• Keine Ergebnisse gefunden

Anomalous shell effect in the transition from a circular to a triangular billiard

N/A
N/A
Protected

Academic year: 2022

Aktie "Anomalous shell effect in the transition from a circular to a triangular billiard"

Copied!
13
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Anomalous shell effect in the transition from a circular to a triangular billiard

Ken-ichiro Arita1,2and Matthias Brack2

1Department of Physics, Nagoya Institute of Technology, 466-8555 Nagoya, Japan

2Institute for Theoretical Physics, University of Regensburg, D-93040 Regensburg, Germany 共Received 5 November 2007; revised manuscript received 14 February 2008; published 20 May 2008兲 We apply periodic orbit theory to a two-dimensional nonintegrable billiard system whose boundary is varied smoothly from a circular to an equilateral triangular shape. Although the classical dynamics becomes chaotic with increasing triangular deformation, it exhibits an astonishingly pronounced shell effect on its way through the shape transition. A semiclassical analysis reveals that this shell effect emerges from a codimension-2 bifurcation of the triangular periodic orbit. Gutzwiller’s semiclassical trace formula, using a global uniform approximation for the bifurcation of the triangular orbit and including the contributions of the other isolated orbits, describes very well the coarse-grained quantum-mechanical level density of this system. We also discuss the role of discrete symmetry for the large shell effect obtained here.

DOI:10.1103/PhysRevE.77.056211 PACS number共s兲: 05.45.Mt, 03.65.Sq

I. INTRODUCTION

The periodic orbit theory 共POT兲 is a very useful tool to study shell structure in single-particle energy spectra. In POT, the quantum-mechanical level density is semiclassi- cally approximated in terms of the periodic orbits of the corresponding classical system. For systems with only iso- lated orbits, Gutzwiller derived the so-called “trace formula”

关1兴which is particularly successful for chaotic systems. The POT for three-dimensional 共3D兲 cavities was developed in 关2兴. In integrable systems, semiclassical trace formulas can be derived from torus quantization 关3,4兴. However, most physical systems lie between the above two extreme situa- tions; i.e., they exhibit mixed phase-space dynamics in which both regular and chaotic motion coexist on the same energy shell. For systems with various types of continuous symmetries, trace formulas have been derived in 关4–6兴.

For an introductory textbook on semiclassical physics and applications of the POT to various physical systems, we refer to关7兴.

In both integrable and mixed systems, bifurcations of periodic orbits can significantly influence the shell struc- ture. The above-mentioned semiclassical trace formulas, which are all based on the stationary-phase approximation 共SPA兲, diverge when bifurcations of periodic orbits occur under variations of energy or potential parameters 共e.g., deformations兲. In the SPA, the classical action integral is expanded around its stationary points 共corresponding to pe- riodic orbits兲up to quadratic terms and the trace integral is evaluated in terms of Gauss-Fresnel integrals. At bifurcation points, the determinant of the coefficient matrix of the qua- dratic terms becomes zero and the Gauss-Fresnel integral becomes singular. In order to obtain a finite semiclassical level density near bifurcation points, higher-order expansion terms of the action integral are needed, which are most con- veniently found from the normal forms appropriate to the types of bifurcation under consideration 关8–10兴.

Quantum billiards 共and their 3D versions, cavities兲, be- sides being quite useful toy models to study POT, reflect important features of finite physical quantum systems such as quantum dots, metallic clusters, and atomic nuclei. E.g., nonintegrable 3D cavities with realistic shapes appropriate for fission barriers of actinide nuclei have been used in POT to explain the onset of the mass asymmetry of nascent fission fragments 关11兴. On the other hand, many integrable billiard systems are well known and fully understood semiclassically—e.g., circular, equilateral triangular, square, and elliptic billiards 共cf. 关7兴兲—and may be used as simple models for physical systems. In these, bifurcations of short periodic orbits may lead to a considerable enhancement of shell effects. E.g., in the elliptic billiard, the short diametric orbit undergoes successive bifurcations with increasing de- formation, and new periodic-orbit families emerge关9,12,13兴.

The same type of bifurcations occur for equatorial orbits in the 3D spheroidal cavity关14兴and provide a schematic expla- nation of nuclear superdeformed and hyperdeformed shell structure关15,16兴. In these studies, it was shown that a system may turn strongly chaotic by adding small reflection- asymmetric共e.g., octupole兲deformations关17兴.

In this paper, we apply the POT to a two-dimensional nonintegrable billiard system whose boundary is continu- ously varied from a circular to an equilateral triangular shape. This study is initiated to explore a possible quantum dot system. Many studies have been undertaken for quantum dots, but this type of deformation has not been studied be- fore. Another important aim is to investigate the role of dis- crete symmetries. As will be discussed, the present model possesses discreteC3vpoint-group symmetry. Recently, sev- eral mean-field studies of nuclei have suggested the possible existence of low-lying states with exotic shapes with point- group symmetries, such as tetrahedral and octahedral defor- mations关18兴. In order for such shapes to be stabilized, rather large quantum shell effects in the single-particle spectra are necessary. We will discuss the role of discrete symmetries in establishing strong shell effects.

(2)

II. SHELL STRUCTURE AND LEVEL STATISTICS A. Model system

We consider the two-dimensional billiard system H= p2

2m+Vr兲, Vr兲=

⬁,0, rrR共R共兲,兲,

2.1

in polar coordinates r=共r,␪兲, whereby the boundary shape R共␪兲is parametrized implicitly by the equation

R2+2

3␣ 9

R3

R0cos共3␪兲=R02, ␪僆关0,2␲兲. 共2.2兲 Figure1shows the shape of the boundary for several values of ␣. This system possesses the discrete symmetries of the point group C3v, consisting of ⫾2␲/3 rotations about the origin and reflections with respect to the three axes through the origin with ␪= 0 ,⫾␲/3. The deformation parameter ␣

= 0 yields a circular shape and ␣= 1 an equilateral triangle.

