$ cold start $x_i=-f$, $ih=1$ hot start $x_i=ran()$ \\
1427: $n_{eq}$ & number of equilibration sweeps before first measurement
1428: $(n_{eq}\sim 100)$ \\
1429: $n_{mc}$ & number of Monte Carlo sweeps ($n_{mc}\sim 10^5$) \\
1430: $\delta x$ &width of Gaussian distribution used for Monte Carlo
1431: update $x_i^{(n)}\to x_i^{(n+1)}$ ($\delta x \sim 0.5$) \\
1432: $n_{p}$ & number of points on which correlation functions are
1433: measured $\langle x_ix_{i+1}\rangle,\ldots,$
1434: $\langle x_i x_{i+n_p}\rangle$ $(n_p\sim 20)$ \\
1435: $n_{mea}$ & number of measurements of the correlation function
1436: in a given Monte Carlo configuration ${x_i}$ ($n_{mea}\sim 5$) \\
1437: $n_{pri}$ & number of Monte Carlo configurations between output
1438: of averages to output file ($n_{pri}\sim 100$)
1439: \end{tabular}
1440: \vspace*{0.3cm}
1441:
1442: \noindent
1443: Output: {\tt qm.dat}
1444:
1445: \begin{tabular}{p{2.0cm}p{13cm}}
1446: $S_{tot}$ &average total action per configuration \\
1447: $V_{av},T_{av}$ & average potential and kinetic energy \\
1448: $\langle x^n\rangle$ & expectation value $\langle x^n\rangle$
1449: ($n=1,\ldots,4$) \\
1450: $\Pi(\tau)$ & euclidean correlation function $\Pi(\tau) =
1451: \langle O(0)O(\tau)\rangle$ for $O=x$, $x^2$, $x^3$.
1452: Results are given in the format $\tau, \Pi(\tau)$,
1453: $\Delta\Pi(\tau)$, $d\log(\Pi)/(d\tau)$, $\Delta[
1454: d\log(\Pi)/(d\tau)]$, where $\Delta\Pi(\tau)$ is
1455: the statistical error in $\Pi(\tau)$.
1456: \end{tabular}
1457:
1458:
1459: \newpage
1460: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1461: \noindent
1462: {\bf 3. qmswitch.for}
1463: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1464:
1465: \noindent
1466: The program {\tt qmswicth.for} computes the free energy $F=-T\log(Z)$
1467: of the anharmonic oscillator using the method of adiabatic switching
1468: between the harmonic and the anharmonic oscillator. The action is
1469: $S_\alpha=S_0+\alpha(S-S_0)$. The code switches from $\alpha=0$ to
1470: $\alpha=1$ and then back to $\alpha=0$. Hysteresis effects are
1471: used in order to estimate errors from incomplete equilibration.
1472: Most input parameters are the same as in {\tt qm.for}. Additional
1473: parameters are given below.
1474:
1475: \noindent
1476: Input: {\tt fort.05}
1477:
1478: \begin{tabular}{p{2.0cm}p{13cm}}
1479: $\omega_0$ & oscillator constant of the reference system
1480: ($\omega_0\sim 4f$) \\
1481: $n_{switch}$ & number of steps in adiabatic switching ($n_{switch}
1482: \sim 20$) \\
1483: \end{tabular}
1484: \vspace*{0.3cm}
1485:
1486: \noindent
1487: Output: {\tt qmswitch.dat}
1488:
1489: \noindent
1490: The output file contains many details of the adiabatic switching
1491: procedure. The final result for the free energy is given as $F=F_0
1492: +\delta F$, where $F_0$ is the free energy of the harmonic oscillator
1493: and $\delta F$ is the integral over $\alpha$. We estimate the uncertainty
1494: in the final result as $F\pm \Delta F(stat)\pm \Delta F(equ) \pm \Delta
1495: F(disc)$, where $\Delta F(stat)$ is the statistical error, $\Delta F(equ)$
1496: is due to incomplete equilibration (hysteresis), and $\Delta F(disc)$ is
1497: due to discretizing the $\alpha$ integral.
