Vous êtes sur la page 1sur 8

Physics Letters A 305 (2002) 1825

www.elsevier.com/locate/pla

Thermal noise caused by an inhomogeneous loss in the mirrors


used in the gravitational wave detector
Kazuhiro Yamamoto , Masaki Ando, Keita Kawabe 1 , Kimio Tsubono
Department of Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
Received 4 September 2002; received in revised form 4 October 2002; accepted 4 October 2002
Communicated by P.R. Holland

Abstract
We have evaluated the thermal noise of mirrors with inhomogeneous loss in interferometric gravitational wave detectors
using a reliable estimation method. Our calculation showed that the traditional estimation by modal expansion is very different
from the actual thermal noise. In a typical model case, when the loss is localized on a surface illuminated by a laser beam, the
thermal noise is about three times larger than the estimation of the modal expansion. When the dissipation is concentrated at
points far from the beam spot, the thermal noise is about fifteen times smaller.
2002 Elsevier Science B.V. All rights reserved.
PACS: 05.40.Jc; 04.80.Nn
Keywords: Thermal noise of mirror; Interferometric gravitational wave detector; Inhomogeneous loss; Direct approach; Modal expansion

1. Introduction
In precise measurements, such as gravitational experiments, the thermal fluctuation of the mechanical
components is a fundamental noise source. Since a direct measurement of thermal noise is generally difficult, an estimation of the thermal motion is important
in studying the noise properties. Modal expansion [1]
is a frequently employed method to estimate thermal
noise. Nevertheless, our recent experiment [2] proved
* Corresponding author. Present address: Gravitational wave
group, Institute for Cosmic Ray Research, University of Tokyo, 51-5 Kashiwa-no-Ha, Kashiwa, Chiba 277-8582, Japan.
E-mail address: yamak@icrr.u-tokyo.ac.jp (K. Yamamoto).
1 Present address: Max-Planck-Institut fr Gravitationsphysik,
Albert-Einstein-Institut, D-85748 Garching, Germany.

that this method is invalid in systems with the loss distributed inhomogeneously. Thermal motions must be
evaluated using other reliable methods when dissipation is not uniform.
The thermal noise of mirrors in interferometric
gravitational wave detectors [37] was evaluated based
on modal expansion [810]. However, the dissipation
in the mirror is distributed inhomogeneously. For example, measurements of the Q-values suggest that the
loss is localized on the surfaces of the mirror [1113]
and near magnets glued on the mirrors to control their
positions [14,15]. Since the thermal noise of a mirror
derived from the modal expansion is already a limiting factor of the sensitivity of the interferometer in the
observation band, an invalidity of the modal expansion is a serious problem. There have only been a few
studies [1619] for the thermal noise of a mirror with

0375-9601/02/$ see front matter 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 5 - 9 6 0 1 ( 0 2 ) 0 1 3 8 9 - 0

K. Yamamoto et al. / Physics Letters A 305 (2002) 1825

inhomogeneous loss. Levins research is one of them.


Levin developed a new method, direct approach, to estimate thermal noise. It is expected that this method is
valid even when the loss is not homogeneous. Levin
discussed the thermal noise of a mirror with inhomogeneous loss qualitatively using this method. Levins conclusion is that the thermal noise is larger than
the estimation of the modal expansion when the loss
is concentrated near the beam spot. Levin also predicted that the thermal noise caused by loss localized
far from the beam spot is smaller than that caused by
loss near the beam spot. Levins discussion showed
the qualitative property of the thermal noise caused
by inhomogeneous loss. However, the thermal motions
caused by various inhomogeneous losses and the difference from the estimation of the modal expansion
have not been discussed quantitatively using a direct
approach.
We calculated the thermal noise of a mirror with
inhomogeneous losses using a direct approach. The
evaluation of the direct approach was compared with
an estimation from modal expansion. We found a
large difference between the results of the direct
approach and modal expansion. When the loss is
concentrated on a surface illuminated by a laser
beam, the result of the direct approach is about
three times larger than the calculation of the modal
expansion. When the loss is concentrated at points
far from beam spot, the evaluation of the direct
approach is about fifteen times smaller. The effects
of our conclusion on the research strategy concerning
thermal noise are considered in the last part of this
article.

