scispace - formally typeset
Open AccessJournal ArticleDOI

Practical speed meter designs for quantum nondemolition gravitational-wave interferometers

Patricia Purdue, +1 more
- 27 Dec 2002 - 
- Vol. 66, Iss: 12, pp 122004
Reads0
Chats0
TLDR
In this paper, the authors describe some practical variants of speed-meter interferometers that can beat the standard quantum limit (SQL) by an arbitrarily large amount, over an arbitrarily wide range of frequencies.
Abstract
In the quest to develop viable designs for third-generation optical interferometric gravitational-wave detectors (e.g., LIGO-III and EURO), one strategy is to monitor the relative momentum or speed of the test-mass mirrors, rather than monitoring their relative position. A previous paper analyzed a straightforward but impractical design for a speed-meter interferometer that accomplishes this. This paper describes some practical variants of speed-meter interferometers. Like the original interferometric speed meter, these designs in principle can beat the gravitational-wave standard quantum limit (SQL) by an arbitrarily large amount, over an arbitrarily wide range of frequencies. These variants essentially consist of a Michelson interferometer plus an extra ``sloshing'' cavity that sends the signal back into the interferometer with opposite phase shift, thereby cancelling the position information and leaving a net phase shift proportional to the relative velocity. In practice, the sensitivity of these variants will be limited by the maximum light power W-circ circulating in the arm cavities that the mirrors can support and by the leakage of vacuum into the optical train at dissipation points. In the absence of dissipation and with squeezed vacuum (power squeeze factor e(-2R)similar or equal to0.1) inserted into the output port so as to keep the circulating power down, the SQL can be beat by h/h(SQL)similar torootW(circ)(SQL)e(-2R)/W-circ at all frequencies below some chosen f(opt)similar or equal to100 Hz. Here W(circ)(SQL)similar or equal to800 kW(f(opt)/100 Hz)(3) is the power required to reach the SQL in the absence of squeezing. (However, as the power increases in this expression, the speed meter becomes more narrow band; additional power and reoptimization of some parameters are required to maintain the wide band. See Sec. III B.) Estimates are given of the amount by which vacuum leakage at dissipation points will debilitate this sensitivity (see Fig. 12); these losses are 10% or less over most of the frequency range of interest (fgreater than or similar to10 Hz). The sensitivity can be improved, particularly at high freqencies, by using frequency-dependent homodyne detection, which unfortunately requires two 4-km-long filter cavities (see Fig. 4).

