ArticlePDF Available

Experimental investigation of nodal domains in the chaotic microwave rough billiard

Authors:

Abstract and Figures

We present the results of experimental study of nodal domains of wave functions (electric field distributions) lying in the regime of Shnirelman ergodicity in the chaotic half-circular microwave rough billiard. Nodal domains are regions where a wave function has a definite sign. The wave functions PsiN of the rough billiard were measured up to the level number N=435 . In this way the dependence of the number of nodal domains [symbol: see text]N on the level number N was found. We show that in the limit N-->infinity a least squares fit of the experimental data reveals the asymptotic number of nodal domains [symbol: see text]N/N approximately equal to 0.058+/-0.006 that is close to the theoretical prediction [symbol: see text]N/N approximately equal to 0.062 . We also found that the distributions of the areas s of nodal domains and their perimeters l have power behaviors ns is proportional to s(-tau) and nl is proportional to l(-tau'), where scaling exponents are equal to tau=1.99+/-0.14 and tau'=2.13+/-0.23 , respectively. These results are in a good agreement with the predictions of percolation theory. Finally, we demonstrate that for higher level numbers N approximately equal to 220-435 the signed area distribution oscillates around the theoretical limit SigmaA approximately 0.0386 N(-1) .
Content may be subject to copyright.
arXiv:0903.1965v1 [nlin.CD] 11 Mar 2009
Experimental investigation of nodal domains in the chaotic
microwave rough billiard
Nazar Savytskyy, Oleh Hul and Leszek Sirko
Institute of Physics, Polish Academy of Sciences,
Aleja Lotnik´ow 32/46, 02-668 Warszawa, Poland
(Dated: August 26, 2004)
Abstract
We present the results of experimental study of nodal domains of wave functions (electric field
distributions) lying in the regime of Shnirelman ergodicity in the chaotic half-circular microwave
rough billiard. Nodal domains are regions where a wave function has a definite sign. The wave
functions ΨNof the rough billiard were measured up to the level number N= 435. In this way the
dependence of the number of nodal domains Non the level number Nwas found. We show that
in the limit N a least squares fit of the experimental data reveals the asymptotic number of
nodal domains N/N 0.058 ±0.006 that is close to the theoretical prediction N/N 0.062.
We also found that the distributions of the areas sof nodal domains and their perimeters lhave
power behaviors nssτand nllτ, where scaling exponents are equal to τ= 1.99 ±0.14
and τ= 2.13 ±0.23, respectively. These results are in a good agreement with the predictions
of percolation theory. Finally, we demonstrate that for higher level numbers N220 435 the
signed area distribution oscillates around the theoretical limit ΣA0.0386N1.
1
In recent papers Blum et al. [1] and Bogomolny and Schmit [2] have considered the
distribution of the nodal domains of real wave functions Ψ(x, y) in 2D quantum systems
(billiards). The condition Ψ(x, y) = 0 determines a set of nodal lines which separate regions
(nodal domains) where a wave function Ψ(x, y) has opposite signs. Blum et al. [1] have
shown that the distributions of the number of nodal domains can be used to distinguish
between systems with integrable and chaotic underlying classical dynamics. In this way
they provided a new criterion of quantum chaos, which is not directly related to spectral
statistics. Bogomolny and Schmit [2] have shown that the distribution of nodal domains
for quantum wave functions of chaotic systems is universal. In order to prove it they have
proposed a very fruitful, percolationlike, model for description of properties of the nodal
domains of generic chaotic system. In particular, the model predicts that the distribution of
the areas sof nodal domains should have power behavior nssτ, where τ= 187/91 [3].
In this paper we present the first experimental investigation of nodal domains of wave
functions of the chaotic microwave rough billiard. We tested experimentally some of impor-
tant findings of papers by Blum et al. [1] and Bogomolny and Schmit [2] such as the signed
area distribution ΣAor the dependence of the number of nodal domains Non the level
number N. Additionally, we checked the power dependence of nodal domain perimeters l,
nllτ, where according to percolation theory the scaling exponent τ= 15/7 [3], which
was not considered in the above papers.
In the experiment we used the thin (height h= 8 mm) aluminium cavity in the shape of
a rough half-circle (Fig. 1). The microwave cavity simulates the rough quantum billiard due
to the equivalence between the Schr¨odinger equation and the Helmholtz equation [4, 5]. This
equivalence remains valid for frequencies less than the cut-off frequency νc=c/2h18.7
GHz, where c is the speed of light. The cavity sidewalls are made of 2 segments. The rough
segment 1 is described by the radius function R(θ) = R0+PM
m=2 amsin( +φm), where
the mean radius R0=20.0 cm, M= 20, amand φmare uniformly distributed on [0.269,0.297]
cm and [0,2π], respectively, and 0 θ < π. It is worth noting that following our earlier
experience [6, 7] we decided to use a rough half-circular cavity instead of a rough circular
cavity because in this way we avoided nearly degenerate low-level eigenvalues, which could
not be possible distinguished in the measurements. As we will see below, a half-circular
geometry of the cavity was also very suitable in the accurate measurements of the electric
field distributions inside the billiard.
2
FIG. 1: Sketch of the chaotic half-circular microwave rough billiard in the xy plane. Dimensions
are given in cm. The cavity sidewalls are marked by 1 and 2 (see text). Squared wave functions
|ΨN(Rc, θ)|2were evaluated on a half-circle of fixed radius Rc= 17.5 cm. Billiard’s rough boundary
Γ is marked with the bold line.