1498:
1499: \vspace*{0.3cm}
1500:
1501: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1502: \noindent
1503: {\bf 4. qmcool.for}
1504: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1505:
1506: \noindent
1507: This programs is identical to ${\tt qm.for}$ except that
1508: expectation values are measured both in the original
1509: and in cooled configurations. We only specify additional
1510: input parameters.
1511:
1512: \noindent
1513: Input: {\tt fort.05}
1514:
1515: \begin{tabular}{p{2.0cm}p{13cm}}
1516: $n_{st}$ & number of Monte Carlo configurations between successive
1517: cooled configurations. The number of cooled configurations
1518: is $n_{conf}/n_{st}$ ($n_{st}\sim 20$). \\
1519: $n_{cool}$ & number of cooling sweeps ($n_{cool}\sim 50$)
1520: \end{tabular}
1521: \vspace*{0.3cm}
1522:
1523: \noindent
1524: Output: {\tt qmcool.dat}
1525:
1526: \begin{tabular}{p{2.0cm}p{13cm}}
1527: $\Pi(\tau)$ & correlation functions are given in the same format
1528: as in ${\tt qm.dat}$ \\
1529: $N_{I+A}$ & total number of instantons extracted from number
1530: of zero crossings \\
1531: & as a function of the number of cooling sweeps \\
1532: $S_{tot}$ & total action vs number of cooling sweeps \\
1533: $S/N$ & action per instanton. $S_0$ is the continuum
1534: result for one instanton
1535: \end{tabular}
1536: \vspace*{0.3cm}
1537:
1538: \newpage
1539: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1540: \noindent
1541: {\bf 5. qmidens.for}
1542: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1543:
1544: \noindent
1545: The program {\tt qmidens.for} calculates non-Gaussian corrections
1546: to the instanton density using adiabatic switching between the
1547: Gaussian action and the full action. The calculation is performed
1548: in both the zero and one-instanton-sector. The details of the
1549: adiabatic switching procedure are very similar to the method
1550: used in {\tt qmswitch.for}. Note that the total length of
1551: the euclidean time domain, $\beta=na$, cannot be chosen too
1552: large in order to suppress transitions between the one-instanton
1553: sector and the three, five, etc.~instanton sector. Most input
1554: parameters are defined as in {\tt qm.for}.
1555:
1556: \noindent
1557: Input: {\tt fort.05}
1558:
1559: \begin{tabular}{p{2.0cm}p{13cm}}
1560: $n_{switch}$ & number of steps in adiabatic switching ($n_{switch}
1561: \sim 20$) \\
1562: \end{tabular}
1563: \vspace*{0.3cm}
1564:
1565: \noindent
1566: Output: {\tt qmidens.dat}
1567:
1568: \noindent
1569: The output file contains many details of the adiabatic
1570: switching procedure. The final result for the instanton
1571: density is compared to the Gaussian (one-loop) approximation.
1572: Note that the method breaks down if $f$ is too small or
1573: $\beta$ is too large.
1574:
1575: \newpage
1576: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1577: \noindent
1578: {\bf 6. rilm.for}
1579: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1580:
1581: \noindent
1582: This program computes correlation functions of the anharmonic
1583: oscillator using a random ensemble of instantons. The
1584: multi-instanton configuration is constructed using the
1585: sum ansatz. Note that, in contrast to RILM calculations
1586: in QCD, the fields and correlation functions are computed
1587: on a lattice.