2. Estimation methods
The details of the direct approach and modal
expansion are described here. Using these methods,
a mirror was treated as an elastic cylinder. It was
assumed that the loss is inhomogeneous structure
damping because ordinary dissipation in materials is
expressed by the structure damping model [20,21].
In order to describe this dissipation, we adopted the
complex Youngs modulus, expressed as


E = E0 1 + i(r) ,

(1)

19

where is called the loss angle. In structure damping,


is independent of the frequency and depends on the
position, r, in the mirror.
2.1. Direct approach
In the direct approach, the thermal noise is derived by applying the fluctuationdissipation theorem
(FDT) [22] to dissipated power in a mirror calculated
directly from the equation of motion. Since the FDT is
a universal and reliable theorem, the direct approach is
valid even when the loss is not uniform.
The FDT predicts the relationship between the
spectrum of thermal noise and the dissipated energy
[16]. This relation is described as
G(f ) =

2kB T Wloss
,
2 f 2 F0 2

(2)

where G is the power spectrum density of the thermal


noise, kB is the Boltzmann constant, T is the temperature, and f is the frequency. Wloss is the average dissipated power in the mirror when oscillatory pressure,
F0 cos(2f t)P (r), is applied to a surface of the mirror. The function, P , represents the laser-beam profile.
Since the beam profile is Gaussian, P is written in the
form


2
2r 2
P (r) =
(3)
exp

,
r0 2
r0 2
where r0 is the beam radius, and r is the distance from
the optical axis.
When the complex Youngs modulus is adopted, the
dissipated power, Wloss , is written as

Wloss = 2f Eelas (r)(r) dV ,
(4)
where Eelas is the elastic energy density when the
strain is maximum. This energy density is calculated
from the equation of motion directly. The observation
band of the gravitational wave detectors (about several
kHz, at most) is lower than the resonant frequency of
the fundamental mode of the mirrors (about several
kHz, at least). Thus, in order to estimate Eelas , it is
an appropriate approximation that a constant pressure,
F0 P (r), is applied to the mirror surface. We used
ANSYS, which is a program used for the finiteelement method (FEM) to calculate the elastic energy
density, Eelas . In FEM, the mirror is divided into fine

20

K. Yamamoto et al. / Physics Letters A 305 (2002) 1825

meshes to solve the equation of motion. The size of the


mesh was ten times smaller than the mirror radius. In
order to confirm that the estimation of thermal noise
did not depend on how to divide a mirror into meshes,
the thermal noise was evaluated from three kinds of
meshing. The results proved that the estimated value
was independent of the meshing.
2.2. Modal expansion
The modal expansion is a frequently used method
to calculate the thermal fluctuation of mechanical
systems. In this method, the thermal motion of the
total system is the summation of those of the resonant
modes. The thermal motion of each mode is obtained
from the Q-values. When the dissipation is expressed
by the structure damping model, the formula for the
thermal noise derived from the modal expansion in a
frequency range which is lower than the first resonant
mode is written as [810]
 4kB T 1
,
G(f ) =
(5)
mn n 2 Qn
n
where mn , n , and Qn are the effective mass, angular
resonant frequency, and Q-value of the nth mode,
respectively.
When these parameters are calculated, the method
proposed by Hutchinson [23] is useful. This method
is a semi-analytical algorithm to simulate resonant
modes of an isotropic elastic cylinder. The resonant
angular frequency, n , and the displacement, wn , of
the nth resonant mode are derived from this method.
The effective mass, mn , is expressed as

|wn (r)|2 dV
mn = volume
(6)
,
( surface wn,z P (r) dS)2
where is the density, wn,z is the z-component of w n .
The z-axis is parallel to the optical axis. The Q-value,
Qn , is written as

En,elas (r) dV
Qn = 
(7)
,
En,elas (r)(r) dV
where En,elas is the elastic energy density of the
nth mode. This energy density is derived from the
displacement, w n . Thus, all parameters of each mode
(mn , n , and Qn ) are derived from the Hutchinsons
method.