The surface roughness of a billiard is characterized by the function k(θ) = (dR/dθ)/R0.
Thus for our billiard we have the angle average ˜
k= (hk2(θ)iθ)1/20.488. In such a
billiard the dynamics is diffusive in orbital momentum due to collisions with the rough
boundary because ˜
kis much above the chaos border kc=M5/2= 0.00056 [8]. The
roughness parameter ˜
kdetermines also other properties of the billiard [9]. The eigenstates
are localized for the level number N < Ne= 1/128˜
k4. Because of a large value of the
roughness parameter ˜
kthe localization border lies very low, Ne1. The border of Breit-
Wigner regime is NW=M2/48˜
k235. It means that between Ne< N < NWWigner
ergodicity [9] ought to be observed and for N > NWShnirelman ergodicity should emerge.
In 1974 Shnirelman [10] proved that quantum states in chaotic billiards become ergodic for
sufficiently high level numbers. This means that for high level numbers wave functions have
to be uniformly spread out in the billiards. Frahm and Shepelyansky [9] showed that in the
rough billiards the transition from the exponentially localized states to the ergodic ones is
more complicated and can pass through an intermediate regime of Wigner ergodicity. In
this regime the wave functions are nonergodic and compose of rare strong peaks distributed
3
over the whole energy surface. In the regime of Shnirelman ergodicity the wave functions
should be distributed homogeneously on the energy surface.
In this paper we focus our attention on Shnirelman ergodicity regime.
One should mention that rough billiards and related systems are of considerable interest
elsewhere, e.g. in the context of dynamic localization [11], localization in discontinuous quan-
tum systems [12], microdisc lasers [13, 14] and ballistic electron transport in microstructures
[15].
In order to investigate properties of nodal domains knowledge of wave functions (elec-
tric field distributions inside the microwave billiard) is indispensable. To measure the wave
functions we used a new, very effective method described in [16]. It is based on the pertur-
bation technique and preparation of the “trial functions”. Below we will describe shortly
this method.
The wave functions ΨN(r, θ) (electric field distribution EN(r, θ) inside the cavity) can be
determined from the form of electric field EN(Rc, θ) evaluated on a half-circle of fixed radius
Rc(see Fig. 1). The first step in evaluation of EN(Rc, θ) is measurement of |EN(Rc, θ)|2.
The perturbation technique developed in [17] and used successfully in [17, 18, 19, 20] was
implemented for this purpose. In this method a small perturber is introduced inside the
cavity to alter its resonant frequency according to
ννN=νN(aB2
NbE2
N),(1)
where νNis the Nth resonant frequency of the unperturbed cavity, aand bare geometrical
factors. Equation (1) shows that the formula can not be used to evaluate E2
Nuntil the term
containing magnetic field BNvanishes. To minimize the influence of BNon the frequency
shift ννNa small piece of a metallic pin (3.0 mm in length and 0.25 mm in diameter) was
used as a perturber. The perturber was moved by the stepper motor via the Kevlar line
hidden in the groove (0.4 mm wide, 1.0 mm deep) made in the cavity’s bottom wall along the
half-circle Rc. Using such a perturber we had no positive frequency shifts that would exceed
the uncertainty of frequency shift measurements (15 kHz). We checked that the presence
of the narrow groove in the bottom wall of the cavity caused only very small changes δνN
of the eigenfrequencies νNof the cavity |δνN|N104. Therefore, its influence into the
structure of the cavity’s wave functions was also negligible. A big advantage of using hidden
in the groove line was connected with the fact that the attached to the line perturber was
4
FIG. 2: Panel (a): Squared wave function |Ψ435(Rc, θ)|2(in arbitrary units) measured on a half-
circle with radius Rc= 17.5 cm (ν435 14.44 GHz). Panel (b): The “trial wave function”
Ψ435(Rc, θ) (in arbitrary units) with the correctly assigned signs, which was used in the recon-
struction of the wave function Ψ435(r, θ) of the billiard (see Fig. 3).
always vertically positioned what is crucial in the measurements of the square of electric field
EN. To eliminate the variation of resonant frequency connected with the thermal expansion
of the aluminium cavity the temperature of the cavity was stabilized with the accuracy of
0.05 deg.
The regime of Shnirelman ergodicity for the experimental rough billiard is defined for N >
35. Using a field perturbation technique we measured squared wave functions |ΨN(Rc, θ)|2
for 156 modes within the region 80 N435. The range of corresponding eigenfrequencies
was from ν80 6.44 GHz to ν435 14.44 GHz. The measurements were performed at 0.36
mm steps along a half-circle with fixed radius Rc= 17.5 cm. This step was small enough to
reveal in details the space structure of high-lying levels. In Fig. 2 (a) we show the example
of the squared wave function |ΨN(Rc, θ)|2evaluated for the level number N= 435. The
5
perturbation method used in our measurements allows us to extract information about the
wave function amplitude |ΨN(Rc, θ)|at any given point of the cavity but it doesn’t allow
to determine the sign of ΨN(Rc, θ) [21]. Our results presented in [16] suggest the following
sign-assignment strategy: We begin with the identification of all close to zero minima of
|ΨN(Rc, θ)|. Then the sign “minus” maybe arbitrarily assigned to the region between the
first and the second minimum, “plus” to the region between the second minimum and the
third one, the next minus” to the next region between consecutive minima and so on.