The system is integrable in these two limits, but noninte- grable in between.

For the calculation of the quantum spectrum of this sys- tem, it is useful to decompose the wave functions into free circular waves:

kr,␪兲=m=−⬁

cmJmkreim, 2.3

wherek=

2me/បis the wave number andethe energy.

Taking theC3vsymmetry into account, we can classify the eigenstates according to the eigenvalues of the symmetry operators

R共␭␮兲=e2␲i␭/3共␭␮兲, 共2.4a兲 P共␭␮兲=共− 1兲共␭␮兲, 共2.4b兲 whereRandPrepresent rotation by 2␲/3 around the origin and reflection with respect to the x axis, respectively. We have simultaneous eigenstates of R and P for ␭= 0 states, and then we can classify the eigenstates into four sets, which

correspond to the irreducible representations共“irreps”兲of the C3v point group关19兴:

k0+共r,␪兲=

m=0

cm0+J3m共kr兲cos共3m␪兲, 共2.5a兲

k共0−兲共r,␪兲=

m=1

cm共0−兲J3m共kr兲sin共3m␪兲, 共2.5b兲

k共⫾1兲共r,␪兲=m=−⬁

cm共⫾1兲J3m1共kr兲ei3m1兲␪. 共2.5c兲

k共+1兲 are the complex conjugates of␺k共−1兲, and the two form degenerate pairs of states. 共The point group C3v has two 1-dimensional irreps and one 2-dimensional irrep. The states 共0⫾兲correspond to the 1-dimensional irreps, and the degen- erate pairs of states 共⫾1兲 correspond to the 2-dimensional irrep.兲 Taking linear combinations of these, one finds the following alternative real expressions for the states共2.5c兲:

k共1⫾兲r,␪兲=m=−⬁

cm共1⫾兲J兩3m+1兩kr

cos关共3msin关共3m+ 1+ 1兲

,

共2.5c⬘兲 which are not eigenstates of the operator R, but of the op- erator P. The eigenvalue spectrum兵kn其 and the coefficients cm are determined by the Dirichlet boundary condition

kn共R,␪兲= 0. We show some eigenfunctions in Fig. 2. The -2

-1 0 1 2

-2 -1 0 1 2

y/R0

x/R0 α=0.0

0.4 0.81.0

FIG. 1. Shapes of the boundary given by Eq. 共2.2兲 for several values of deformation parameter␣.

λµ=0+ 0− 1+ 1−

03.778

05.115

06.637

07.733

07.979 02.409

05.235

06.475 06.387

08.143

09.181 09.338

FIG. 2. Wave functions of some lowest eigenstates for ␣= 0.4.

0+, 0−, and 1⫾are states of type共2.5a兲,共2.5b兲, and共2.5c⬘兲, respec- tively. Theirkvalues are also indicated.

(3)

states belonging to the sets共2.5a兲and共2.5b兲have symmetry under the rotation R. The former are even under the reflec- tion P, while the latter are odd.

Figure 3 shows the lower part of the energy spectrum 兵en其=兵ប2kn2/2m其, plotted against the deformation parameter

␣. We note that the levels bunch into small energy intervals in the region ␣⯝0.5– 0.7, forming large gaps in the spec- trum. It is, in fact, quite surprising to realize that the shell gaps in this nonintegrable region are much larger than in the two integrable limits ␣= 0 and ␣= 1. The gaps cause large shell effects in the total energy of a system ofNnoninteract- ing fermions described by the Hamiltonian共2.1兲. We split the energy into a smooth and an oscillating part

EN兲= 2

n=1 N/2

en=N兲+␦EN兲 共N even兲, 共2.6兲 whereby the factor of 2 accounts for the spin s= 1/2. The smooth partN兲is equivalently given by a Strutinsky aver- aging 关20兴, the extended Thomas-Fermi model, or the Weyl expansion共cf关7兴, Chap. 4兲, while the shell-correction energy

E共N兲 reflects the quantum effects; it is dominated by the shortest periodic orbits of the classical system as demon- strated below for the case of the level density. Large gaps in the spectrum lead to large amplitudes of␦E共N兲. This is dem- onstrated in Fig. 4 where we present ␦E共N兲, scaled by a factor

N, for three values of␣. We note that the oscillatory pattern for ␣= 0.5 is quite regular and on the average has a much larger amplitude than in the integrable limits.

B. Level statistics

Nearest-neighbor spacing 共NNS兲 distributions are com- monly used to identify signatures of chaos in a quantum

system. Generically, classically chaotic systems exhibit level repulsion and the NNS distributions are of Wigner type, while regular systems typically have degeneracies and the NNS distributions are Poisson like关21兴. To extract these uni- versal fluctuation properties, one has to useunfoldedspectra whose mean level density is normalized to unity. Thus, for systems with discrete symmetries, one has to study the NNS of the subsets of levels belonging to the irreps of the corre- sponding point group.