3. Result

The thermal noise of a mirror with inhomogeneous


losses was derived from the direct approach, and compared with an evaluation from the modal expansion.
The mirrors of the interferometric gravitational wave
detector projects are similar. As a typical example for
calculations, we selected the specification of the mirror of TAMA300 [6,7], which is the Japanese interferometric gravitational wave detectors. The material
of the TAMA mirror is fused silica. The shape is a
cylinder. This mirror is 50 mm in radius and 60 mm
in height. The center of the beam spot is at the center of the flat surface of the mirror. Since the Fabry
Perot cavity geometry affects the beam radii at mirrors, the dependence of the amplitude of the thermal
noise on the beam radius was investigated. The range
of the beam radius was between 5 mm and 30 mm. The
beam radius of TAMA300 is 8 mm at the front mirror
and 15 mm at the end mirror.
To simplify the discussion, the loss angle, (r),
was assumed to be a constant value, , in the damped
volume and (r) to be zero in other regions. Two
types of distributions of dissipation were considered.
In the first type, the loss was concentrated on a surface.
These losses corresponded to the dissipation in the reflective coating or surface damage caused by inadequate polishing. In the second type, the dissipation was
localized at points. These losses represent glued magnets to control the position of the mirrors and stand-off
to fix the wire on the mirror.

3.1. Homogeneous loss

In order to check the validity of our calculation,


the thermal noise of a mirror with homogeneous loss
was calculated. This result must be the same as the
estimation of the modal expansion. Fig. 1 shows the
estimation of the amplitude of the thermal noise at 100
Hz. The closed circles and open squares represent the
results of the direct approach and modal expansion,
respectively. In this calculation, the loss angle, ,
was 106 . These results of the direct approach agree
with those of the modal expansion. Therefore, our
calculation did not have any serious problems.

K. Yamamoto et al. / Physics Letters A 305 (2002) 1825

21

Fig. 1. Thermal noise caused by homogeneous loss. This graph


shows the dependence of the amplitude of thermal noise at 100 Hz
on the beam radius. The closed circles and open squares represent
the results of the direct approach and modal expansion, respectively.

3.2. Surface
Three distributions of the loss concentrated on
the surface were considered. These distributions are
shown in Fig. 2. The shadows show a surface on which
the dissipation was localized. The other part had no
losses. The loss was concentrated on the flat surface
illuminated by the beam, the other flat surface, and the
cylindrical surface in each distribution. These models
are called Front, Back, and Cylindrical surface,
respectively. It was supposed that the loss layer is 5 m
in thickness. This is a typical thickness of the coating
used in interferometric gravitational wave detectors.
In our calculation, the loss angle, , was 104 in the
loss layer. The recent measurements show that the loss
angle of the coating layer is about 104 [24,25].
The results of calculations concerning these surface
models are shown in Fig. 3. The closed, open, and
grey circles represent the thermal motions of the Front,
Back, and Cylindrical surface models calculated from
the direct approach, respectively. The open squares are
an estimation of the Front and Back surface models
from the modal expansion. The estimations from the
modal expansion in both cases are the same because
the Q-values are the same. The grey squares show
the estimation of the Cylindrical surface derived from
the modal expansion. These results show that there
is a large discrepancy between the actual thermal
noise and the estimation from the modal expansion.
The thermal noise of the Front surface model is

Fig. 2. Dissipation concentrated on a surface. The shadows show a


surface on which the dissipation was localized. The other part had no
losses. These losses corresponded to the dissipation in the reflective
coating or surface damage caused by inadequate polishing. In the
upper part, the loss was concentrated on the flat surface illuminated
by the beam. This distribution is called Front surface. In the middle
part, the dissipation was localized on the other flat surface. This
is called Back surface. In the lower part, the dissipation was
concentrated on the cylindrical surface. This is called Cylindrical
surface.

about three times larger than the evaluation from the


modal expansion. The thermal motion of the Back and
Cylindrical surface model is about four and six times
smaller, respectively. This result agrees with Levins
qualitative consideration: the thermal noise caused by
the loss near the beam spot (Front surface model) is
larger than that caused by the dissipation which is
far from the beam spot (Back and Cylindrical surface
models).
Levin predicted that the amplitude of the thermal

noise of the Front surface model is at least R/r0


times larger than that calculated based on modal expansion [16], where R is the mirror radius. The crosses
in Fig. 3 represent the product of the estimation of

22

K. Yamamoto et al. / Physics Letters A 305 (2002) 1825

Fig. 3. Thermal fluctuations of surface models. This graph shows


the dependence of the amplitude of the thermal noise at 100 Hz
on the beam radius. The closed, open, and grey circles represent
the thermal motions of the Front, Back, and Cylindrical surface
models calculated from the direct approaches, respectively. The
open squares are an estimation of the Front and Back surface
models from the modal expansion. The estimations from the modal
expansion in both cases are the same because the Q-values are
the same. The grey squares show the evaluation of the Cylindrical
surface models derived from the modal expansion. The crosses
represent Levins prediction about the Front surface model. The line
shows an analytical estimation by Nakagawa et al. of the thermal
noise of the Front surface model.

the modal expansion and R/r0 . Even though Levins discussion was not strict, the difference between
Levins prediction and our result is small. Nakagawa
et al. [26] calculated analytically the thermal noise of
the Front surface model using the direct approach and
the other method developed by them [27]. This analytical result is shown in Fig. 3. It is consistent with
our numerical result. When the beam radius is large,
there is a slight difference. This is because the mirror
was treated as a half-infinite elastic body in the analytical calculation. When the beam radius is comparable
to the mirror radius, the infinite elastic-body approximation in the analytical method is not appropriate.
The dependence of the thermal motion on the beam
radius is different from the estimation of modal expansion. Modal expansion predicts that the amplitude is almost inversely proportional to the square
root of the beam radius. However, the amplitude of
the Front surface model is almost inversely proportional to the beam radius. This result is consistent with
Levins qualitative discussion [16]. The amplitudes of
the Back and Cylindrical surface models are almost

Fig. 4. Dissipation concentrated at points. The black dots indicate


points at which the loss was concentrated. The other part had no
losses. Every model had four points. These losses corresponded to
the loss of glued magnets and stand-offs. In the upper part, points
were on the flat surface illuminated by the beam. This distribution is
called Front point. In the middle part, points were on the other flat
surface. This is called Back point. In the lower part, points were
on the cylindrical surface. This is called Cylindrical point.

independent of the beam radius. This is because the


loss is far from the beam spot, and the elastic energy
density at the lossy areas is almost independent of the
beam radius.
3.3. Point
Three distributions of loss concentrated at points
were considered. These distributions are shown in
Fig. 4. The black dots indicate points at which the
loss was concentrated. The other part had no losses.
Every model had four points. These four points were
on the flat surface illuminated by the beam, the other
flat surface, and the cylindrical surface in each distribution. They are called Front, Back, and Cylindrical point, respectively. In TAMA, the magnet po-

K. Yamamoto et al. / Physics Letters A 305 (2002) 1825

Fig. 5. Thermal motions of the point models. This graph shows


the dependence of the amplitude of thermal noise at 100 Hz on
the beam radius. The closed circles represent the thermal noise of
the Front point models evaluated from the direct approach. The
open circles are the thermal noise of the Back and Cylindrical point
models evaluated from the direct approach. The open squares are
the estimation of the Front and Back point models from the modal
expansion. The estimations from the modal expansion in both the
cases are the same because the Q-values are the same. The grey
squares are an estimation of the Cylindrical point model from the
modal expansion.

sitions are the same as that of the Back point model.