In this way we construct our “trial wave function” ΨN(Rc, θ). If the assignment of the
signs is correct we should reconstruct the wave function ΨN(r, θ) inside the billiard with the
boundary condition ΨN(rΓ, θΓ) = 0.
The wave functions of a rough half-circular billiard may be expanded in terms of circular
waves (here only odd states in expansion are considered)
ΨN(r, θ) =
L
X
s=1
asCsJs(kNr) sin(),(2)
where Cs= (π
2Rrmax
0|Js(kNr)|2rdr)1/2and kN= 2πνN/c.
In Eq. (2) the number of basis functions is limited to L=kNrmax =lmax
N, where
rmax = 21.4 cm is the maximum radius of the cavity. lmax
N=kNrmax is a semiclassical
estimate for the maximum possible angular momentum for a given kN. Circular waves with
angular momentum s > L correspond to evanescent waves and can be neglected. Coefficients
asmay be extracted from the “trial wave function” ΨN(Rc, θ) via
as= [π
2CsJs(kNRc)]1Zπ
0ΨN(Rc, θ) sin(). (3)
Since our “trial wave function” ΨN(Rc, θ) is only defined on a half-circle of fixed radius Rc
and is not normalized we imposed normalization of the coefficients as:PL
s=1 |as|2= 1. Now,
the coefficients asand Eq. (2) can be used to reconstruct the wave function ΨN(r, θ) of the
billiard. Due to experimental uncertainties and the finite step size in the measurements of
|ΨN(Rc, θ)|2the wave functions ΨN(r, θ) are not exactly zero at the boundary Γ. As the
quantitative measure of the sign assignment quality we chose the integral γRΓ|ΨN(r, θ)|2dl
calculated along the billiard’s rough boundary Γ, where γis length of Γ. In Fig. 2 (b)
we show the “trial wave function” Ψ435(Rc, θ) with the correctly assigned signs, which was
used in the reconstruction of the wave function Ψ435(r, θ) of the billiard (see Fig. 3). Using
the method of the “trial wave function” we were able to reconstruct 138 experimental wave
6
FIG. 3: The reconstructed wave function Ψ435(r, θ) of the chaotic half-circular microwave rough
billiard. The amplitudes have been converted into a grey scale with white corresponding to large
positive and black corresponding to large negative values, respectively. Dimensions of the billiard
are given in cm.
functions of the rough half-circular billiard with the level number Nbetween 80 and 248
and 18 wave functions with Nbetween 250 and 435. The wave functions were reconstructed
on points of a square grid of side 4.3·104m. The remaining wave functions from the
range N= 80 435 were not reconstructed because of the accidental near-degeneration
of the neighboring states or due to the problems with the measurements of |ΨN(Rc, θ)|2
along a half-circle coinciding for its significant part with one of the nodal lines of ΨN(r, θ).
These problems are getting much more severe for N > 250. Furthermore, the computation
time trrequired for reconstruction of the ”trial wave function” scales like tr2nz2, where
nzis the number of identified zeros in the measured function |ΨN(Rc, θ)|. For higher N,
the computation time tron a standard personal computer with the processor AMD Athlon
XP 1800+ often exceeds several hours, what significantly slows down the reconstruction
procedure.
Ergodicity of the billiard’s wave functions can be checked by finding the structure of the
energy surface [8]. For this reason we extracted wave function amplitudes C(N)
nl =hn, l|Niin
the basis n, l of a half-circular billiard with radius rmax , where n= 1,2,3...enumerates the
7
FIG. 4: Structure of the energy surface in the regime of Shnirelman ergodicity. Here we show the
moduli of amplitudes |C(N)
nl |for the wave functions: (a) N= 86, (b) N= 435. The wave functions
are delocalized in the n, l basis. Full lines show the semiclassical estimation of the energy surface
(see text).
zeros of the Bessel functions and l= 1,2,3... is the angular quantum number. The moduli
of amplitudes |C(N)
nl |and their projections into the energy surface for the representative
experimental wave functions N= 86 and N= 435 are shown in Fig. 4. As expected, in the
regime of Shnirelman ergodicity the wave functions are extended homogeneously over the
whole energy surface [6]. The full lines on the projection planes in Fig. 4(a) and Fig. 4(b)
mark the energy surface of a half-circular billiard H(n, l) = EN=k2
Nestimated from the
semiclassical formula [7]: q(lmax
N)2l2larctan(l1q(lmax
N)2l2) + π/4 = πn. The peaks
|C(N)
nl |are spread almost perfectly along the lines marking the energy surface.