The average total level density of a two-dimensional bil- liard is given by Weyl’s asymptotic formula关22兴

¯g共e兲 ⬃ m

2␲ប2A+O共e−1/2兲, 共2.7兲 in which the leading-order term is proportional to the areaA surrounded by the boundary and does not depend on the energy. This means that for large energies ethe mean level spacing⌬¯ becomes asymptotically constant:

¯→¯

0=1 g

¯ ⬃2␲ប2

mA . 共2.8兲

As discussed in the previous subsection, the quantum levels of our system fall into four sets 兵en其 with ␬= 0⫾ and⫾1, corresponding to the eigenstates共2.5a兲,共2.5b兲, and共2.5c兲or 共2.5c⬘兲, respectively. The numbers of levels in these sets have the relative ratios

N0+:N0−:N+1:N−1= 1:1:2:2,

and hence the mean level spacing in each set is given by

¯0⫾兲= 6⌬¯

0, ⌬¯共⫾1= 3⌬¯

0. 共2.9兲

The unfolded level spacings are then obtained by sn共␬兲=en+1共␬兲en共␬兲

¯共␬兲 共␬= 0⫾,⫾1兲. 共2.10兲 Figure5shows the NNS distributionsPs兲, averaged over all four sets. At ␣= 0, the system is integrable and the distribu- tion is Poisson-like as expected. At ␣= 0.3, the distribution changes into Wigner form. A similar situation is also found at

0 50 100 150 200

0 0.2 0.4 0.6 0.8 1

e[h2/mR02]

α

(0,+) (0,−) (1,±)

FIG. 3. Lowest part of the energy spectrum 共in units of ប2/mR02兲, plotted against the deformation parameter␣.

-1 0 1 2

5 10 15 20 25 30

δE/N

√N α=1.0

-1 0 1 2

δE/N

α=0.5 -1

0 1 2

δE/N

α=0.0

FIG. 4. Scaled shell-correction energy ␦EN兲/冑N 共in units of ប2/mR02兲versus冑N.

(4)

␣= 0.7. At␣= 0.5, however, the distribution deviates consid- erably from the Wigner form and becomes closer to a Pois- son distribution. For mixed systems, the NNS distributions P共s兲are often fitted by a Brody distribution关23兴

B共s,␻兲=␤共␻+ 1兲sexp共−␤s␻+1兲, ␤=

+ 2+ 1

冊 册

␻+1,

共2.11兲 which interpolates between the Poisson 共␻= 0兲 and Wigner 共␻= 1兲 distributions. The Brody parameter ␻ can then be taken as a measure for the chaoticity of the NNS distribution.

In Fig.6we show␻as obtained by fitting theP共s兲distribu- tions of Fig. 5 to 共2.11兲. We clearly recognize two peaks around ␣⯝0.3 and ⯝0.7, exhibiting near-chaoticity, sepa- rated by a deep minimum around␣⯝0.5 where the system appears to approach regularity. As we will discuss below, this near-regularity is related to an approximate restoration of local dynamical symmetry due to a periodic-orbit bifurca- tion.

III. FOURIER ANALYSIS AND CLASSICAL PERIODIC ORBITS

In the POT, the quantum level density ge兲 is approxi- mated in terms of classical periodic orbits by the semiclassi- cal trace formula关1–6兴

g共e兲=

n

␦共e−en兲 ⯝¯g共e兲+

A共e兲cos

S共e兲−␲ 2␯

,

共3.1兲 where the first term, like for the energy in 共2.6兲, represents the smooth part, while the second term contains the quantum shell effects. In the latter, the sum is taken over all periodic orbits ␰ of the classical system 共or the orbit families ␰ in systems with continuous symmetries 关3–5兴兲,Sis the action integral around ␰, andis the Maslov index 关24,25兴. The amplitudeAdepends on the stability of the orbit␰共and, for an orbit family, on the phase-space volume covered by the family兲. For isolated orbits, the amplitudeA was given by Gutzwiller关1兴:

A共e兲= 1

␲ប

T共e兲

兩det关1 −M共e兲兴兩, 共3.2兲 whereT共e兲=dS共e兲/de is the period of the orbit andM共e兲 its stability matrix defined below. If an orbit has a discrete degeneracy f 共i.e., if there exist f replicas with identical ac- tions, stabilities, and Maslov indices, but different orienta- tions兲 due to discrete symmetries, it has to be included f times into the sum in共3.1兲. This holds also for time-reversed rotational orbits.

Transforming variables from energy eto wave number k and using the relationS=បkLfor billiards, whereLis the length of the orbit ␰, the trace formula becomes

g共k兲=ប2k

m g共e兲 ⯝¯g共k兲+

A共k兲cos

kL2␯

.

共3.3兲 Let us now consider the Fourier transform of the level den- sityg共k兲with respect tok:

F共L兲=

g共k兲e−ikLe−共k⌬兲2/2dk. 共3.4兲 The Gaussian damping factor is included for the truncation of the high-energy part of the spectrum. If we insert Eq.共3.3兲 and ignore the k dependence of the amplitude factors A 共which are constant for isolated orbits in billiards and cavi- ties兲, we obtain

Fsc共L兲=F0共L兲+␲

e−i␲␯/2A共L−L兲. 共3.5兲

x兲is a normalized Gaussian of width⌬, which turns into Dirac’s delta function in the limit ⌬0. Equation共3.5兲in- dicates that the Fourier transform F共L兲 is a function with successive peaks at the lengths of the periodic orbitsL=L, with heights proportional to the amplitudesA. We can there- fore extract information about the classical periodic orbits

0 0.2 0.4 0.6 0.8 1

0 1 2 3 4 5

P(s)

s

α=0.00

0 0.2 0.4 0.6 0.8 1

0 1 2 3 4 5

P(s)

s

α=0.30

0 0.2 0.4 0.6 0.8 1

0 1 2 3 4 5

P(s)

s

α=0.50

0 0.2 0.4 0.6 0.8 1

0 1 2 3 4 5

P(s)

s

α=0.70

FIG. 5. Nearest-neighbor spacing distributionPs兲for four val- ues of the deformation parameter␣. The lowest 600 levels共i.e., 100 and 200 levels of the subsets 0⫾ and⫾1, respectively兲were used to obtain the statistics. Solid and dashed lines show Poisson and Wigner distributions, respectively.

0 0.2 0.4 0.6 0.8 1

0 0.2 0.4 0.6 0.8 1

ω

α

FIG. 6. Brody parameter␻for nearest-neighbor spacing distri- bution, plotted as a function of␣.

(5)

from the Fourier transform of the quantum-mechanical level density:

Fqm共L兲=

n

e−iknL−共kn⌬兲2, kn=

2men/ប. 共3.6兲 Figure7shows the modulus of this Fourier transform of our system versus the deformation parameter␣. At= 0, where the periodic orbits form degenerate families corresponding to rational tori, we find the peaks at the well-known orbit lengths of the circular billiard关2兴:

Lvw= 2vR0sin共w␲/v兲. 共3.7兲 Herewandvare the winding number around the origin and the number of vertices 共vⱖ2w兲 of each orbit, respectively.

For instance, L= 4R0 for the diametric orbit 共w= 1 ,v= 2兲, L

= 3

3R0 for the triangular orbit 共w= 1 ,v= 3兲, L= 4

2R0 for the square orbit 共w= 1 ,v= 4兲, and so on. With increasing␣, we observe a dramatic enhancement of the peak height cor- responding to the triangular orbit, L⯝5.2R0 共and its second repetition L⯝10.4兲, starting around␣⯝0.4 and culminating around ␣⯝0.55. As we shall see below, this can be traced back to a codimension-2 bifurcation of the triangular orbit.

In fact, for␣⬎0 all rational tori of the circular billiard are broken into pairs of stable and unstable isolated orbits. With increasing␣, bifurcations of the stable orbits occur and new periodic orbits emerge, which makes the phase-space in- creasingly chaotic. In the following, we first demonstrate this for the two shortest orbits and then focus on the triangular orbit.

The stability of a periodic orbit is described by the stabil- ity matrix M, which is defined by the linearized Poincaré map around the periodic orbit:

M=⳵„q共T兲,p共T兲…

⳵„q共0兲,p共0兲… , 共3.8兲 where (q共t兲,p共t兲) are the coordinates and momenta perpen- dicular to the periodic orbit␰as functions of timet, andTis the period of the orbit. In a two-dimensional autonomous Hamiltonian system, Mis a symplectic 共2⫻2兲 matrix and the stability of a orbit is easily identified by looking at its

trace TrM. For elliptic共stable兲orbits, the eigenvalues ofM are of the form 共eiv,e−iv兲 with real vⱖ0, and thus −2 ⱕTrMⱕ2. For direct- and inverse-hyperbolic 共unstable兲 orbits, the eigenvalues are 共eu,e−u兲 and共−eu, −e−u兲, respec- tively, with realu⬎0, and hence TrM⬎2 and TrM⬍−2.

Bifurcations of isolated orbits occur whenever TrM= + 2.

Figures8and9 show TrM共␣兲for the two shortest pairs of periodic orbits in our system. In Fig. 8, the stable branch 共2A兲 of the diametric orbit is seen to undergo a period- doubling pitchfork bifurcation at␣= 0.166, where a symmet- ric wedge-shaped orbit共4V兲emerges. The latter undergoes a further pitchfork bifurcation at ␣= 0.227, where a pair of asymmetric wedge-shaped orbits共4U兲emerges.

2 4 6 8 10 12

0 0.2 0.4 0.6 0.8 1 0 2 4 6 8 10

L/R0 α

|F(L)| (arb. unit)

FIG. 7. Modulus兩FqmL兲兩of Fourier transform of the quantum- mechanical level density versus deformation parameter␣.

-6 -4 -2 0 2 4 6

0 0.1 0.2 0.3

TrM

α 2A

2B

2A2

4V 4U

FIG. 8. Trace of stability matrix TrM for the diameter orbits 共2A兲 and the second repetition of the stable diameter orbit共2A2兲, plotted versus deformation parameter␣.

-6 -4 -2 0 2 4 6

0 0.2 0.4 0.6 0.8 1

0.8 0.5 0.4 0.3 0.2 0.1 0.05

TrM

√α α

3A

3B 3C

3D

1.99 1.992 1.994 1.996 1.998 2 2.002 2.004 2.006 2.008 2.01

0.39 0.395 0.4 0.405 0.41 0.415 0.42

TrM

α 3A

3C

3D

FIG. 9. Same as Fig. 8, but for the triangular orbits, in the upper panel plotted versus 冑. The lower panel shows a magnifi- cation around the bifurcation of the stable orbit 3A occurring near

␣⯝0.407.

(6)

The bifurcation scenario of the triangular orbits is more complicated. In the upper panel of Fig. 9, where TrM is plotted against

, the stable triangular orbit 3A is seen to touch the critical line TrM= + 2 near

⯝0.63 共␣⯝0.4兲 and to remain stable on either side. A pair of new stable 3D and unstable 3C triangular orbits emerge from the touching point. A magnification of the situation around␣⯝0.4, shown in the lower panel against ␣, reveals that the scenario con- sists of two connected near-lying bifurcations, also called a

“codimension-2 bifurcation.” At␣= 0.400, a tangent共saddle- node兲bifurcation occurs, at which the new orbits 3D and 3C are born. Shortly after this bifurcation, the共old兲stable orbit 3A and the new unstable orbit 3C encounter in a touch- and-go bifurcation at␣= 0.407共see Appendix A for its ana- lytic value兲. Note that the new 3C and 3D orbits do not possessC3vsymmetry, in contrast to the old 3A orbit, so that they occur in degenerate triplets obtained by successive ro- tations about 2␲/3; i.e., these orbits have a discrete degen- eracy of f= 3.