In other projects, the magnets are also glued near the
edge of the flat surface of mirrors. The stand-offs are
glued on the TAMA mirror in order to fix wires on the
mirror. The positions of the stand-offs are the same as
that of the Cylindrical point model. The radius of the
damped region was 0.5 mm. This is the same as that
of the magnet of TAMA. In the Front and Back point
models, the point was 45 mm away from the center of
the flat surface. In the Cylindrical point model, all four
points were on a horizontal plane which included the
center of the mirror. The distance between a point and
a flat surface of the mirror was 20 mm. These positions
are the same as those of the magnets and stand-offs of
TAMA. It was assumed that the thickness of the equivalent loss layer was 0.1 mm and that the loss angle, ,
was 1 in the damped volume. From these distributions
of the losses and Eq. (7), the calculated Q-values were
on the order of 105 . The measured Q-values of the
mirror with magnets and stand-offs of TAMA were
larger than 105 [15].
The results of point models are shown in Fig. 5. The
closed circles represent the thermal fluctuations of the
Front point models evaluated from the direct approach.

23

The open circles are the thermal motions of the Back


and Cylindrical point model from the direct approach.
The open squares are an estimation of the Front and
Back point models from the modal expansion. The estimations from the modal expansion in both cases are
the same, because the Q-values are the same. The grey
squares show the thermal fluctuation of the Cylindrical point model derived from modal expansion. In all
of the models, the amplitude estimated from the direct approach is about fifteen times smaller than the
evaluation from the modal expansion. The discrepancy
between the calculations of the direct approach and
modal expansion is larger than that of the surface models, because of a larger inhomogeneity. The thermal
noise of the Front point model is larger than those of
the Back and Cylindrical point models because the distance between the loss and the beam spot in the Front
point is smaller.
The dependence of the thermal motion on the
beam radius is different from the estimation of the
modal expansion. The modal expansion predicts that
the amplitude is almost inversely proportional to
the square root of the beam radius. However, the
amplitudes of the thermal fluctuations in all point
models are almost independent of the beam radius, just
as for the Back and Cylindrical surface models. The
reason is the same as that of the Back and Cylindrical
surface models: the loss is far from the beam spot and
the elastic energy density at the lossy areas is almost
independent of the beam radius.

4. Discussion
Our calculation proves that there is a large discrepancy between the actual thermal noise and the estimation from the modal expansion. Since the study of
thermal noise is based on modal expansion, our results
have large effects on the strategy of research of thermal fluctuation. The implications of our results on the
research of the thermal noise of mirrors are discussed.
Our results reveal that the loss concentrated on
the surface illuminated by a laser beam is serious.
For example, Fig. 3 shows that the thermal noise
caused bythe coating in TAMA300 is about 6.0
1020 m/ Hz at 100 Hz. This value is larger than the
goal level of the
thermal noise of the mirror of TAMA
(5 1020 m/ Hz at 100 Hz), even though the

24

K. Yamamoto et al. / Physics Letters A 305 (2002) 1825

modal expansion predicts


that this thermal fluctuation

(2.2 1020 m/ Hz at 100 Hz) is smaller than


the goal sensitivity. In km-scale interferometers, the
beam radius is about 3 cm. The thermal noise caused
by the coating loss inkm-scale interferometers is
about 2.0 1020 m/ Hz at 100 Hz. This value
is larger than the goal sensitivity of future projects,
like LCGT
[28] and LIGO II [29], which is about
1020 m/ Hz. Nevertheless, the properties of the loss
of a reflective coating and a polished surface were not
well researched. Our calculation shows the necessity
of further investigations of the loss in the reflective
coating and the polished surface.
Our research proves that the thermal noise caused
by glued magnets and stand-offs is not a serious
problem. For example, according to Fig. 5, modal
expansion predicts that the thermal noise
caused by

magnets in TAMA is 4.3 1019 m/ Hz at 100 Hz.