1588:
1589: \noindent
1590: Input: {\tt fort.05}
1591:
1592: \begin{tabular}{p{2.0cm}p{13cm}}
1593: $f$ & minimum of anharmonic oscillator potential $V(x)=(x^2-f^2)^2$ \\
1594: $n$ & number of lattice points in the euclidean time direction
1595: ($n\sim 800$) \\
1596: $a$ & lattice spacing ($a\sim 0 05$) \\
1597: $N_{I+A}$ & number of instantons (has to be even). The program
1598: displays the one and two-loop result for the parameters
1599: $(f,\beta=na)$. \\
1600: $n_{mc}$ & number of configurations ($n_{mc}\sim 10^3$) \\
1601: $n_{p}$ & number of points on which correlation functions are
1602: measured $\langle x_ix_{i+1}\rangle,\ldots,$
1603: $\langle x_i x_{i+n_p}\rangle$ $(n_p\sim 20)$ \\
1604: $n_{mea}$ & number of measurements of the correlation function
1605: in a given Monte Carlo configuration ${x_i}$
1606: ($n_{mea}\sim 5$) \\
1607: $n_{pri}$ & number of Monte Carlo configurations between output
1608: of averages to output file ($n_{pri}\sim 100$)
1609: \end{tabular}
1610: \vspace*{0.3cm}
1611:
1612: \noindent
1613: Output: {\tt rilm.dat}
1614:
1615: \begin{tabular}{p{2.0cm}p{13cm}}
1616: $S_{tot}$ &average total action per configuration. \\
1617: $V_{av},T_{av}$ & average potential and kinetic energy.\\
1618: $\langle x^n\rangle$ & expectation value $\langle x^n\rangle$
1619: ($n=1,\ldots,4$) \\
1620: $\Pi(\tau)$ & euclidean correlation function $\Pi(\tau) =
1621: \langle O(0)O(\tau)\rangle$ for $O=x$, $x^2$, $x^3$.
1622: Results are given in the format $\tau$, $\Pi(\tau)$,
1623: $\Delta\Pi(\tau)$, $d\log(\Pi)/(d\tau)$, $\Delta[
1624: d\log(\Pi)/(d\tau)]$.
1625: \end{tabular}
1626: \vspace*{0.3cm}
1627:
1628: \newpage
1629: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1630: \noindent
1631: {\bf 7. rilm\_gauss.for}
1632: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1633:
1634: \noindent
1635: This program generates the same random instanton ensemble as
1636: {\tt rilm.for} but it also includes Gaussian fluctuations
1637: around the classical path. This is done by performing a
1638: few heating sweeps in the Gaussian effective potential.
1639: Most input parameters are defined as in {\tt rilm.for}.
1640: Additional input parameters are given below.
1641:
1642: \noindent
1643: Input: {\tt fort.05}
1644:
1645: \begin{tabular}{p{2.0cm}p{13cm}}
1646: $n_{heat}$ & number of heating steps ($n_{heat} \sim 10$) \\
1647: $\delta x$ & coordinate update ($\delta x\sim 0.5$)
1648: \end{tabular}
1649:
1650: \vspace*{0.5cm}
1651:
1652: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1653: \noindent
1654: {\bf 8. iilm.for}
1655: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1656:
1657: \noindent
1658: This program computes correlation functions of the anharmonic
1659: oscillator using an interacting ensemble of instantons. The
1660: multi-instanton configuration is constructed using the sum
1661: ansatz. The configuration is discretized on a lattice and
1662: the total action is computed using the discretized lattice
1663: action. Very close instanton-anti-instanton pairs are
1664: excluded by adding an nearest neighbor interaction with
1665: a repulsive core. Most input parameters are the same as
1666: in ${\tt rilm.for}$. Additional input parameters are
1667:
1668: \noindent
1669: Input: {\tt fort.05}
1670:
1671: \begin{tabular}{p{2.0cm}p{13cm}}
1672: $\tau_{core}$ & range of hard core interaction ($\tau_{core}\sim 0.3$) \\
1673: $A_{core}$ & strength of hard core interaction ($A_{core}\sim 3.0$) \\
1674: $dz$ & average position update ($dz\sim 1$)
1675: \end{tabular}
1676:
1677:
1678: \end{document}
1679:
1680:
1681:
1682:
1683:
1684: