Sunteți pe pagina 1din 88

Testing General Relativity with

Spherical Resonant Mass Detectors

A Thesis

Presented to the

Graduate Faculty of the

University of Louisiana at Lafayette

In Partial Fulfillment of the

Requirements for the Degree

Master of Science

Alex J. Sylvester

Fall 2015
ProQuest Number: 10002423

All rights reserved

INFORMATION TO ALL USERS


The quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if material had to be removed,
a note will indicate the deletion.

ProQuest 10002423

Published by ProQuest LLC (2016). Copyright of the Dissertation is held by the Author.

All rights reserved.


This work is protected against unauthorized copying under Title 17, United States Code
Microform Edition ProQuest LLC.

ProQuest LLC.
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, MI 48106 - 1346

c Alex J. Sylvester

2015

All Rights Reserved


Testing General Relativity with
Spherical Resonant Mass Detectors

Alex J. Sylvester

APPROVED:

James B. Dent, Chair Natalia A. Sidorovskaia


Assistant Professor of Physics Professor of Physics

Gabriele Morra Mary Farmer-Kaiser


Assistant Professor of Physics Dean of the Graduate School
ACKNOWLEDGMENTS

I would like to thank the UL-Lafayette Physics department for making my time spent

with the University so personally rewarding. I would also like to thank the members of my

committee and the Graduate School for your patience and helpful critism. Finally, I would

like to thank my advisor James Dent for helping me make this manuscript possible and for

teaching me GR twice.
TABLE OF CONTENTS

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

CHAPTER 1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

CHAPTER 2: Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

CHAPTER 3: The Basics of General Relativity . . . . . . . . . . . . . . . . . . 9


3.1. Mathematical Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2. Principle of Equivalence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.3. Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3.1. The Covariant Derivative . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3.2. The Riemann Curvature Tensor . . . . . . . . . . . . . . . . . . . . . . 20
3.4. Einstein Field Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

CHAPTER 4: Gravitational Waves . . . . . . . . . . . . . . . . . . . . . . . . . . 27


4.1. Basics of Gravitational Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.2. Energy and Momentum of Gravitational Waves . . . . . . . . . . . . . . . . . 31
4.3. Sources of Gravitational Waves . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.4. Detection of Gravitational Waves . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.4.1. Hulse-Taylor Observation . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.4.2. Detectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.4.3. GW Strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.5. Tidal Force on a Detector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

CHAPTER 5: Modified Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46


5.1. GWs in f (R) Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.2. Energy-Momentum of GWs in f (R) gravity . . . . . . . . . . . . . . . . . . . 48

CHAPTER 6: Acoustic Response of a Spherical Detector . . . . . . . . . . . . 53


6.1. Mathematical Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.2. GW Tidal Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.3. Spherical Detector Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
6.4. Absorption Cross Section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.4.1. GR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.4.2. f (R) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.5. Observational Possibilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6.5.1. Minimum Strain for the Mario Schenberg Detector . . . . . . . . . . . 68
6.5.2. Scalar Wave Mass Constraints . . . . . . . . . . . . . . . . . . . . . . . 70

CHAPTER 7: Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

vi
LIST OF TABLES

Table 6.1. Mario Schenberg detector parameters [33] . . . . . . . . . . . . . . . . . 69


LIST OF FIGURES

Figure 3.1: A vector parallel transported along a closed loop on a sphere. . . . . . 21

Figure 4.1: A passing gravitational wave displaces a circular ring of test masses. The
top ring shows the perturbation caused by h11 . The bottom ring is perturbed
by h12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

Figure 4.2: Hulse-Taylor Results[27]: graph of the change in time to reach periastron
(the point where orbiting bodies are closest to each other) of PSR 1913+16 vs
time. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

Figure 4.3: Incident GW perpendicular to the plane of the interferomter. The test
masses are low-transmissivity mirrors that provide optical cavities to increase
the travel distance of the beam. . . . . . . . . . . . . . . . . . . . . . . . . . . 40

Figure 4.4: Detectable GW amplitude strain vs GW frequency[28] . . . . . . . . . 42

Figure 6.1: Displacement u of a mass element on the surface of a resonating sphere.


The dark circle represents the cross section of the sphere in equilibrium, while
the gray represents the sphere in a perturbed state. . . . . . . . . . . . . . . . 54

Figure 6.2: Quadrupolar modes of a sphere[33]. The dark regions signify the max-
imal radial perturbation of the sphere from a GW incident with the zenith
direction of the sphere. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
CHAPTER 1: Introduction

This year of 2015 marks the 100th anniversary of Einsteins General Theory of

Relativity, and since its birth, the theory has passed every test. Experimental evidence such

as the deflection of light by the sun, measuring the perihelion shift of the inner planets, and

observing gravitational redshift have instilled the utmost confidence of the astrophysical

communtity [1].

General relativity (GR) has made some very interesting predictions such as time

dilation, gravitational lensing, black holes, and the expansion of the universe. GR also

predicts the existence of gravitational waves, traveling fluctuations of the gravitational field.

Gravitational waves have not been directly observed, but there are ongoing programs such

as LIGO and VIRGO where interferometers have been built several kilometers long to

detect changes of an interferometers arm length by one part in 1023 . Once detected, we will

have a new spectrum of radiation to study that is complementary to and completely distinct

from electromagnetic radiation [2, 3].

According to GR, gravitational waves (GWs) travel at the speed of light and are

emitted through a quadrupole moment of a gravitating system, similarly to how

electromagnetic radiation is emited through a dipole moment of a charged system.

Electromagnetic radiation is emitted by charged particles like electrons, and studying EM

waves has taught us a great deal about its microscopic sources. GWs are emitted by massive

celestial objects, and detection of these waves would allow humanity to study gravity in the

strong, highly relativistic and non-linear limit. Detection of GWs are expected within the

next few years via the beam detectors of LIGO and VIRGO, and with detection comes new

ways to observe the universe [2, 3].


The first attempts to dectect GWs were by building resonant mass detectors that

matched the resonance frequency of the detector with the expected frequency of an incident

GW[1, 4]. These detectors have a huge disadvantage compared to beam detectors in that

the range of of detectable frequencies are limited to the specific resonance frequencies of the

detectors. One thing that beam detectors and conventional resonance detectors have in

common is that the direction of an incident wave affects its detectability. There has been a

recent push to build resonance detectors with a spherical shape as opposed to their

conventional cylindrical shape [4, 5]. Spherical detectors have the advantage of being

indifferent to the direction of an incident wave. This allows scientists to not only detect

GWs, but also measure the polarizations of the wave with a single detector, while

conventional resonance and beam detectors would require multiple detectors to measure the

waves polarizations. Once frequencies of detectable GW sources are known, detectors can

be built to match these frequencies, and because resonant mass detectors have an

economical advantage over their interferometer counterparts, it will provide a less expensive

way for people to build GW observatories.

Being able to measure the polarizations of a gravitational wave would provide a very

good test of general relativity versus modified theories of gravity, since modified gravity

typically introduces new polarization modes. As stated earlier GR has passed every test,

however since its birth people have been devising new experimental means of testing the

theory, as well as looking to extend or modify it theoretically [2, 6]. Recent motivation to

look for modified theories of GR comes from the observation of the universe expanding at an

accelerated rate [6, 7, 8, 9, 10, 11, 12, 13]. The simplest cosmological model to explain this

accelerated expansion from within the confines of GR is to introduce a cosmological

2
constant into the field equations of GR. The main problem with this solution is that the

measured value of this constant is 10120 times smaller than quantum field theory predicts [6].

A very popular branch of modified gravity theories known as f (R) gravity generalizes

the Lagrangian of GR and hence the field equations of GR [6]. This generalization of the

Lagrangian gives rise to an additonal field of gravitation, and when looking at GWs in f (R)

this additional field manifests itself as a scalar mode. f (R) has become a theoretical test

bed for modifications of GR, from cosmological to solar system scales [6].

Gravitational waves in f (R) gravity differ from GR in that an f (R) GW has a scalar

piece that is massive (and therefore does not travel at the speed of light)[14]. This

manuscript endeavors to calculate how the massive scalar mode of a f (R) gravitational wave

would affect a spherical detector. It begins by attempting to familiarize the reader with the

basic ideas and mathematical notation of general relativity and gravitational waves, and

then show how spherical resonance detectors can be used to observe the nature of GWs.

The next chapter contains a glossary of the main quantities of general relativity and

gravitational waves and how to express them mathematically. Chapter 3 begins with a

review of the mathematical notation used in this document. It then shifts the discussion to

the principle of equivalence and how it is the foundation of general relativity. The chapter

concludes with brief reviews of the Riemann tensor and the Einstein field equations.

Chapter 4 deals with the basics of gravitational wave physics and current searches for

GWs. Chapter 5 begins with a review of f (R) GWs and concludes with the calculation of

the energy and momentum carried by an f (R) GW that has been derived independently.

Chapter 6 details the acoustic response of the spherical resonant mass detector

perturbed by a GW, and then derives how an f (R) GW affects the detector differently than

3
a wave in GR. The end of the chapter calculates the theoretically smallest scalar wave

amplitude that the Mario Schenberg spherical detector can observe, and studies how the

mass of a scalar wave provides a test of cosmological models.

The main goals of this document are to review the basic concepts of general relativity

and gravitational wave physics and to apply those ideas by calculating the effects an f (R)

GW would have on a spherical detector.

4
CHAPTER 2: Glossary

This document is written using natural units, which is the standard for particle

physics and cosmology. That is

c = kB = ~ = 1 , (2.1)

where c is the speed of light, kB is the Boltzmann constant, and ~ is the reduced Planck

constant. In natural units, most quantities are written in terms of electron volts (eV) and

have the following conversions from S.I. units,

1m = 5.08 106 eV1 length is measured in 1/energy

1s = 1.52 1015 eV1 time is measured in 1/energy

1kg = 5.62 1035 eV mass is measured in energy

Four-vector notation will be used with Greek indices (, ,..) running from 0 3, with 0

the temporal component and 1, 2, and 3 the spatial components. Latin indices (i, j, ...) run

through only the spatial components.

Unless otherwise noted, the Einstein summation convention will be employed -

subscripts and superscripts with the same symbol are summed over their components, i.e.
3
X

A B = A B = A0 B 0 + A1 B 1 + A2 B 2 + A3 B 3 . (2.2)
=0

The Minkowski metric tensor will follow a mostly positive signature


1 0 0 0
0 1 0 0
=
0
. (2.3)
0 1 0
0 0 0 1
Indices can be raised and lowered with the metric

T = 0 T 0 + 1 T 1 + 2 T 2 3 T 3 = T . (2.4)
The metric has unique properties when its indices are raised or contracted. When an index

is raised, it takes the form of the Kronecker delta


(
0 6=
=
= 0 0
+ 1 1
+ 2 2
+ 3 3
= . (2.5)
1 =

The contraction of the metric is

= 0 0 + 1 1 + 2 2 + 3 3 , (2.6)

where

0 0 = 1 1 = 2 2 = 3 3 = 30 30 + 31 31 + 32 32 + 33 33 = 1 , (2.7)

so the contraction of the metric is simply

= 4 . (2.8)

The Minkowski metric is used when one is working in flat space like in special relativity.

The metric tensor g must be used when one is working in an arbitrarily curved space,

and it has all the properties shown above for the Minkowski metric tensor. The metric is a

dimensionless quantity that is used to describe conventional gravitational potential.

The metric commonly used for gravitational waves is the Minkowski metric with a

small perturbation h

g = + h , (2.9)

The perturbation h is where the GW manifests itself

h = e exp(ik x ) + e exp(ik x ) , (2.10)

6
where e is the polarization tensor of the wave and k is the wave vector.

The affine connection is written in terms of the metric

1
= g ( g + g g ) , (2.11)
2

with = /x . It contains derivatives of the metric, and therefore its units are in energy.

The Riemann tensor R is defined,

R + . (2.12)

The Ricci tensor R is written as the contraction of the Riemann tensor,

R = R = g R = g g R . (2.13)

The Ricci scalar R is the contraction of the Ricci tensor,

R = g R (2.14)

The Riemann tensor, Ricci tensor, and Ricci scalar contain terms with second derivatives of

the metric and quadratic terms of first derivatives of the metric, so their units are energy

squared (eV2 ).

The Einstein field equations for general relativity are written as

G = 8GT , (2.15)

where G is known as the Einstein Tensor and is written in terms of the Ricci tensor and

scalar

1
G = R g R , (2.16)
2

7
and G is Newtons constant (and not the contraction of the Einstein tensor G = g G )

N m2
G = 6.74 1011 = 6.79 1057 eV2 . (2.17)
kg2

T in equation (2.15) is the energy-momentum tensor that is defined based on the

physical system one is working in, and its units are eV4 . It is a common practice in GR to

model spacetime as a perfect fluid in thermodynamic equilibrium. In this case, T takes the

form

dx dx
T = ( + p) + g p , (2.18)
d d

where and p are the mass-energy density and pressure of spacetime, respectively, and

dx /d is the four-velocity vector.

8
CHAPTER 3: The Basics of General Relativity

3.1. Mathematical Background

Many quantities in General Relativity (GR) are described as tensors, where tensors

are generalizations of vectors. Scalars are tensors of rank zero; while vectors are tensors of

rank one. Tensors of rank two can be thought of as matrices. The indices of vectors and

tensors span the 4 dimensions of spacetime. A position vector x can then represent

Cartesian coordinates,

x0 = t

x1 = x

x2 = y

x3 = z,

where the indices 0, 1, 2, and 3 are superscripts. In S.I. units x0 = ct, but the factor of c

disappears when working in natural units. Any superscript / subscript described with a

Greek character is understood to span all four dimensions of spacetime, while Latin indices

will describe only space. Any quantity written in bold (for example x) represents a standard

three dimensional vector. The vector x can also be expressed in its covariant form x

x = x . (3.1)

is the the Minkowski metric tensor which is used when one is working in flat Cartesian

space.

The metric tensor g describes how spacetime is curved and is used in an arbitrarily

curved space. It is often necessary to shift quantities into their covariant (when the index is
a subscript) or contravariant (when the index is a superscript) forms. To do this, one needs

to compensate for the curvature of spactetime, so it is necessary to use the metric tensor to

raise and lower indices. The components of the Minkowski metric tensor are written in a

44 matrix,


1 0 0 0
0 1 0 0
=
0
. (3.2)
0 1 0
0 0 0 1

Repeated subscripts and superscripts in a term are referred to as dummy indices

and are understood to be summed over. Their symbols can be changed from term to term.

Indices not repeated are referred to as free. The dot product of two vectors A and B in

this notation is written

A B = A B = A B = A0 B 0 + A1 B 1 + A2 B 2 + A3 B 3 (3.3)

A spacetime interval ds2 is written in Minkowski space as

ds2 = dx dx = dt2 + dx2 + dy 2 + dz 2 . (3.4)

In the frame moving with a clock in motion, the clocks spatial displacement is zero

(dx2 + dy 2 + dz 2 = 0). This leads us to define proper time

d 2 ds2 (3.5)

The proper time is the measure of the time elapsed on a clock carried along the spacetime

interval.

10
0
One can shift coordinates x x with a Lorentz transformation

0
0 x
x = x , (3.6)
x

0
where x /x is the Lorentz tranformation matrix. General tensors can be transformed in

this way, for example a rank-3 tensor T transforms like

0 0
0 0 x x x
T 0 = T (3.7)
x x x 0

Einstein developed the mechanics of what is now known as special relativity based on

two postulates. The first being the speed of light in a vacuum is the same in all reference

frames; the second is that the laws of physics are the same in all inertial reference frames.

Lorentz transformations are used in special relativity to describe how basic laws of mechanics

look in inertial reference frames. However, there is a problem extending relativistic

mechanics to gravitation because classical gravity does not correctly change under a Lorentz

transformation of coordinates. Thus, Einstein endeavored to extend relativity to gravitation,

and the result was general relativity. Just as special relativity has its two postulates, general

relativity is rooted in what is known as the Principle of Equivalence[1, 15] .

3.2. Principle of Equivalence

The Principle of Equivalence is based on the idea that inertial mass mi from Newtons

2nd law,

F = mi a , (3.8)

is the same mass that gravitates mg from Newtons law of gravitation,

GMr
F = mg . (3.9)
r2

11
Newton had suspected mi = mg , but there were no conclusive results to prove this until the

late 19th century[1]. Einstein used this idea to come up with the following thought

experiment. Imagine a passenger standing in a closed elevator with no windows. The

elevator is hanging from a high cliff by a wire. While the elevator is hanging, the passenger

feels Earths gravity pulling him to the floor of the elevator. Now imagine the wire being

cut, allowing the elevator to fall freely. Just like astronauts in orbit around Earth, the

passenger is in free fall (an inertial reference frame) and no longer feels the effect of Earths

gravitational field (for the moment, let us ignore the positional dependence of Earths

gravitational field). Now imagine the same elevator scenario in deep empty space. The

passenger is still in an inertial frame, so as long as he is in the elevator, he cannot tell the

difference between falling towards Earth and being in deep space. Finally we attach a rocket

to the bottom of the elevator, and the thrust of the engine accelerates the elevator at 9.81

m/s2 . According to the Principle of Equivalence, the passenger will not be able to tell the

difference between the inertial force felt from the rocket in deep space, and the force of

gravity felt hanging from a wire on Earth.

In the elevator thought experiment we had to ignore the fact that Earths gravitational

field varies with distance from Earths center of mass. In reality we cannot expect inertial

forces to cancel gravitational forces for free falling frames in gravitational fields that are

space and time dependent. We can assume an approximate cancellation if we limit our view

to a very small region of spacetime. There are different forms of the principle, and I will

now use Steven Weinbergs statement: at every spacetime point in an arbitrary gravitational

field, it is possible to choose a locally inertial coordinate system such that, within a

sufficiently small region of the point in question, the laws of nature take the same form as in

12
unaccelerated coordinate systems in the absence of gravitation. [1]

Lets go back to the elevator falling towards Earth. Ignoring forces like air friction, the

elevator and passenger are moving freely through a gravitational field. Now putting the

Equivalence Principle to use, we can observe it in a locally interial frame from the
0
perspective of the passenger inside the elevator, which we will call x . In this inertial frame,

its equation of motion is a straight line in spacetime

0
d 2 x
=0 , (3.10)
d 2

where d is the proper time

d 2 = dx0 dx0 . (3.11)

Lets now leave the elevator, and observe the elevator falling while we stand safely

somewhere on the surface of Earth in some arbitrary coordinate system x . We can model

the inertial frames coordinates as a function of our own. That is,

0 0
x = x (x ). (3.12)

From this we can see

0
0 x
dx = dx , (3.13)
x

and the proper time can be expressed in these arbitrary coordinates

x 0 x 0
d 2 = 0 dx 0 dx = g dx dx (3.14)
x x

13
where g is the metric tensor and is defined
0 0
x x
g . (3.15)
x x

The metric tensor is the most important tensor in GR. It contains all the information

about the geometry of spacetime, and the Einstein field equations which describe the

dynamics of a gravitational field are just a set of differential equations that relate the

components of the metric tensor to the energy-momentum tensor of the system under

consideration.

Equation (3.4) uses the Minkowski metric to write the Pythagorean theorem in

four dimensional Cartesian spacetime. If the local spacetime is curved, (3.4) does not hold.

The metric tensor g allows us to write the Pythagorean theorem in an arbitrarily curved

spacetime.

Tensors are the main items of interest in GR because tensor equations obey what is

called the Principle of General Covariance, which allows us to put the equivalence principle

to work mathematically. Weinberg defines the Principle of General Covariance as follows: a

physical equation holds in a gravitational field if two conditions are met: 1) The equation

holds in the absence of gravitation, and 2) the equation is generally covariant. This means

that if an equation between two tensors holds in one coordinate system, it holds in all

coordinate systems[1, 15].

The elevators aforementioned freely moving path can also be described in our

non-inertial coordinate system on Earth

0 0 0 0
d 2 x x dx 2 x dx dx x d2 x
 
d
2
= = + =0 , (3.16)
d d x d x x d d x d 2

From here the fact that

14
0
x x
= , (3.17)
x x0

can be used to write the straight path as

0
d 2 x x 2 x dx dx
+ =0 . (3.18)
d 2 x0 x x d d

We define

0
x 2 x
(3.19)
x0 x x

to be the affine connection, and what is left is the geodesic equation:

d2 x
dx dx

+ =0 . (3.20)
d 2 d d

Geodesics can be thought of as generalizations of straight lines in curved space. More

accurately, geodesics are extremal paths in curved spacetime, and for timelike paths, they

maximize the proper time. In GR, freely moving objects like our elevator move along

geodesics. The geodesic equation can be thought of as the GR version of Newtons 2nd law,

when F = ma = 0.