An additional confirmation of ergodic behavior of the measured wave functions can be also
8
FIG. 5: Amplitude distribution PA1/2) for the eigenstates: (a) N= 86 and (b) N= 435 con-
structed as histograms with bin equal to 0.2. The width of the distribution P(Ψ) was rescaled
to unity by multiplying normalized to unity wave function by the factor A1/2, where Ade-
notes billiard’s area. Full line shows standard normalized Gaussian prediction P0A1/2) =
(1/2π)eΨ2A/2.
sought in the form of the amplitude distribution P(Ψ) [22, 23]. For irregular, chaotic states
the probability of finding the value Ψ at any point inside the billiard, without knowledge
of the surrounding values, should be distributed as a Gaussian, P(Ψ) eβΨ2. It is worth
noting that in the above case the spatial intensity should be distributed according to Porter-
Thomas statistics [5]. The amplitude distributions PA1/2) for the wave functions N= 86
and N= 435 are shown in Fig. 5. They were constructed as normalized to unity histograms
with the bin equal to 0.2. The width of the amplitude distributions P(Ψ) was rescaled to
unity by multiplying normalized to unity wave functions by the factor A1/2, where Adenotes
billiard’s area (see formula (23) in [22]). For all measured wave functions in the regime of
Shnirelman ergodicity there is a good agreement with the standard normalized Gaussian
9
FIG. 6: The number of nodal domains N(full circles) for the chaotic half-circular microwave
rough billiard. Full line shows a least squares fit N=a1N+b1Nto the experimental data (see
text), where a1= 0.058 ±0.006, b1= 1.075 ±0.088. The prediction of the theory of Bogomolny
and Schmit [2] a1= 0.062.
prediction P0A1/2) = (1/2π)eΨ2A/2.
The number of nodal domains Nvs. the level number Nin the chaotic microwave rough
billiard is plotted in Fig. 6. The full line in Fig. 6 shows a least squares fit N=a1N+b1N
of the experimental data, where a1= 0.058 ±0.006, b1= 1.075 ±0.088. The coefficient
a1= 0.058 ±0.006 coincides with the prediction of the percolation model of Bogomolny and
Schmit [2] N/N 0.062 within the error limits. The second term in a least squares fit
corresponds to a contribution of boundary domains, i.e. domains, which include the billiard
boundary. Numerical calculations of Blum et al. [1] performed for the Sinai and stadium
billiards showed that the number of boundary domains scales as the number of the boundary
intersections, that is as N. Our results clearly suggest that in the rough billiard, at low
level number N, the boundary domains also significantly influence the scaling of the number
of nodal domains N, leading to the departure from the predicted scaling NN.
The bond percolation model [2] at the critical point pc= 1/2 allows us to apply other
results of percolation theory to the description of nodal domains of chaotic billiards. In par-
ticular, percolation theory predicts that the distributions of the areas sand the perimeters
10
FIG. 7: Distribution of nodal domain areas. Full line shows the prediction of percolation theory
log10(hns/ni) = 187
91 log10(hs/smin i). A least squares fit log10(hns/ni) = a2τlog10(hs/smini) of
the experimental results lying within the vertical lines yields the scaling exponent τ= 1.99 ±0.14
and a2=0.05 ±0.04. The result of the fit is shown by the dashed line.
lof nodal clusters should obey the scaling behaviors: nssτand nllτ, respec-
tively. The scaling exponents [3] are found to be τ= 187/91 and τ= 15/7. In Fig. 7
we present in logarithmic scales nodal domain areas distribution hns/nivs. hs/sminiob-
tained for the microwave rough billiard. The distribution hns/niwas constructed as nor-
malized to unity histogram with the bin equal to 1. The areas sof nodal domains were
calculated by summing up the areas of the nearest neighboring grid sites having the same
sign of the wave function. In Fig. 7 the vertical axis hns/ni=1
NTPNT
i=1 n(N)
s/n(N)repre-
sents the number of nodal domains n(N)
sof size sdivided by the total number of domains
n(N)averaged over NT= 18 wave functions measured in the range 250 N435. In
these calculations we used only the highest measured wave functions in order to minimize
the influence of boundary domains on nodal domain areas distribution. Following Bogo-
molny and Schmit [2], the horizontal axis is expressed in the units of the smallest possi-
ble area s(N)
min,hs/smini=1
NTPNT
i=1 s/s(N)
min, where s(N)
min =π(j01/kN)2and j01 2.4048 is
the first zero of the Bessel function J0(j01) = 0. The full line in Fig. 7 shows the pre-
diction of percolation theory log10(hns/ni) = 187
91 log10(hs/smin i). In a broad range of
11
FIG. 8: Distribution of nodal domain perimeters. Full line shows the prediction of percolation
theory log10(hnl/ni) = 15
7log10(hl/lmini). A least squares fit log10(hnl/ni) = a3τlog10 (hl/lmin i)
of the experimental results lying within the range marked by the vertical lines yields τ= 2.13±0.23
and a3= 0.04 ±0.21. The result of the fit is shown by the dashed line.
log10(hs/smin i), approximately from 0.2 to 1.3, which is marked by the two vertical lines
the experimental results follow closely the theoretical prediction. Indeed, a least squares fit
log10(hns/ni) = a2τlog10(hs/smini) of the experimental results lying within the vertical
lines yields the scaling exponent τ= 1.99 ±0.14 and a2=0.05 ±0.04, which is in a good
agreement with the predicted τ= 187/91 2.05. The dashed line in Fig. 7 shows the
results of the fit. In the vicinity of log10(hs/smini)1 and 1.2 small excesses of large areas
are visible. A similar situation, but for larger log10(s/smin)>4, can be also observed in the
nodal domain areas distribution presented in Fig. 5 in Ref. [2] for the random wave model.
The exact cause of this behavior is not known but we can possible link it with the limited
number of wave functions used for the preparation of the distribution.