read more

Content maybe subject to copyright    Report

Practical speed meter designs for quantum nondemolition gravitational-wave interferometers
Patricia Purdue and Yanbei Chen
Theoretical Astrophysics, California Institute of Technology, Pasadena, California 91125
Received 27 July 2002; published 27 December 2002
In the quest to develop viable designs for third-generation optical interferometric gravitational-wave detec-
tors e.g., LIGO-III and EURO, one strategy is to monitor the relative momentum or speed of the test-mass
mirrors, rather than monitoring their relative position. A previous paper analyzed a straightforward but imprac-
tical design for a speed-meter interferometer that accomplishes this. This paper describes some practical
variants of speed-meter interferometers. Like the original interferometric speed meter, these designs in prin-
ciple can beat the gravitational-wave standard quantum limit SQL by an arbitrarily large amount, over an
arbitrarily wide range of frequencies. These variants essentially consist of a Michelson interferometer plus an
extra ‘sloshing’ cavity that sends the signal back into the interferometer with opposite phase shift, thereby
cancelling the position information and leaving a net phase shift proportional to the relative velocity. In
practice, the sensitivity of these variants will be limited by the maximum light power W
circ
circulating in the
arm cavities that the mirrors can support and by the leakage of vacuum into the optical train at dissipation
points. In the absence of dissipation and with squeezed vacuum power squeeze factor e
2R
0.1) inserted into
the output port so as to keep the circulating power down, the SQL can be beat by h/h
SQL
W
circ
SQL
e
2R
/W
circ
at all frequencies below some chosen f
opt
100 Hz. Here W
circ
SQL
800 kW(f
opt
/100 Hz)
3
is the power required to reach the SQL in the absence of squeezing. However, as the
power increases in this expression, the speed meter becomes more narrow band; additional power and reopti-
mization of some parameters are required to maintain the wide band. See Sec. III B. Estimates are given of the
amount by which vacuum leakage at dissipation points will debilitate this sensitivity see Fig. 12; these losses
are 10% or less over most of the frequency range of interest (f 10 Hz). The sensitivity can be improved,
particularly at high freqencies, by using frequency-dependent homodyne detection, which unfortunately re-
quires two 4-km-long filter cavities see Fig. 4.
DOI: 10.1103/PhysRevD.66.122004 PACS numbers: 04.80.Nn, 03.67.a, 42.50.Dv, 95.55.Ym
I. INTRODUCTION
This paper is part of the effort to explore theoretically
various ideas for a third-generation interferometric
gravitational-wave detector. The goal of such detectors is to
beat, by a factor of 5 or more, the standard quantum limit
SQL—a limit that constrains interferometers 1 such as
the first generation of the Laser Interferometric Gravitational
Wave Observatory LIGO-I which have conventional opti-
cal topology 2,3, but does not constrain more sophisticated
‘quantum nondemolition’ QND interferometers 4,5.
The concepts currently being explored for third-
generation detectors fall into two categories: external read-
out and intracavity readout. In interferometer designs with
external readout topologies, light exiting the interferometer
is monitored for phase shifts, which indicate the motion of
the test masses. Examples include conventional interferom-
eters and their variants such as LIGO-I 2,3, LIGO-II 6,
and those discussed in Ref. 7兴兲, as well as the speed-meter
interferometers discussed here and in a previous paper 8.In
intracavity readout topologies, the gravitational-wave force
is fed via light pressure onto a tiny internal mass, whose
displacement is monitored with a local position transducer.
Examples include the optical bar, symphotonic state, and op-
tical lever schemes discussed by Braginsky, Khalili, and
Gorodetsky 9–11. These intracavity readout interferom-
eters may be able to function at much lower light powers
than external readout interferometers of comparable sensitiv-
ity because the QND readout is performed via the local po-
sition transducer perhaps microwave-technology based, in-
stead of via the interferometers light; however, the designs
are not yet fully developed.
At present, the most complete analysis of candidate de-
signs for third-generation external-readout detectors has been
carried out by Kimble, Levin, Matsko, Thorne, and Vyatcha-
nin 7兴共KLMTV. They examined three potential designs for
interferometers that could beat the SQL: a squeezed-input
interferometer, which makes use of squeezed vacuum being
injected into the dark port; a variational-output scheme in
which frequency-dependent homodyne detection was used;
and a squeezed-variational interferometer that combines the
features of both. Because the KLMTV designs measure the
relative positions of the test masses, we shall refer to them as
position meters, particularly when we want to distinguish
them from the speed meters that, for example, use
variational-output techniques. Although at least some of the
KLMTV position-meter designs have remarkable perfor-
mance in the lossless limit, all of them are highly susceptible
to losses.
In addition, we note that the KLMTV position meters
each require four kilometer-scale cavities two arm cavities
two filter cavities. The speed meters described in this pa-
per require at least three kilometer-scale cavities two arm
cavitiesone ‘sloshing’ cavity described below兲兴.Ifwe
use a variational-output technique, as KLMTV did, the re-
sulting interferometer will have five kilometer-scale cavities
two arm cavities one sloshing cavitytwo filter cavities
again, see below兲兴. The speed meter described in
PHYSICAL REVIEW D 66, 122004 2002
0556-2821/2002/6612/12200424/$20.00 ©2002 The American Physical Society66 122004-1

this paper can achieve a performance significantly better than
a conventional position meter, as shown in Fig. 1. By ‘con-
ventional,’ we mean ‘without any QND techniques.’ An
example is LIGO-I. The squeezed-input speed meter SISM
noise curve shown in Fig. 1 beats the SQL by a factor of
10
in amplitude and has fixed-angle squeezed vacuum injected
into the dark port this allows the interferometer to operate at
a lower circulating power than would otherwise be necessary
to achieve that level of sensitivity, as described by Eq. 3
below. The squeezed-variational position meter SVPM,
which requires squeezed vacuum and frequency-dependent
homodyne detection, is more sensitive than the squeezed-
input speed meter over much of the frequency range of in-
terest, but the speed meter has the advantage at low frequen-
cies. It should also be noted that the squeezed-variational
position meter requires four kilometer-scale cavities as de-
scribed in the previous paragraph, whereas the squeezed-
input speed meter requires three.
If frequency-dependent homodyne detection is added to
the squeezed-input speed meter, the resulting squeezed-
variational speed meter SVSM can be optimized to beat the
squeezed-variational position meter over the entire frequency
range. Figure 1 contains two squeezed-variational speed
meter curves; one is optimized to match the squeezed-input
speed meter curve at low frequencies, and the other is opti-
mized for comparison with the squeezed-variational postion-
meter curve resulting in less sensitivity at high frequencies.
The original idea for a speed meter, as a device for mea-
suring the momentum of a single test mass, was conceived
by Braginsky and Khalili 12 and was further developed by
Braginsky, Gorodetsky, Khalili, and Thorne BGKT兲关13.In
their appendix, BGKT sketched a design for an interferomet-
ric gravity wave speed meter and speculated that it would be
able to beat the SQL. This was verified in Ref. 8兴共Paper I,
where it was demonstrated that such a device could in prin-
ciple beat the SQL by an arbitrary amount over a wide range
of frequencies. However, the design presented in that paper,
which we shall call the two-cavity speed-meter design, had
three significant problems: it required i a high circulating
power (8 MW to beat the SQL by a factor of 10 in noise
power at 100 Hz and below, ii a large amount of power
coming out of the interferometer with the signal 共⬃0.5 MW,
and iii an exorbitantly high input laser power
(300 MW). The latter two problems are effectively elimi-
nated by the alternate class of speed meters presented here—
designs that are based on the same QND mechanism de-
scribed in Refs. 8,12,13 but implemented by different
optical configurations. In addition, techniques for reducing
the needed circulating power are also discussed. These im-
provements bring interferometric speed meters into the realm
of practicality.
A simple version of the three-cavity speed-meter design to
be discussed in this paper is shown in Fig. 2. In an idealized
theorist’s version of this speed meter, the input laser light
with electric field denoted I(
) in Fig. 2 passes through a
power-recycling mirror into a standard Michelson interfer-
ometer. The relative phase shifts of the two arms are adjusted
so that all of the input light returns to the input port, leaving
the other port dark i.e., the interferometer is operating in the
symmetric mode so D(
)0 in Fig. 2. In effect, we have a
resonant cavity shaped like . When the end mirrors move,
they will put a phase shift on the light, causing some light to
enter the antisymmetric mode shaped like ) and come out
the dark port. So far, this is the same as conventional inter-
FIG. 1. Comparison of noise curves with losses of several
interferometer configurations. Each of these curves has been opti-
mized in a way that is meant to illustrate their relative advantages
and disadvantages. The conventional position meter CPM兲关7 has
W
circ
820 kW and bandwidth
cT/4L 2
100 Hz. The
squeezed-input speed meter SISM—optimized to agree with the
conventional position meter at high frequencies—has power
squeeze factor e
2R
0.1, optimal frequency
opt
2
105 Hz,
extraction rate
2
opt
, and sloshing frequency
3
opt
. The
squeezed-variational position meter SVPM兲关7 has the same pa-
rameters as the conventional position meter, with power squeeze
factor e
2R
0.1. There are two squeezed-variational speed-meter
curves SVSM.Oneblack dashes uses the same parameters as
the squeezed-input speed meter. The other solid curve has been
optimized to compare more directly with the squeezed-variational
position meter; it has 2
95 Hz and
2
100 Hz note
that our
is equivalent to the bandwidth
used to describe the
interferometers in Ref. 7兴兲.
FIG. 2. Simple version of three-cavity design for speed-meter
interferometer. The main laser input port is denoted by I(
), where
t z/c. The signal is extracted at the bottom mirror denoted
Q(
), where
t z/c]. The difference between the one- and
two-port versions is the mirror shown in gray.
P. PURDUE AND Y. CHEN PHYSICAL REVIEW D 66, 122004 2002
122004-2

ferometer designs but without the optical cavities in the two
interferometer arms.
Next, we feed the light coming out of the dark port
D(
)
into a sloshing cavity labeled K(
) and L(
)in
Fig. 2. The light carrying the position information sloshes
back into the ‘antisymmetric cavity’ with a phase shift of
180°, cancelling the position information in that cavity and
leaving only a phase shift proportional to the relative veloc-
ity of the test masses.
1
The sloshing frequency is
c
T
s
2L
, 1
where T
s
is the power transmissivity of the sloshing mirror, L
is the common length of all three cavities, and c is the speed
of light. We read the velocity signal
Q(
)
out at an extrac-
tion mirror with transmissivity T
o
), which gives a signal-
light extraction rate of
cT
o
L
. 2
We have used the extraction mirror to put the sloshing cavity
parallel to one of the arms of the Michelson part of the in-
terferometer, allowing this interferometer to fit into the ex-
isting LIGO facilities. The presence of the extraction mirror
essentially opens two ports to our system. We can use both
outputs, or we can add an additional mirror to close one port
the gray mirror in Fig. 2. We will focus on the latter case in
this paper.
The sensitivity h of this interferometer, compared to the
SQL, can be expressed as
2
h
h
SQL
W
circ
SQL
e
2R
W
circ
800 kW
e
2R
W
circ
, 3
where W
circ
is the power circulating in the arms, W
circ
SQL
800 kW(f
opt
/100 Hz)
3
is the power required to reach the
SQL in the absence of squeezing for the arms of length L
4 km and test masses with mass m 40 kg), and e
2R
is the
power squeeze factor.
3
With no squeezed vacuum, the
1
The net signal is proportional to the relative velocities of the test
masses, assuming that the frequencies
of the test masses’ motion
are
(sloshing frequency). However, the optimal regime of
operation for the speed meter is
⬃⍀. As a result, the output
signal contains a sum over odd time derivatives of position see the
discussion in Sec. III A. Therefore, the speed meter monitors not
just the relative speed of the test masses, but a mixture of all odd
time derivatives of the relative positions of the test masses.
2
It should be noted that, as the power increases in Eq. 3, the
speed-meter performance becomes more narrow band. Additional
power and a re-optimization of some of the speed meters param-
eters are required to maintain the same bandwidth at higher sensi-
tivities. See Sec. III B for details.
3
For an explanation of squeezed vacuum and squeeze factors, see,
for example, KLMTV and references cited therein. In particular,
their work was based on that of Caves 14 and Unruh 4. Also,
KLMTV state that a likely achievable value for the squeeze factor
in the LIGO-III time frame is e
2R
10, so we use that value in our
discussion.
FIG. 3. Schematic diagram showing the practical version of the
three-cavity speed-meter design, which reduces the power flowing
through the beam splitter. Three additional mirrors, with transmis-
sivity T
i
, are placed around the beam splitter. The ’’ and ‘‘ ’’
signs near the mirrors indicate the sign of the reflectivities in the
junction conditions for each location. The mirror shown in gray
closes the second port of the interferometer.
FIG. 4. Schematic diagram showing the practical three-cavity
speed-meter design with squeezed vacuum injected at the dark port
and two filter cavities on the output. Note that the circulator is a
four-port optical device that separates the injected squeezed input
and the interferometers output.
PRACTICAL SPEED METER DESIGNS FOR QUANTUM . . . PHYSICAL REVIEW D 66, 122004 2002
122004-3