Figure10shows excerpts of Poincaré surfaces of section 共␪,␸兲in the relevant regions. Here␪ is the polar angle of a reflection point of an orbit at the boundary, while ␸ repre- sents the reflection angle measured from the normal to the boundary at the reflection point. The three upper panels il- lustrate the tangent bifurcation of the 3C and 3D orbits. For

␣= 0.390共upper left兲, one sees one major regular island cor- responding to the stable equilateral triangular orbit 3A. It

contains its fixed point at 共␪,␸兲=共0 ,␲/6兲, surrounded by quasitori 共small aperiodic perturbations of the 3A orbit兲. At the bifurcation point ␣⬇0.400 共upper center兲, three cusps are formed by one of the surrounding quasitori in the island, and for ␣= 0.401共upper right兲, the three stable fixed points of the new 3D orbits, surrounded by small regular islands, are seen to have emerged from each of the three cusps in the major island. The three saddles separating these three islands from the central island contain the unstable fixed points of the 3C orbits. The stable fixed point of the 3A orbit still persists at the original position at the center of the major island. The numberf= 3 of the new stable and unstable fixed points in the island is due to the threefold discrete degen- eracy of the orbits 3D and 3C mentioned above.

The lower three panels in Fig. 10 illustrate the touch- and-go bifurcation of the orbits 3A and 3C. At the bifurcation point␣= 0.407共lower left兲, the three unstable fixed points of the 3C orbits have contracted into a starlike intersection of three quasitori, located at the central fixed point 共␪,␸兲

=共0 ,␲/6兲 of the 3A orbit. The three nearby stable fixed points in the major island belong to the 3D orbit. For ␣

= 0.420 共lower center兲, small stable islands have formed again around the fixed points of the 3A orbit, and nearby we recognize the three unstable fixed points of the 3C orbit. For

= 0.500共lower right兲, the central island of the 3A orbits has grown, the three islands of the stable 3D orbits have shrunk, and the three unstable fixed points of the 3C orbits are about to be buried in the increasing chaotic structure.

0 0.25 0.5 0.75 1

-60 -30 0 30 60

sinϕ

θ α=0.400

0 0.25 0.5 0.75 1

-60 -30 0 30 60

sinϕ

θ α=0.401

0 0.25 0.5 0.75 1

-60 -30 0 30 60

sinϕ

θ[deg]

α=0.390

0 0.25 0.5 0.75 1

-60 -30 0 30 60

sinϕ

θ α=0.407

0 0.25 0.5 0.75 1

-60 -30 0 30 60

sinϕ

θ α=0.420

0 0.25 0.5 0.75 1

-60 -30 0 30 60

sinϕ

θ α=0.500

FIG. 10. Poincaré surfaces of section共␪, sin␸兲in the fundamental domain of theC3group, plotted for several values of␣共given at the upper left of each panel兲in the bifurcation region of the triangular orbit 3A.␪is the polar angle of a reflection point and␸the reflection angle measured from the normal to the boundary; the set共␪, sin␸兲is approximately area preserving.

(7)

The bifurcation of the triangular orbit 3A occurs at the deformation where we observed the onset of the large Fou- rier peak in Fig.7. It is due to the appearance of the pair of sixfold-degenerate new 3D and 3C orbits. In particular, the stable one 3D, besides the doubly degenerate 3A orbit, adds to the local regularity of the phase space observed in Figs.5, 7, and 10. It is therefore apparent that this bifurcation is responsible for the remarkable shell structure in the quantum spectrum seen in Fig.3. Note that the maximum of the Fou- rier peak in Fig.7occurs at␣⯝0.55—i.e., above the bifur- cation where the new orbits are born. The fact that the sig- nature of a bifurcation is strongestafterthe bifurcation point 共on the side where the new orbits exist兲 is a rather general trend for shell effects which can also be understood in terms of semiclassical uniform approximations关9,10兴 共see the fol- lowing section and Fig.12below兲. It has also been observed recently in the level statistics of a Hamiltonian system with mixed phase space关26兴.

IV. SEMICLASSICAL ANALYSIS A. Semiclassical level density near bifurcations As mentioned above, the Gutzwiller trace formula 共3.1兲 with the amplitudes 共3.2兲 for isolated orbits diverges at bi- furcation points. This is due to the breakdown of the stationary-phase approximation, and higher-order terms in the expansion of the action integral must be included in the derivation of the trace formula. For this purpose, it is neces- sary 关8兴 to express the trace integral in phase space. After integrating over the coordinate and momentum parallel to an isolated orbit ␰in a two-dimensional system, its semiclassi- cal contribution to the oscillating part of the level density becomes 关4,10,16兴

ge兲= 1

2␲22Re

dq

⬘ 冕

dp1 nr

e

p2q

⬘ 冏

1/2

⫻exp

iSˆ共q

,p兲iq

pi2

. 共4.1兲

Here q

is the final coordinate and p the initial momentum perpendicular to the orbit␰, the two forming a canonical pair of variables共q

,p兲to describe the “natural” Poincaré surface of section 共PSS兲 of the orbit on which 共q

,p兲=共0 , 0兲 is its fixed point. For generic bifurcations,nris the repetition num- ber of the primitive orbit. The functionSˆ共q

,p兲denotes the Legendre transform ofS共q

,q兲:

Sˆ共q

,p兲=S共q

,q兲+q p, 共4.2兲 whereS共q

,q兲is the共open兲action integral along the orbit␰ 共at fixed energye兲,

S共r

,r兲=

r r

p共r

·dr

, 共4.3兲 projected onto the q axis. The function Sˆ共q

,p兲 is actually the generating function of the Poincaré map:

Sˆ共q

,p兲: 共q,p兲共q

,p

兲. 共4.4兲 If one expands the function ⌽共q

,p兲=Sˆ共q

,p兲q

p in the phase of the integrand of 共4.1兲 around q

=p= 0 up to qua- dratic terms inq

andp and evaluates the integrals in共4.1兲 by the standard SPA, one obtains the contribution of the orbit

␰ to Gutzwiller’s trace formula 共3.1兲 with the amplitude 共3.2兲. Near bifurcations of the orbit␰, the SPA breaks down and one has to include higher-order terms in the expansion of

⌽共q

,p兲. The minimum number of terms needed to describe a given bifurcation scenario on the PSS leads to the so-called

“normal forms” of ⌽共q

,p兲, which depend on the type of bifurcation. Doing the integrations in 共4.1兲 using such nor- mal forms leads to a finite combined contribution of the orbit

␰ and all the orbits involved in its bifurcation to the trace formula. In order to conform with standard notation, we re- name the function ⌽共q

,p兲 as S共q

,p兲 which, according to 共4.2兲, is identical with the projected action integral Sq

,q兲, but taken as a function of the variables q

andp.