Nodal domain perimeters distribution hnl/nivs. hl/lminiis shown in logarithmic scales
in Fig. 8. The distribution hnl/niwas constructed as normalized to unity histogram
with the bin equal to 1 . The perimeters of nodal domains lwere calculated by identi-
fying the continues paths of grid sites at the domains boundaries. The averaged values
hnl/niand hl/lminiare defined similarly as previously defined hns/niand hs/smini, e.g.
12
FIG. 9: The normalized signed area distribution NΣAfor the chaotic half-circular microwave rough
billiard. Full line shows predicted by the theory asymptotic limit NΣA0.0386, Blum et al. [1].
hl/lmini=1
NTPNT
i=1 l/l(N)
min, where l(N)
min = 2πqs(N)
min = 2π(j01/kN) is the perimeter of the
circle with the smallest possible area s(N)
min. The full line in Fig. 8 shows the prediction
of percolation theory log10 (hnl/ni) = 15
7log10(hl/lmini). Also in this case the agreement
between the experimental results and the theory is good what is well seen in the range
0.2<log10(hl/lmini)<1.2, which is marked by the two vertical lines. A least squares fit
log10(hnl/ni) = a3τlog10 (hl/lmini) of the experimental results lying within the marked
range yields τ= 2.13 ±0.23 and a3= 0.04 ±0.21. The result of the fit is shown in Fig. 8
by the dashed line. As we see the scaling exponent τ= 2.13 ±0.23 is close to the exponent
predicted by percolation theory τ= 15/72.14. The above results clearly demonstrate
that percolation theory is very useful in description of the properties of wave functions of
chaotic billiards.
Another important characteristic of the chaotic billiard is the signed area distribution
ΣAintroduced by Blum et al. [1]. The signed area distribution is defined as a variance:
ΣA=h(A+A)2i/A2, where A±is the total area where the wave function is positive
(negative) and Ais the billiard area. It is predicted [1] that the signed area distribution
should converge in the asymptotic limit to ΣA0.0386N1. In Fig. 9 the normalized
signed area distribution NΣAis shown for the microwave rough billiard. For lower states
13
80 N250 the points in Fig. 9 were obtained by averaging over 20 consecutive eigenstates
while for higher states N > 250 the averaging over 5 consecutive eigenstates was applied. For
low level numbers N < 220 the normalized distribution NΣAis much above the predicted
asymptotic limit, however, for 220 < N 435 it more closely approaches the asymptotic
limit. This provides the evidence that the signed area distribution ΣAcan be used as a
useful criterion of quantum chaos. A slow convergence of NΣAat low level numbers N
was also observed for the Sinai and stadium billiards [1]. In the case of the Sinai billiard
this phenomenon was attributed to the presence of corners with sharp angles. According
to Blum et al. [1] the effect of corners on the wave functions is mainly accentuated at low
energies. The half-circular microwave rough billiard also possesses two sharp corners and
they can be responsible for a similar behavior.
In summary, we measured the wave functions of the chaotic rough microwave billiard
up to the level number N= 435. Following the results of percolationlike model proposed
by [2] we confirmed that the distributions of the areas sand the perimeters lof nodal
domains have power behaviors nssτand nllτ, where scaling exponents are equal
to τ= 1.99 ±0.14 and τ= 2.13 ±0.23, respectively. These results are in a good agreement
with the predictions of percolation theory [3], which predicts τ= 187/91 2.05 and τ=
15/72.14, respectively. We also showed that in the limit N a least squares fit of the
experimental data yields the asymptotic number of nodal domains N/N 0.058 ±0.006
that is close to the theoretical prediction N/N 0.062 [2]. Finally, we found out that
the signed area distribution ΣAapproaches for high level number Ntheoretically predicted
asymptotic limit 0.0386N1[1].
Acknowledgments. This work was partially supported by KBN grant No. 2 P03B 047
24. We would like to thank Szymon Bauch for valuable discussions.
[1] G. Blum, S. Gnutzmann, and U. Smilansky, Phys. Rev. Lett. 88, 114101-1 (2002).
[2] E. Bogomolny and C. Schmit, Phys. Rev. Lett. 88, 114102-1 (2002).
[3] R. M. Ziff, Phys. Rev. Lett. 56, 545 (1986).
[4] H.-J. St¨ockmann, J. Stein, Phys. Rev. Lett. 64, 2215 (1990).
[5] H.-J. St¨ockmann, Quantum Chaos, an Introduction, (Cambridge University Press, 1999).
14
[6] Y. Hlushchuk, A. B edowski, N. Savytskyy, and L. Sirko, Physica Scripta 64, 192 (2001).
[7] Y. Hlushchuk, L. Sirko, U. Kuhl, M. Barth, H.-J. St¨ockmann, Phys. Rev. E 63, 046208-1
(2001).
[8] K.M. Frahm and D.L. Shepelyansky, Phys. Rev. Lett. 78, 1440 (1997).
[9] K.M. Frahm and D.L. Shepelyansky, Phys. Rev. Lett. 79, 1833 (1997).
[10] A. Shnirelman, Usp. Mat. Nauk. 29, N6, 18 (1974).
[11] L. Sirko, Sz. Bauch, Y. Hlushchuk, P.M. Koch, R. Bl¨umel, M. Barth, U. Kuhl, and H.-J.
St¨ockmann, Phys. Lett. A 266, 331 (2000).
[12] F. Borgonovi, Phys. Rev. Lett. 80, 4653 (1998).
[13] Y. Yamamoto and R.E. Sluster, Phys. Today 46, 66 (1993).
[14] J.U. ockel and A.D. Stone, Nature 385, 45 (1997).
[15] Ya. M. Blanter, A.D. Mirlin, and B.A. Muzykantskii, Phys. Rev. Lett. 80, 4161 (1998).
[16] N. Savytskyy and L. Sirko, Phys. Rev. E 65, 066202-1 (2002).
[17] L.C. Maier and J.C. Slater, J. Appl. Phys. 23, 68 (1952).
[18] S. Sridhar, Phys. Rev. Lett. 67, 785 (1991).
[19] C. Dembowski, H.-D. Gr¨af, A. Heine, R. Hofferbert, H. Rehfeld, and A. Richter, Phys. Rev.
Lett. 84, 867 (2000).
[20] D.H. Wu, J.S.A. Bridgewater, A. Gokirmak, and S.M. Anlage, Phys. Rev. Lett. 81, 2890
(1998).
[21] J. Stein, H.-J. St¨ockmann, and U. Stoffregen, Phys. Rev. Lett. 75, 53 (1995).
[22] S.W. McDonald and A.N. Kaufman, Phys. Rev A 37, 3067 (1988).
[23] M.V. Berry, J. Phys. A 10, 2083 (1977).
15
... Formally, such nodal domains may be defined as the maximally connected regions wherein the wave function does not change sign. Experimentally, nodal domains have also been the focus of much attention [43][44][45] . ...
Article
Full-text available
We consider a semiclassical approach to study the possible stabilization of a zigzag phase for a system of interacting electrons confined within a thin annulus region with infinite inner/outer walls. The electrons are considered spinless and interact via the standard Coulomb interaction potential. The classical minimum energy configuration of the system due to the Coulomb repulsion between electrons is accurately determined by using the simulated annealing method. As the number of electrons increases we see the appearance and stabilization of a two-ring structure with zigzag patterns. Further increase of the number of electrons appears to eventually lead to the collapse of the two-ring zigzag structure. A simple quantum treatment of the nodal features of the free many particle wave function shows the appearance of nodal domain patterns that are consistent with the zigzag features that were observed in the classical model.
... It should be emphasized that the other complex quantum systems can be simulated by microwave plane billiards [29][30][31][32][33][34][35][36][37][38][39][40][41][42][43][44][45] and atoms excited in strong microwave fields [46][47][48][49][50][51][52][53][54][55]. ...
Article
Full-text available
We present the experimental study of the distributions of the reflection amplitudes ri =|Sii| of the two-port scattering matrix S for networks with unitary and symplectic symmetries for the intermediate absorption strength parameter y. The experimental results confirm the theoretical predictions obtained within the framework of the Gaussian unitary and symplectic ensembles of the random matrix theory.
... The experimental verification of wave chaos in microwave billiards has been addressed by several international research groups over the last few decades. This extensive work led to the observation of a rich phenomenology originating from complex wave dynamics [7], including: Wave-function scars [8], chaotic dynamics in superconducting billiards [9], time-reversal symmetry breaking [10], nodal domains in rough billiards [11], time-invariance violation at and around exceptional points [12] as well as electromagnetic reverberation [13]. Concerning optical resonators, it has been shown that chaotic dynamics are a key mechanism in asymmetric geometries [14][15][16]. ...
Article
Full-text available
The propagation of electromagnetic waves in a closed domain with a reflecting boundary amounts, in the eikonal approximation, to the propagation of rays in a billiard. If the inner medium is uniform, then the symplectic reflection map provides the polygonal rays’ paths. The linear response theory is used to analyze the stability of any trajectory. The Lyapunov and reversibility error invariant indicators provide an estimate of the sensitivity to a small initial random deviation and to a small random deviation at any reflection, respectively. A family of chaotic billiards is considered to test the chaos detection effectiveness of the above indicators.
... The RWM has led to many conjectures concerning L p norms, semi-classical measures or volume and topology of nodal domains of chaotic eigenfunctions. Several of these conjectures have been addressed numerically ( [16], [3], [4], [6]) or experimentally ( [18], [7], [20]). ...
Preprint
Full-text available
The Random Wave Conjecture of M. V. Berry is the heuristic that eigenfunctions of a classically chaotic system should behave like Gaussian random fields, in the large eigenvalue limit. In this work we collect some definitions and properties of Gaussian random fields, and show that the formulation of the Berry's conjecture proposed using local weak limits is equivalent to the one that is based on the Benjamini-Schramm convergence. Finally, we see that both these formulations of the Berry's property imply another property known as inverse localization that relates high energy eigenfunctions and solutions to the Euclidean Helmholtz equation.
... Thus, any theoretical or numerical results obtained for quantum graphs can be verified experimentally. One should remark that in simulations of the other complex quantum systems, the model systems are often used, such as flat microwave billiards [25][26][27][28][29][30][31][32][33][34][35][36][37][38] and exited atoms in strong microwave fields [39][40][41][42][43][44]. ...
Article
Full-text available
We discuss a connection between the generalized Euler characteristic Eo(|VDo|) of the original graph which was split at edges into two separate subgraphs and their generalized Euler characteristics Ei(|VDi|), i=1,2, where |VDo| and |VDi| are the numbers of vertices with the Dirichlet boundary conditions in the graphs. Applying microwave networks which simulate quantum graphs, we show that the experimental determination of the generalized Euler characteristics Eo(|VDo|) and Ei(|VDi|), i=1,2 allows finding the number of edges in which the subnetworks were connected.
... Microwave networks are the only ones that allow for the experimental simulation of quantum systems with all three types of symmetry within the framework of the random matrix theory (RMT): Gaussian orthogonal ensemble (GOE)systems with preserved time reversal symmetry (TRS) [16,21,24,25,27,[30][31][32], Gaussian unitary ensemble (GUE)-systems with broken TRS [24,28,[33][34][35][36], and Gaussian symplectic ensemble (GSE)-systems with TRS and half-spin [37]. The other model systems, which are not as versatile as microwave networks, but are often used in simulations of complex quantum systems, are flat microwave billiards [38][39][40][41][42][43][44][45][46][47][48][49][50][51][52][53][54], and exited atoms in strong microwave fields [55][56][57][58][59][60][61][62][63][64][65][66][67]. ...
Article
Full-text available
We show that there is a relationship between the generalized Euler characteristic Eo (|VDo |)of the original graph that was split at vertices into two disconnected subgraphs i = 1, 2 and theirgeneralized Euler characteristics Ei (|VDi |). Here, |VDo | and |VDi | denote the numbers of vertices withthe Dirichlet boundary conditions in the graphs. The theoretical results are experimentally verifiedusing microwave networks that simulate quantum graphs. We demonstrate that the evaluation ofthe generalized Euler characteristics Eo (|VDo |) and Ei (|VDi |) allow us to determine the number ofvertices where the two subgraphs were initially connected .
Article
We investigate properties of the transmission amplitude of quantum graphs and microwave networks composed of regular polygons such as triangles and squares. We show that for the graphs composed of regular polygons, with the edges of the length l, the transmission amplitude displays a band of transmission suppression with some narrow peaks of full transmission. The peaks are distributed symmetrically with respect to the symmetry axis kl=π, where k is the wave vector. For microwave networks the transmission peak amplitudes are reduced and their symmetry is broken due to the influence of internal absorption. We demonstrate that for the graphs composed of the same polygons but separated by the edges of length l′<l, the transmission spectrum is generally not symmetric according to the axis kl′=π. We also show that graphs composed of regular polygons of different size with the edges being irrational numbers are not fully chaotic and their level spacing distribution and the spectral rigidity are well described by the Berry-Robnik distributions. Moreover, the transmission spectrum of such a graph displays peaks which are very close to one. Furthermore, the microwave networks are investigated in the time-domain using short Gaussian pulses. In this case the delay-time distributions, though very sensitive to the internal structure of the networks, show the sequences of transmitted peaks with the amplitudes much smaller than the input one. The analyzed properties of the graphs and networks suggest that they can be effectively used to manipulate quantum and wave transport.
Chapter
We discuss a role of the boundary conditions in the graphs split at vertices. In the analysis we take advantage of the properties of a new spectral invariant for quantum graphs and networks: the generalized Euler characteristics \(\mathcal E_i(|V_{D_i}|) \), where \(|V_{D_i}|\) denotes the number of vertices with the Dirichlet boundary conditions in the ith graph. The theoretical results are verified numerically.KeywordsQuantum graphsEuler characteristicNeumann and Dirichlet boundary conditionsMicrowave networks
Article
Full-text available
We analyze the situation when the original graph is split at edges and vertices into two disconnected subgraphs. We show that there is a relationship between the generalized Euler characteristic Eo(|VDo|) of the original graph and the generalized Euler characteristics Ei(|VDi|), i = 1, 2, of two disconnected subgraphs, where |VDo| and |VDi|, i = 1, 2, are the numbers of vertices with the Dirichlet boundary conditions in the graphs. Theoretical predictions are verified experimentally using microwave networks which simulate quantum graphs.
Article
We investigated properties of a singular billiard, that is, a quantum billiard which contains a pointlike (zero-range) perturbation. A singular billiard was simulated experimentally by a rectangular microwave flat resonator coupled to microwave power via wire antennas which act as singular scatterers. The departure from regularity was quantitatively estimated by the short-range plasma model in which the parameter η takes the values 1 and 2 for the Poisson and semi-Poisson statistics, respectively. We show that in the regime of semi-Poisson statistics the experimental power spectrum and the second nearest-neighbor-spacing distribution P(2,s) are in good agreement with their theoretical predictions. Furthermore, the measurement of the two-port scattering matrix allowed us to evaluate experimentally the enhancement factor F(γtot) in the regime of the semi-Poisson statistics as a function of the total absorption factor γtot. The experimental results were compared with the analytical formula for F(γtot) evaluated in this article. The agreement between the experiment and theory is good.
Article
Full-text available
We have measured the probability density \psi(r)\(2) in the semiclassical limit of a classically chaotic square well potential with and without time reversal symmetry, and compared our findings with theoretical predictions. We find that wave functions with time-reversal symmetry have larger fluctuations than those without time-reversal symmetry. To quantify the degree of these fluctuations, eigenmodes both with and without time-reversal symmetry are statistically analyzed and the two-point spatial correlation function and the probability density distribution function of the eigenmodes are found to agree with theoretical predictions.