Since this value is nine times larger than the goal
sensitivity, it was considered that the thermal noise
caused by the magnets is a serious problem. However,
the direct approach proves that the actual amplitude
of thermal noise caused
by the magnets is smaller

(2.7 1020 m/ Hz at 100 Hz) than the goal


sensitivity of TAMA. Since the mirror of the kmscale interferometer is larger than that of TAMA,
it is expected that the loss and thermal noise of
a km-scale interferometer is smaller than those of
TAMA. Probably, the thermal noise of the km-scale
interferometer is about as large as the goal of the
future projects. From above discussion, the loss of the
magnets and stand-offs is not a serious problem, even
in future interferometers.

5. Conclusion
The modal expansion is frequently used to estimate
the thermal motions of mirrors in interferometric gravitational wave detectors. However, our previous experiment [2] showed that this method is not valid when
the dissipation is distributed inhomogeneously. On the
other hand, the direct approach proposed by Levin [16]
is correct even when the loss is distributed inhomogeneously. Nevertheless, the thermal fluctuations caused
by various inhomogeneous losses have not been investigated quantitatively using Levins approach. We calculated the thermal noise of the mirror of TAMA with

the inhomogeneous losses using the direct approach,


and found that there is a large difference between the
evaluated values of the direct approach and the modal
expansion. When the loss is concentrated on a surface
illuminated by a beam, the amplitude of thermal noise
is about three times larger than the calculated value
based on the modal expansion. When the dissipation is
localized near the glued magnets and stand-offs, the
amplitude is about fifteen times smaller. These conclusions show that studying the loss in the reflective
coating and the polished surface is important, and that
the thermal noise is not seriously affected by the glued
magnets and stand-offs despite the large reduction of
the Q-values.

Acknowledgements
We are grateful to T. Suzuki and T. Tomaru for
useful suggestions about ANSYS. This research is
supported in part by Research Fellowships of the
Japan Society for the Promotion of Science for Young
Scientists, and by a Grant-in-Aid for Creative Basic
Research of the Ministry of Education.

References
[1] P.R. Saulson, Phys. Rev. D 42 (1990) 2437.
[2] K. Yamamoto, S. Otsuka, M. Ando, K. Kawabe, K. Tsubono,
Phys. Lett. A 280 (2001) 289.
[3] A. Abramovici, W.E. Althouse, R.W.P. Drever, Y. Grsel,
S. Kawamura, F.J. Raab, D. Shoemaker, L. Sievers, R.E. Spero,
K.S. Thorne, R.E. Vogt, R. Weiss, S.E. Whitcomb, M.E.
Zucker, Science 256 (1992) 325.
[4] VIRGO Collaboration, VIRGO Final Design Report, 1997.
[5] K. Danzmann, H. Lck, A. Rdiger, R. Schilling, M. Schrempel, W. Winkler, J. Hough, G.P. Newton, N.A. Robertson, H. Ward, A.M. Campbell, J.E. Logan, D.I. Robertson, K.A. Strain, J.R.J. Bennett, V. Kose, M. Khne, B.F.
Schutz, D. Nicholson, J. Shuttleworth, H. Welling, P. Aufmuth, R. Rinkleff, A. Tnnermann, B. Willke, Proposal for
a 600 m Laser-Interferometric Gravitational Wave Antenna,
Max-Planck-Institut fr Quantenoptik Report 190, Garching,
Germany, 1994.
[6] K. Tsubono, TAMA Collaboration, in: Proc. TAMA International Workshop on Gravitational Wave Detection, Saitama,
Japan, November 1996, Univ. Academy Press, 1997, p. 183.
[7] M. Ando, TAMA Collaboration, Phys. Rev. Lett. 86 (2001)
3950.
[8] A. Gillespie, F. Raab, Phys. Rev. D 52 (1995) 577.
[9] F. Bondu, J.-Y. Vinet, Phys. Lett. A 198 (1995) 74.