squeeze factor is e
2R
1, so the circulating power W
circ
must
be 8 MW in order to beat the SQL at f
opt
100 Hz by a
factor of
10 in sensitivity. With a squeeze factor of e
2R
10, we can achieve the same performance with W
circ
800 kW, which is the same as LIGO-II is expected to be.
This performance in the lossless limit is the same as that
of the two-cavity Paper I speed meter for the same circu-
lating power, but the three-cavity design has an overwhelm-
ing advantage in terms of required input power. However,
there is one significant problem with this design that we must
address: the uncomfortably large amount of laser power,
equal to W
circ
, flowing through the beam splitter. Even with
the use of squeezed vacuum, this power will be too high.
This type of problem was addressed by Weiss and Drever,
who showed, respectively, that inserting optical delay lines
15 or Fabry-Pe
´
rot FP cavities 16 into the arms can
achieve a high circulating power with relatively low input
power at the beam splitter. In particular, using FP cavities in
the arms is now the standard design for most conventional
interferometers, such as LIGO-I. However, applying these
techniques alone will alter the propagation of the
gravitational-wave sidebands inside the interferometer and
jeopardize the performance of our speed meter. Fortunately,
there is a technique, based on the work of Mizuno 17 that
allows us to use FP cavities in the arms without affecting the
propagation of the sidebands. This method requires an addi-
tional mirror between the beam splitter and the extraction
mirror, placed such that light with the carrier frequency reso-
nates in the subcavity formed by this mirror and the arms’
internal mirrors. We shall call this design the practical three-
cavity speed meter; the three new mirrors are labeled T
i
in
Fig. 3.
As claimed by Mizuno 17 and tested experimentally by
Freise et al. 18 and Mason 19, when the transmissivity of
the third mirror decreases from 1, the storage time of side-
band fields in the arm cavity due to the presence of the in-
ternal mirrors will decrease. This phenomenon is called reso-
nant sideband extraction RSE; consequently, the third
mirror is called the RSE mirror. One special case, which is of
great interest to us, occurs when the RSE mirror has the same
transmissivity as the internal mirrors. In this case, the effect
of the internal mirrors on the gravitational-wave sidebands
should be exactly cancelled out by the RSE mirror. The three
new mirrors then have just one effect: they reduce the carrier
power passing through the beam splitter—and they can do so
by a large factor.
Indeed, we have confirmed that this is true for our speed
meter, as long as the distances between the three additional
mirrors the length of the ‘RSE cavity’ are small a few
meters, so that the phase shifts added to the slightly off-
resonance sidebands by the RSE cavity are negligible. We
can then adjust the transmissivities of the power-recycling
mirror and of the three internal mirrors to reduce the amount
of carrier power passing through the beam splitter to a more
reasonable level.
With this design, the high circulating power is confined to
the Fabry-Pe
´
rot arm cavities, as in conventional LIGO de-
signs. There is some question as to the level of power that
mirrors will be able to tolerate in the LIGO-III time frame.
Assuming that several megawatts is not acceptable, we shall
show that the circulating power can be reduced by injecting
fixed-angle squeezed vacuum into the dark port, as indicated
by Eq. 3.
Going a step farther, we shall show that if, in addition to
injected squeezed vacuum, we also use frequency-dependent
FD homodyne detection, the sensitivity of the speed meter
is dramatically improved at high frequencies above f
opt
100 Hz); this is shown in Fig. 1. The disadvantage of this
is that FD homodyne detection requires two filter cavities of
the same length as the arm cavities 4 km for LIGO,as
shown in Fig. 4.
Our analysis of the losses in these scenarios indicates that
our speed meters with squeezed vacuum and/or variational-
output are much less sensitive to losses than a position meter
using those techniques as analyzed by KLMTV. Losses for
the various speed meters we discuss here are generally quite
low and are due primarily to the losses in the optical ele-
ments as opposed to mode-mismatching effects. Without
squeezed vacuum, the losses in sensitivity are less than 10%
in the range 50105 Hz, lower at higher frequencies, but
higher at low frequencies. Injecting fixed-angle squeezed
vacuum into the dark port allows this speed meter to operate
at a lower power see Eq. 3兲兴, thereby reducing the domi-
nant losses which are dependent on the circulating power
because they come from vacuum fluctations contributing to
the back action. In this case, the losses are less than 4% in
the range 25150 Hz. As before, they are lower at high fre-
quencies, but they increase at low frequencies. Using FD
homodyne detection does not change the losses significantly.
This paper is organized as follows: In Sec. II we give a
brief description of the mathematical method that we use to
analyze the interferometer. In Sec. III A, we present the re-
sults in the lossless case, followed in Sec. III B by a discus-
sion of optimization methods. In Sec. III C, we discuss some
of the advantages and disadvantages of this design, including
the reasons it requires a large circulating power. Then in Sec.
IV, we show how the circulating power can be reduced by
injecting squeezed vacuum through the dark port of the in-
terferometer and how the use of frequency-dependent homo-
dyne detection can improve the performance at high frequen-
cies. In Sec. V, we discuss the effect of losses on our speed
meter with the various modifications made in Sec. IV, and we
compare our interferometer configurations with those of
KLMTV. Finally, we summarize our results in Sec. VI.
II. MATHEMATICAL DESCRIPTION
OF THE INTERFEROMETER
The interferometers in this paper are analyzed using the
techniques described in Paper I Sec. II. These methods are
based on the formalism developed by Caves and Schumaker
20,21 and used by KLMTV to examine more conventional
interferometer designs. For completeness, we will summa-
rize the main points here.
The electric field propagating in each direction down each
segment of the interferometer is expressed in the form
P. PURDUE AND Y. CHEN PHYSICAL REVIEW D 66, 122004 2002
122004-4

E
field
4
0
Sc
A
. 4
Here A(
) is the amplitude which is denoted by other
letters—B(
), P(
), etc.—in other parts of the interferom-
eter; see Fig. 2,
t z/c,
0
is the carrier frequency, is
the reduced Planck’s constant, and S is the effective cross-
sectional area of the light beam; see Eq. 8 of KLMTV. For
light propagating in the negative z direction,
t z/c is
replaced by
t z/c. We decompose the amplitude into
cosine and sine quadratures,
A
A
1
cos
0
A
2
sin
0
, 5
where the subscript 1 always refers to the cosine quadrature,
and 2 to sine. Both arms and the sloshing cavity have length
L4 km, whereas all of the other lengths z
i
are short com-
pared to L. We choose the cavity lengths to be exact half
multiples of the carrier wavelength so e
i2
0
L/c
1 and
e
i2
0
z
i
/c
1. There will be phase shifts put onto the sideband
light in all of these cavities, but only the phase shifts due to
the long cavities are significant.
The aforementioned sidebands are put onto the carrier by
the mirror motions and by vacuum fluctuations. We express
the quadrature amplitudes for the carrier plus the sidebands
in the form
A
j
A
j
0
a
˜
j
e
i
␻␨
a
˜
j
e
i
␻␨
d
2
. 6
Here A
j
(
) is the carrier amplitude, a
˜
j
(
) is the field ampli-
tude a quantum mechanical operator for the sideband at
sideband frequency
absolute frequency
0
) in the j
quadrature, and a
˜
j
(
) is the Hermitian adjoint of a
˜
j
(
); cf.
Eqs. 68 of KLMTV, where commutation relations and
the connection to creation and annihilation operators are dis-
cussed. In other portions of the interferometer Fig. 2, A
j
(
)
is replaced by, e.g., C
j
(
); A
j
(
), by C
j
(
); a
˜
j
(
), by
c
˜
j
(
), etc.
Since each mirror has a power transmissivity and comple-
mentary reflectivity satisfying the equation TR 1, we can
write out the junction conditions for each mirror in the sys-
tem, for both the carrier quadratures and the sidebands see
particularly Eqs. 5 and 1214 in Paper I. We shall de-
note the power transmissivities for the sloshing mirror as T
s
,
for the extraction output mirror as T
o
, the power-recycling
mirror as T
p
, for the beam-splitter as T
b
0.5, for the inter-
nal mirrors as T
i
, and for the end mirrors as T
e
; see Figs. 2
and 3.
The resulting equations can be solved simultaneously to
get expressions for the carrier and sidebands in each segment
of the interferometer. Since those expressions may be quite
complicated, we use the following assumptions to simplify
our results. First, we assume that only the cosine quadrature
is being driven so that the carrier sine quadrature terms are
all zero. Second, we assume that the transmissivities obey
1T
o
T
s
T
e
and 1
T
p
,T
i
T
e
. 7
The motivations for these assumptions are that i they lead
to speed-meter behavior; ii as with any interferometer, the
best performance is achieved by making the end-mirror
transmissivities T
e
as small as possible; and iii good per-
formance requires a light extraction rate comparable to the
sloshing rate,
⬃⍀ cf. the first paragraph of Sec. III B in
Paper I, which with Eqs. 1 and 2 implies T
o
T
s
so
T
o
T
s
. Throughout the paper, we will be using these as-
sumptions, together with
L/c 1, to simplify our expres-
sions.
III. SPEED METER IN THE LOSSLESS LIMIT
For simplicity, in this section we will set T
e
0 end mir-
rors perfectly reflecting. We will also neglect the vacuum-
fluctuation noise coming in the main laser port (i
˜
1,2
) since
that noise largely exits back toward the laser and produces
negligible noise on the signal light exiting the output port. As
a result of these assumptions, the only vacuum-fluctuation
noise that remains is that which comes in through the output
port (p
˜
1,2
). An interferometer in which this is the case and in
which light absorption and scattering are unimportant (R
T 1 for all mirrors, as we have assumed is said to be
‘lossless.’ In Sec. V, we shall relax these assumptions; i.e.,
we shall consider lossy interferometers.
It should be noted that the results and discussion in this
section and in Sec. IV apply to both the simple and practical
versions of the three-cavity speed meter Figs. 2 and 3. The
two versions are completely equivalent in the lossless limit.
A. Mathematical analysis
The lossless interferometer output for the speed meters in
Fig. 2 and 3, as derived by the analysis sketched in the pre-
vious section, is then
q
˜
1
⫽⫺
L
*
L
p
˜
1
, 8a
q
˜
2
2i
0
W
circ
cLL
x
˜
L
*
L
p
˜
2
. 8b
Here p
˜
j
(
) is the side-band field operator analogue of
a
˜
j
(
) in Eq. 6兲兴 associated with the dark-port input P(
),
and q
˜
j
(
) associated with the output Q(
); see Fig. 2. Also,
in Eqs. 8, L(
)isac number given by
L
2
2
i
9
recalling that c
T
s
/2L is the sloshing frequency,
cT
o
/L the extraction rate, the asterisk in L
*
(
) denotes
the complex conjugate, x
˜
(
) is the Fourier transform of the
relative displacement of the four test masses—i.e., the Fou-
rier transform of the difference in lengths of the interferom-
eters two arm cavities—and W
circ
is circulating power in the
each of the interferometers two arms. Note that the circulat-
PRACTICAL SPEED METER DESIGNS FOR QUANTUM . . . PHYSICAL REVIEW D 66, 122004 2002
122004-5

Figures
Citations
More filters
Journal ArticleDOI

Squeezed states of light and their applications in laser interferometers

TL;DR: In this article, the authors revisited the concept of squeezed states and two-mode squeezed states of light, with a focus on experimental observations, and the distinct properties of the two modes of light displayed in quadrature phase-space as well as in the photon number representation are described.
Journal ArticleDOI

Gravitational Radiation Detection with Laser Interferometry

TL;DR: Gravitational-wave detection has been pursued relentlessly for over 40 years as mentioned in this paper, and with the imminent operation of a new generation of laser interferometers, it is expected that detections will become a common occurrence.
Journal ArticleDOI

Macroscopic quantum mechanics: theory and experimental concepts of optomechanics

TL;DR: In this article, the authors review a set of techniques of quantum measurement theory that are often used to analyse quantum optomechanical systems, including quantum entanglement, quantum teleportation and quantum Zeno effect.
Journal ArticleDOI

Squeezing in the audio gravitational-wave detection band.

TL;DR: It is shown that low frequency noise sources, such as seed noise, pump noise, and detuning fluctuations, present in optical parametric amplifiers, have negligible effect on squeezing produced by a below-threshold OPO.
Journal ArticleDOI

Continuous force and displacement measurement below the standard quantum limit

TL;DR: In this article, strong quantum correlations in an ultracoherent optomechanical system were exploited to demonstrate off-resonant force and displacement sensitivity reaching 1.5 dB below the standard quantum limit.
References
More filters
Journal ArticleDOI

Quantum Optics, Experimental Gravitation, and Measurement Theory

TL;DR: In this paper, Wigner's friends and their amnesia are described and compared with the information transfer in Quantum Measurements: Irreversibility and Amplification, and Information Transfer of Quantum Measurement: Information Transfer in Quantum Information Transfer and Non-Demolition Measurements.
Proceedings ArticleDOI

Energetic Quantum Limit in Large-Scale Interferometers

TL;DR: For each optical topology of an interferometric gravitational wave detector, quantum mechanics dictates a minimum optical power (the "energetic quantum limit") to achieve a given sensitivity as mentioned in this paper.
Related Papers (5)
Frequently Asked Questions (12)
Q1. What are the contributions mentioned in the paper "Practical speed meter designs for quantum nondemolition gravitational-wave interferometers" ?

This paper describes some practical variants of speed-meter interferometers. 

In fact, K* can be obtained from the speed meter k* by putting V→0 and d→g .-16exorbitantly high input laser power required by the twocavity speed meter. 

The dominant sources of loss-induced noise at low frequencies ( f & f opt) are the radiation-pressure noise from losses in the arm, extraction, and sloshing cavities. 

the circulator by which the squeezed vacuum is injected; ~iv! and the external filter cavities used for the variational-output scheme. 

While it is unfortunate that losses limit the performance of interferometers, the speed meter is at least able to retain a wide-band sensitivity even in the presence of a loss limit. 

The reason for this high power is the energetic quantum limit ~EQL!, which was first derived for gravitational-wave interferometers by Braginsky, Gorodetsky, Khalili and Thorne @22#. 

~43!Since k is proportional to the circulating power @see Eqs. ~14!#, gaining a factor e2R in k is equivalent to gaining this factor in Wcirc . 

As a result, a position meter optimized at some frequency f opt may be able to reach its ‘‘loss limit’’ ~the equivalent of ShL) at that frequency f opt , but doing so will result in a sharp growth of noise at frequencies below f opt . 

noise coming in the main laser port ( ĩ 1,2) since that noise largely exits back toward the laser and produces negligible noise on the signal light exiting the output port. 

that will dramatically improve the performance of the speed meter at frequencies f * f opt .A. Injection of squeezed vacuum into dark portBecause the amount of circulating power required by their speed meter remains uncomfortably large, it is desirable to reduce it by injecting squeezed vacuum into the dark port. 

In particular, the noise curve for the speed meter with Wcirc5800 kW ~and f opt5107 Hz) matches the curve of the conventional position meter at high frequencies, while it beats the SQL by a factor of ;8 ~in power! 

A more conventional solution for their externalreadout interferometer is to inject squeezed light into the dark port, as the authors shall discuss in Sec. IV A ~and as was also discussed in the original paper @22# on the EQL!.