B. Normal form for codimension-2 bifurcation For the description of the codimension-2 bifurcation sce- nario of the orbit 3A and its satellites 3C and 3D discussed in the previous section and illustrated on the PSS in Fig.10, the following normal form is appropriate关27兴:

SI,␸兲=S0共␣兲−⑀IaI3/2cos共3␸兲−bI2. 共4.5兲 Here one has transformed the Poincaré variables 共q

,p兲 to quasipolar variables 共I,␸兲by

p=

2Icos␸, q

=

2Isin␸. 共4.6兲 In共4.5兲,S0共␣兲is the共closed兲action integral of the central 3A orbit as a function of␣. The “bifurcation parameter”⑀ is a monotonously decreasing function of␣such that⑀= 0 at the touch-and-go bifurcation point 共here ␣=0= 0.407兲 of the central orbit and that⑀⬍0 for␣⬎␣0.aandbare parameters which are specific for the system and will be determined below.

That共4.5兲is able to describe the correct fixed-point struc- ture not only of the 3A orbit, but also of its satellites 3C and 3D, will now be shown explicitly. We first rewrite 共4.5兲 in terms ofq

andp:

S共q

,p兲=S0− ⑀

2共p2+q

2

a8共p3− 3pq

2 b

4共p2+q

22. 共4.7兲 The stationary-phase conditions are

qS

⬘ 冏

qi

= 0,

Sp

pi

= 0. 共4.8兲

One of the solutions is q0= 0 and p0= 0 and corresponds to the central 3A orbit. This is so by default, due to the choice of the Poincaré variables共q,p兲. Two further stationary points are found to be

(8)

q1,2= 0, p1,2= − 3a

4

2b⫾

32b9a22− ⑀

b. 共4.9兲

The fixed points 共qi,pi兲 with i= 1 , 2 belong to the satellite orbits 3C and 3D respectively. Two more pairs of fixed points for each of them are found by rotations in the 共I,␸兲 plane: ␸␸⫾2␲/3. ⑀1= 9a2/32b is the critical value for p1,2to have real values—i.e., for the real orbits 3C and 3D to exist. For our system we can chooseb⬎0, and therefore⑀1

⬎0. For⑀⬎⑀1, the 3C and 3D orbits become imaginary and only the central orbit 3Ais real. With decreasing⑀, the three pairs of stable and unstable satellite orbits appear at ⑀=1, which we identify with the tangent bifurcation point. At ⑀

= 0 we have the touch-and-go bifurcation as stated above.

For⑀⬍0共␣⬎␣0兲, all orbits are real.

The normal-form parameters ⑀, a, and b can be deter- mined by fitting the actionsSand stability traces TrM of the orbits, which have been obtained numerically, to their local behaviors predicted by the normal form 共4.5兲. The stability trace is given in terms of by关10兴

TrM=

p2q

⬘ 冊

−1

1 +

p2q

⬘ 冊

22p2

2

q

2

.

共4.10兲 For the central orbit, it becomes

TrM0= TrMA= 2 −⑀2, 共4.11兲 so that␧can be determined as⑀=

2 − TrMA, choosing the correct sign on either side of the bifurcation. The other pa- rameters are obtained from the action difference of the sat- ellite orbits 1 and 2 共i.e., 3C and 3D兲. Inserting 共q1,2,p1,2兲 from共4.9兲into共4.7兲, one finds

SDSC

បkR0

= 4

3b

1共⑀−⑀13/2. 共4.12兲 By fitting the numerical data forSDSCas function of⑀, we obtain ⑀1 andb, and thereforea. Thus we can uniquely de- termine all the normal form parameters. The results for the triangular orbits discussed above are

a=0.519252

kR0 , b=2.34950 បkR0

, ⑀1= 0.0322755.

共4.13兲 The formulas for the stability traces for the 3C and 3D orbits are

TrMC,D= 2⫾12

13共⑀1−⑀兲− 24⑀1共⑀1−⑀兲

⫾12

1共⑀1−⑀兲3, 共4.14兲 and the action difference between the 3C and 3A orbits is

SCSA

បkR0

= 1

12b关3⑀2− 121+ 812− 8

1共⑀1−⑀兲3兴. 共4.15兲 We have checked that the numerical results in the neighbor- hood of the bifurcations are nicely reproduced by these equa- tions.

C. Uniform approximations

Inserting the normal form共4.5兲into the integral共4.1兲, one obtains a “local” uniform approximation 关8兴 which is finite and valid near the bifurcation—i.e., for not too large absolute values of ⑀. Due to the C3v symmetry of our system, the touch-and-go bifurcation is nongeneric and isochronous. The factor nr in the denominator of 共4.1兲 here must be chosen 关28兴 asnr=f= 3. We have another degeneracy factor of 2 in the numerator due to the time-reversal symmetry of all or- bits. Using the variables共I,␸兲and changing the energyeto the wave number k, we finally obtain the following expres- sion for the local uniform approximation, in which the inte- gration over␸ can be done analytically:

g共k兲= 1 3␲2បRe

0

dI

0 2␲

dL

I2

1/2exp

i兵S0I

−⑀IaI3/2cos共3␸兲−bI2其−i␲ 2␯

= 2L

3␲បReeikL−i␲␯/2

0

dI J0

aI3/2

ei/ប兲共−⑀I−bI2.