Book
This book introduces the quantum mechanics of classically chaotic systems, or quantum chaos for short. The author's philosophy has been to keep the discussion simple and to illustrate theory, wherever possible, with experimental or numerical examples. The microwave billiard experiments, initiated by the author and his group, play a major role in this respect. Topics covered include the various types of billiard experiment, random matrix theory, systems with periodic time dependences, the analogy between the dynamics of a one-dimensional gas with a repulsive interaction and spectral level dynamics, where an external parameter takes the role of time, scattering theory distributions and fluctuation, properties of scattering matrix elements, semiclassical quantum mechanics, periodic orbit theory, and the Gutzwiller trace formula. This book will be of great value to anyone working in quantum chaos.
Article
Studies of optical microresonators with dimensions between 0.1 and 10 microns are now under way in a wide variety of condensed matter systems. Ideally, one can isolate a single mode of the optical field in a cube a halfwavelength on a side with perfectly reflecting walls. Liquiddroplets,polymer spheres and semiconductor Fabry‐Perot microcavities with dielectricmirrors are examples of microresonators with which one can approach this ideal limit and nearly isolate a few modes of the electromagnetic field from the continuum of surrounding free‐space modes. A new generation of optical microresonators is making possible the exploration of quantum electrodynamic phenomena in condensed matter systems and providing microlasers with a wide range of potential applications.
Article
The form of the wavefunction psi for a semiclassical regular quantum state (associated with classical motion on an N-dimensional torus in the 2N-dimensional phase space) is very different from the form of psi for an irregular state (associated with stochastic classical motion on all or part of the (2N-1)-dimensional energy surface in phase space). For regular states the local average probability density Pi rises to large values on caustics at the boundaries of the classically allowed region in coordinate space, and psi exhibits strong anisotropic interference oscillations. For irregular states Pi falls to zero (or in two dimensions stays constant) on 'anticaustics' at the boundary of the classically allowed region, and psi appears to be a Gaussian random function exhibiting more moderate interference oscillations which for ergodic classical motion are statistically isotropic with the autocorrelation of psi given by a Bessel function.
Article
It is shown that the frequency of a resonant electromagnetic cavity is perturbed by inserting a metallic sphere, needle, or disk of dimensions small compared to a wavelength by an amount depending upon the local electric and magnetic field at the position of the perturbing object. This perturbation is calculated for ellipsoidal objects of needle‐shaped, spherical, and disk‐shaped form. The perturbations by the different objects depend upon different components of electric and magnetic fields, and by combining measurements with all three, it is in theory possible to measure all the field components. Experimental checks of the calculations are described, resulting in satisfactory agreement between theory and experiment except with the needles, in which the perturbation is very sensitive to the precise shape of the object, and the needles used were not accurate enough ellipsoids to give satisfactorily quantitative results.
Article
We measure the angular momentum content of modes in a flat, near-circular microwave cavity with a rough perimeter and demonstrate localization in angular momentum space. Because Schrödinger's wave mechanics and Maxwell's electrodynamics are equivalent for a 2d cavity, we compare our experimental results directly with the quantum theory of rough 2d cavities [K.M. Frahm and D.L. Shepelyansky, Phys. Rev. Lett. 78 (1997) 1440]. Introducing the concept of effective roughness we find good qualitative agreement.
Article
We numerically investigate statistical properties of short-wavelength normal modes and the spectrum for the Helmholtz equation in a two-dimensional stadium-shaped region. As the geometrical optics rays within this boundary (billiards) are nonintegrable, this wave problem serves as a simple model for the study of quantum chaos. The local spatial correlation function and the probability distribution Pn(psi) of wave amplitude for normal modes psin are computed and compared with predictions based on semiclassical arguments applied to this nonintegrable Hamiltonian. The spectrum is analyzed in terms of the probability P(DeltaE) of neighboring energy-eigenvalue separations, which is shown to be similar to a Wigner distribution for the eigenvalues of a random matrix.
Article
The formula for the fractal dimensionality of the perimeter (hull) of a percolation cluster, d'f=1+1/nu, proposed recently by Sapoval, Rosso, and Gouyet, is shown to imply for the perimeter scaling exponents tau'=1+2nu/(1+nu), sigma'=1/(1+nu), and gamma'=2. Monte Carlo simulations of very large perimeter-generating walks yield tau'=2.143+/-0.002 and d'f=1.751+/-0.002, consistent with these predictions (on the assumption that nu=(4/3)). The walks are also used to determine pc=0.5927 5+/-0.0000 3 for site percolation on a square lattice.
Article
The eigenfrequencies of resonance cavities shaped as stadium or Sinai billiards are determined by microwave absorption. In the applied frequency range 0-18.74 GHz the used cavities can be considered as two dimensional. For this case quantum-mechanical and electromagnetic boundary conditions are equivalent, and the resonance spectrum of the cavity is, if properly normalized, identical with the quantum-mechanical eigenvalue spectrum. Spectra, containing up to a thousand eigenfrequencies, are obtained within minutes. Statistical properties of the spectra as well as their correlation with classical periodic orbits are discussed.
Article
The wave functions of Sinai-billiard-shaped microwave cavities are experimentally studied. Some of the general features observed are parity breaking in the lowest eigenstates, 'bouncing-ball' states, and states with quasi-rectangular or quasi-circular symmetry. The above features are associated with nonisolated periodic orbits. Some states are observed which can be associated with isolated periodic orbits, leading to scars. This work represents the first direct experimental observation of scarred eigenfunctions. At high frequencies the eigenfunctions are very complex, and are yet to be classified in a general scheme.