This is where Newtonian physics deviates from GR. From a Newtonian point of view,

the elevator is moving while being under the influence of a gravitational force. According to

GR, the elevator is simply moving along the curvature of spacetime. When the wire is cut,

it begins its descent because the geometry of spacetime is curved due to the mass of Earth.

This is similar to when a ball is released from the side of a hill; it follows the hills shape

towards the bottom.

15
GR contains Newtonian gravity as a limiting case. For example, it can be shown that

the geodesic equation can reduce to the Newtonian gravitational acceleration

d2 x
= , (3.21)
dt2

where is the gradient and is the scalar potential of the gravitational field

GM
= . (3.22)
|x|

It should be noted that (3.20) is a fully relativistic equation. To reduce it to classical

gravity, we must take non-relativistic limits. In this limit, the field is static, and the spatial

velocities are very small compared to the speed of light. Thus we have

dx0 dt dxi
= 1 , (3.23)
d d d

and (3.20) becomes

d 2 x
= 00 . (3.24)
dt2

In GR, the affine connection is a function of the metric tensor

1
= g ( g + g g ) , (3.25)
2

where = /x . Because the field is stationary, all time derivatives in the metric must

vanish. We adopt a Minkowski metric with a very small pertubation h , so that we are

only interested in to the first order in h (this approximation holds for Earth, the Sun,

and any other celestial object in the solar system).

16
g = + h |h |  1 . (3.26)

Combining equations (3.25) and (3.26) one obtains

d2 x 1
2
= h00 , (3.27)
dt 2

where h00 = 2.

3.3. Curvature

3.3.1. The Covariant Derivative

In physics it is necessary to study the dynamics of quantities, and in GR these

quantities are tensors. This means one must be able to take derivatives of tensors. However,

there arises a problem when one takes an ordinary derivative of a tensor. Consider a vector

V . Because vectors are tensors, we have a transformation like one would expect for a rank

one tensor,

0
0 x
V = V . (3.28)
x
0
Taking the derivative of V , we see the problem,

 0  0 0
0 x x x V x 2 x
0 V = V = + 0 V (3.29)
x0 x x0 x x x x x

Tensors must transform like equation (3.7), and looking at the last expression in
0
(3.29), the first term is what one would expect to see if 0 V was a tensor. However, the

last term shows that it is not. This problem can be fixed by introducing a new type of

derivative that will hold all the required tensor properties. The last term looks a lot like the

17
affine connection defined in (3.25), and by including the affine connection in our new

derivative, the last term in (3.29) is cancelled, resulting in a proper tensor transformation.

Thus for a vector V , its covariant derivative is written as

V = V + V , (3.30)

where is the symbol for the covariant derivative operator. We can use this derivative for

a tensor of any number of indices. For a scalar we find that it is simply the standard

partial derivative,

= . (3.31)

For a tensor like in equation (3.7), we have

T = T + T + T T . (3.32)

The covariant derivative has two important properties. The first is the fact that it

converts tensors to other tensors as has been shown. The second is that in the absence of

gravitation, = 0, and it reduces to ordinary differentiation. Thus one can take any

relation in special relativity, replace with g , and with , and one finds that the

relation will still be true in the presence of gravitational fields as long as one works in a

spacetime scale that is sufficiently small compared to the gravitational field. [1]

As an example consider Maxwells equations

B=0 (3.33)

18
t B + E = 0 (3.34)

E= (3.35)

t E + B = J , (3.36)

where E and B are the electric and magnetic fields respectively, and J are the charge and

current densities, respectively, and is the differential operator

 

= , , . (3.37)
x y z

To write these equations in the current notation one defines the Electomagnetic field

strength tensor F to be


0 E1 E2 E3
E1 0 B3 B2
F = . (3.38)
E2 B3 0 B1
E3 B2 B1 0
Equations (3.35) and (3.36) are combined

F = J (3.39)

with J = (, J1 , J2 , J3 ). Equations (3.33) and (3.34) are united as

F + F + F = 0 , (3.40)

and

19
F = F . (3.41)

These relations are true in special relativity where one is working in Minkowskian

space. To extend these relations to GR, simply replace with g , and with .[1, 15]