共4.16兲 The integration over I can be performed numerically using an expansion formula given in关27兴.

As stated above, the result 共4.16兲 is only useful in the neighborhood of the bifurcation. Far away from it, where all orbits involved become isolated, it can be evaluated asymp- totically 共corresponding to the SPA兲, but the amplitudes of the orbits then do not agree with the Gutzwiller values共3.2兲.

In order to achieve this, one must develop “global” uniform approximations关9,10兴. To that purpose, one needs to include higher-order expansion terms in the normal form. For the codimension-2 bifurcation of our type one obtains, after suit- able coordinate transformations关29兴,

gk兲= 1 3␲2បRe

0

dI

0 2␲

d␸⌿共I,␸兲exp

iS0II

aI3/2cos共3␸兲−bI2其−i

2 ␯

, 共4.17兲

with

⌿共I,␸兲=L−␣II3/2cos共3␸兲, 共4.18兲 where ␣ and ␤ are expressed by a certain combination of higher-order expansion coefficients. In practice, these param- eters are determined such that共4.17兲yields the sum of con- tributions with amplitudes 共3.2兲 of all involved isolated or-

(9)

bits far away from the bifurcation point. They are determined by equating 共cf.关30兴兲

LC

兩2 − TrMC兩=LA−␣IC−␤IC3/2

兩det⌽

共IC兲兩 , 共4.19兲 LD

兩2 − TrMD兩=LA−␣ID−␤ID3/2

兩det⌽

ID兲兩 , 共4.20兲 with

det⌽

=

⳵␸I22SS2 ⳵␸22ISS2I

␸=0=

pq22S

S2q

q

22pSS2p

q=0

共4.21兲 and

IC,D=1

2共q1,22 +p1,22 兲. 共4.22兲 We now have all ingredients ready for calculating the semi- classical level density of our system. Since it is well known 关1兴that the sum over all periodic orbits does not converge in systems with mixed dynamics, we have to introduce a certain truncation. This is achieved关7,31兴by focusing on the gross- shell structure of the level density, coarse-graining it by a Gaussian convolution overkwith width␥:

gk兲= 1

␲␥

dk

e−关共kk兲/␥兴2gk

. 4.23

The semiclassical trace formula共3.3兲then becomes g共k兲 ⯝¯g共k兲+

e−共␥L/2兲2A共k兲cos

kL2␯

.

共4.24兲 Due to the Gaussian factor, periodic orbits with lengths L ⲏ2␲/␥ are now exponentially suppressed. Choosing ␥

= 0.6, we need only consider periodic orbits withL⬍10R0. From Fig. 8, we see that no isochronous bifurcations of the diametric orbit occur for␣⬎0, so that its contribution can be evaluated by the standard Gutzwiller formula with ampli- tudes共3.2兲. A period-doubling bifurcation of the stable diam-

eter occurs at␣⯝0.15, but does not contribute much to the coarse-grained level density. The same is true for the tetrag- onal and pentagonal orbits. The contributions of the triangu- lar orbit 3A and its satellites 3C and 3D are included in the global uniform approximation共4.17兲, including the Gaussian damping factor exp关−共␥LA/2兲2兴.

The hexagonal orbit encounters also a codimension-2 bi- furcation of the same type as the triangular orbit, as seen in Fig. 11. Determining the normal-form parameters in the same manner as above, we find the values

a=14.0046

បkR0

, b=414.669 បkR0

, 共4.25兲

which are much larger than those for the triangular orbit given in 共4.13兲. Since they contribute inversely to the local uniform approximation 共4.16兲, this bifurcation has a much less dramatic influence on the level density than that of the triangular orbit. In order to demonstrate this more explicitly, we plot in Figs.12and13the modulus of the integral

F共⑀;a,b兲=

0

dI J0共aI3/2兲ei−⑀I−bI2 共4.26兲 appearing in共4.16兲. Figure12is for the parameters共4.13兲of the triangular and Fig. 13 for the parameters 共4.25兲 of the hexagonal orbits. We see that for largeraandb, this integral has smaller values and is a more monotonous function of␧.

For smallaandb, however, it takes large values and exhibits a considerable peak near the bifurcation point ⑀= 0. 共Note that the maximum actually occurs slightly after the bifurcation—i.e., for −⑀⬎0—as discussed at the end of Sec.

III.兲 This can be understood as follows. If the normal-form parametersa andb are small, the actionS does not depend

1.9 1.95 2 2.05 2.1

0.35 0.355 0.36 0.365 0.37

TrM

x 6A 6C

6D

FIG. 11. Same as Fig.9for the hexagonal orbits.

0.2 0.4 0.6 0.8 1.0

~a

-2 -1 0 1 2

0

1 2 3

|F(;a,b)|

2 3 4 5

b~

-2 -1 0 1 2

0

1 2 3

|F(;a,b)|

FIG. 12. Profile of the functionF共⑀;a,b兲in共4.26兲for the nor- mal form parameters共4.13兲of the triangular orbit. The left panel is for fixedband varying˜a=a冑kR0and the right panel for fixeda and varying˜b=bបkR0, both forkR0= 20.

10 20 30

~a

-2 -1 0 1 2

0

0.1 0.2 0.3

|F(;a,b)|

200 400 600

b~

-2 -1 0 1 2

0

0.1 0.2 0.3

|F(;a,b)|

FIG. 13. Same as Fig.12, but with the normal form parameters 共4.25兲of the hexagonal orbit.