K. Yamamoto et al. / Physics Letters A 305 (2002) 1825

[10] K. Yamamoto, K. Kawabe, K. Tsubono, in: Proc. TAMA International Workshop on Gravitational Wave Detection, Saitama,
Japan, November 1996, Univ. Academy Press, 1997, p. 373.
[11] K. Numata, G.B. Bianc, N. Ohishi, A. Sekiya, S. Otsuka,
K. Kawabe, M. Ando, K. Tsubono, Phys. Lett. A 276 (2000)
37.
[12] W.J. Startin, M.A. Beilby, P.R. Saulson, Rev. Sci. Instrum. 69
(1998) 3681.
[13] A.M. Gretarsson, G.M. Harry, Rev. Sci. Instrum. 70 (1999)
4081.
[14] N. Ohishi, K. Kawabe, K. Tsubono, in: Proc. TAMA International Workshop on Gravitational Wave Detection, Saitama,
Japan, November 1996, Univ. Academy Press, 1997, p. 369.
[15] TAMA Collaboration, TAMA Status Report 99, 1999, p. 109
(in Japanese).
[16] Y. Levin, Phys. Rev. D 57 (1998) 659.
[17] A. Gillespie, Ph.D. thesis, California Institute of Technology,
1995.
[18] J.E. Logan, N.A. Robertson, J. Hough, P.J. Veitch, Phys. Lett.
A 161 (1991) 101.
[19] K. Yamamoto, S. Otsuka, M. Ando, K. Kawabe, K. Tsubono,
Class. Quantum Grav. 19 (2002) 1689.
[20] T.J. Quinn, C.C. Speake, L.M. Brown, Philos. Mag. A 65
(1992) 261.
[21] P.R. Saulson, R.T. Stebbins, F.D. Dumont, S.E. Mock, Rev. Sci.
Instrum. 65 (1994) 182.

25

[22] H.B. Callen, R.F. Greene, Phys. Rev. 86 (1952) 702.


[23] J.R. Hutchinson, J. Appl. Mech. 47 (1980) 901.
[24] D.R.M. Crooks, P. Sneddon, G. Cagnoli, J. Hough, S. Rowan,
M.M. Fejer, E. Gustafson, R. Route, N. Nakagawa, D. Coyne,
G.M. Harry, A.M. Gretarsson, Class. Quantum Grav. 19 (2002)
883.
[25] G.M. Harry, A.M. Gretarsson, P.R. Saulson, S.E. Kittelberger,
S.D. Penn, W.J. Startin, S. Rowan, M.M. Fejer, D.R.M.
Crooks, G. Cagnoli, J. Hough, N. Nakagawa, Class. Quantum
Grav. 19 (2002) 897.
[26] N. Nakagawa, A.M. Gretarsson, E.K. Gustafson, M.M. Fejer,
Phys. Rev. D 65 (2002) 102001.
[27] N. Nakagawa, B.A. Auld, E. Gustafson, M.M. Fejer, Rev. Sci.
Instrum. 68 (1997) 3553.
[28] K. Kuroda, M. Ohashi, S. Miyoki, D. Tatsumi, S. Sato,
H. Ishizuka, M.-K. Fujimoto, S. Kawamura, R. Takahashi,
T. Yamazaki, K. Arai, M. Fukushima, K. Waseda, S. Telada,
A. Ueda, T. Shintomi, A. Yamamoto, T. Suzuki, Y. Saito,
T. Haruyama, N. Sato, K. Tsubono, K. Kawabe, M. Ando,
K. Ueda, H. Yoneda, M. Musha, N. Mio, S. Moriwaki,
A. Araya, N. Kanda, M.E. Tobar, Int. J. Mod. Phys. D 8 (1999)
557.
[29] LIGO project, LIGO II Conceptual Project Book, 1999, http://
www.ligo.caltech.edu/docs/M/M990288-A1.pdf.

Vous aimerez peut-être aussi