(10)

much on I and ␸, and the quasiperiodic orbits around the central orbit give a more coherent contribution to the inte- gral, yielding a better restoration of local dynamical symme- try around the central periodic orbit. Actually, the limit a

0 is sufficient to locally restore integrability: the normal form共4.5兲then only depends onIand forb⫽0 describes the bifurcation of a torus from an isolated orbit共cf.关32兴兲.

This local dynamical symmetry is equivalent to a reso- nance condition for quasitori winding around the bifurcating periodic orbit over a rather wide region. The quantization of these quasitori yields a number of quasidegenerate levels, seen as the bunches in the spectrum of Fig.3. This increases the probability for small level spacings and renders the NNS distribution more Poisson-like. The considerable changes in the NNS distributions shown in Figs. 5 and 6 around ␣

= 0.5 therefore account for the emergence of these quasireso- nant tori.

In Fig.14we compare the oscillating part of the semiclas- sical level density with that of the quantum-mechanical re- sult. In the semiclassical calculation we have included the primitive diametric, triangular, tetragonal, pentagonal, and hexagonal orbits as explained above. We see that, besides the rapid oscillations due to the average length of the shortest orbits included共i.e., mainly of the short diameter orbit兲, there is a beating pattern that comes from the interferences be- tween the different orbits. We note that both amplitudes and phases are nicely reproduced in the semiclassical result for all deformation parameters including the bifurcation points

␣⯝0.4. For ␣= 0.5 and 0.6, the interference effects are re- duced and the oscillating pattern is quite regular, indicating the dominance of the bifurcating periodic orbits and the ac- companying quasitori which have approximately the same lengths. This is also seen in the shell-correction energy for

␣= 0.5 in Fig.4, where the spacing⌬Nof the regular oscil- lation corresponds to the length of the dominating triangular orbit.

V. DISCUSSION OF DISCRETE SYMMETRIES We conclude by some remarks on the role of discrete symmetries. In the above model, the triangular periodic orbit undergoes a bifurcation in the course of which a pair of threefold-degenerate new orbits are created. 共We only men- tion the relative degeneracies here; all orbits have two extra degeneracy factors of 2 due to reflections at thexaxis and to time reversal.兲Stated more generally: the bifurcation brings about new periodic orbits of reduced symmetry, which have several degenerate replicas, connected with the symmetry operations RandP in共2.3兲, which give coherent contribu- tions to the level density. This is one of the reasons why we obtain a strong shell effect due to this bifurcation. If the C3 symmetry is slightly broken, the equilateral triangular orbit will split into three nonequilateral triangular orbits with dif- ferent lengths and different stabilities. They will give de- structive contributions to the level density in most of the parameter region. We would therefore expect that in billiards with Cnv symmetry, the regular polygonal orbits withn re- flections will play the most prominent role, such as the tri- angular orbit in the present system.

This behavior can also be predicted from a perturbative trace formula, developed by Creagh 关33兴, which describes semiclassically the breaking of continuous symmetries. In systems with continuous symmetries, the leading periodic orbits occur in degenerate families corresponding to rational tori. When a perturbation breaks the continuous symmetries, the tori are broken into isolated orbits. For weak perturba- tions, only those orbit families 共tori兲␰are broken for which the action change⌬Sin lowest order of the classical pertur- bation theory is nonzero. Furthermore, if the perturbed sys- tem still has a discrete point symmetry, the orbit families which have the same point symmetry are in resonance with the perturbation and typically suffer the largest first-order action change. In our billiard system 共2.2兲, we can treat the deviation from the circular billiard, for sufficiently small val- ues of ␣, as a perturbation that breaks the continuous U共1兲 symmetry. For small␣, the shape of the boundary is given by R共␪兲 ⬇R0共1 −⑀cos 3␪兲, 共5.1兲 and ␧=

3␣/9 is the appropriate perturbation parameter. As we show in Appendix B, the length of the periodic orbit family 共v,w兲 is changed in first order of ⑀ only forv= 3w;

i.e., only triangular orbits are affected in first-order perturba- tion theory. The situation is similar in a spherical cavity per- turbed by deformations with 2j-pole deformations. For this system it was shown explicitly 关34兴 that the orbit families

-2 -1 0 1 2

δg(k)

α=0.3,γR0=0.6

-2 -1 0 1 2

δg(k)

α=0.4

-2 -1 0 1 2

δg(k)

α=0.41

-2 -1 0 1 2

δg(k)

α=0.5

-2 -1 0 1 2

0 10 20 30 40

δg(k)

kR0 α=0.6

FIG. 14. Oscillating part of the coarse-grained level density.

Solid and dashed lines represent the semiclassical and quantum- mechanical results, respectively.

Referenzen

ÄHNLICHE DOKUMENTE

The new generation of appropriators claims to take artistic concepts such as novelty, originality, authorship, ingenuity, intentionality, creativity, expression,

Previous research indicates that benefits of sprouting may be negated by net DM loss from sprouting coupled with no significant improvement in nutrient concentrations or

For different starting conditions and container geometries, we observe an interesting interplay between the moving swimmers and the elevated liquid surface, which results in

binding of a low molecular weight rhodium complex to a poly- electrolyte layer assembled on the surface of a latex particle, and application of the resulting materials

After representing first the generating function in its full form, we examine in the next section the be- haviour of this function in the special (zero tempera- ture) case w = 0

A previous theory of integer-valued probability distributions is extended to many dimensions, to develop a really systematic way of treating mode mixing.. The mixing of the

First conceived as one of the major elements in the Alliance’s transformation in the early years of the 21st century, it was supposed to bring a rapid military response to emerging

EU accession needs transformations different from the earlier ones in a way that they will be less autonomous, demand active government control and reliance on bureaucracies, will