3.3.2. The Riemann Curvature Tensor

It has been shown how the affine connection helps describe the generalizations of

straight lines and derivatives in arbitrary coordinate systems. As stated before, when

= 0, one knows the local spacetime is flat. However, 6= 0 does not always imply that

there is a gravitational field present. It can be shown that when one is working in

non-Cartesian coordinates (for example, spherical polar coordinates) the connection terms

are not zero. To see if spacetime is actually curved, one needs a quantity that can describe

the intrinsic properties of spacetime.

This quantity is known as the Riemann curvature tensor. It is defined in terms of

the affine connection

R + . (3.42)

The overall sign of this tensor is a matter of convention, so it varies with from author to

author. This discussion follows the conventions of [1].

A geometric interpretation of the Riemann tensor can be explained with a simple 2-D

example. Imagine a triangle in a flat 2-D space. Define a vector V to be located somewhere

on the triangle. If one shifts the location of the vector without rotating it, it will continue to

point in the same direction. Now imagine a triangle on the surface of a sphere defined with

20
Figure 3.1. A vector parallel transported along a closed loop on a sphere.

points A and C located on its equator and a point B located on its north pole. Starting

vector V at C and pointing at B and shifting it from C to A without rotating, one finds it

is still pointing at B. Next V is shifted from A to B then from B back to C. When this

circuit is completed, one finds that V is no longer pointing at B, despite the fact that it has

not been not manually rotated (See Figure 3.1). The shifting of a vector around a closed

circuit of spacetime is called parallel transport. The rotation of V while parallel

transporting it along CAB comes from the fact CAB lies in a curved surface. On a flat

surface, there is no rotation. The Riemann tensor measures the change in the direction a

vector is pointing when it is parallel transported along a closed loop. If spacetime is flat, the

Riemann tensor will always be zero, despite the coordinates one uses.

The Riemann tensor has two contractions that are used in GR, the Ricci tensor R ,

R = R = g R = g g R , (3.43)

and the Ricci scalar R,

R = R = g R . (3.44)

21
The symmetries in the Riemann tensor makes the Ricci tensor and scalar the only unique

contractions of R ,[1, 15] and it will be shown in the next section how these contractions

are useful in describing the Einstein Field Equations.

3.4. Einstein Field Equations

The equations of motion for the metric are the Einstein field equations. There are

multiple ways to go about deriving them, but for this discussion, we will use the calculus of

variations method. The action I is composed of the action from gravitational fields IG and

the action from everything else (such as matter and electromagnetic fields) IM ,

Z
I = IG + IM = d4 xL . (3.45)

The action for the gravitational field in its simplest form is


Z
1
IG = gRd4 x , (3.46)
16G

where g =det|g |. Equation (3.46) is known as the Einstein-Hilbert action[16]. It will be

shown later that the Einstein-Hilbert action can actually contain a function of R rather

than just simply R, and one of the most common ways to modify GR is to replace R in the

action with f (R).

We take the action to be an extremum, so

I = IG + IM = 0 . (3.47)

The variation of the actions are defined as

22

Z
1
IM d4 x gT g , (3.48)
2

and


Z
1
IG ( gR)d4 x , (3.49)
16G

with G being Newtons constant and


( gR) = ( gR g ) = gR g + R g + gg R . (3.50)

Using the relations[1]

R = 0
1
g= gg g
2
g = g g g ,

it is found that

1
( gR) = g[R g R]g . (3.51)
2

Then one is left with

1
R g R + 8GT = 0 , (3.52)
2

which are the Einstein field equations. [1, 16] The energy-momentum tensor T is

dependent on what matter and fields are present, but its components can be thought of

qualitatively as [2]

23
T 00 = (Energy Density)

T 0i = T i0 = (Momentum Density or Energy Flux)

T ij = (Stress Tensor) .

The field equations are more commonly written in their covariant form

G = 8GT , (3.53)

where G is the Einstein tensor

1
G = R g R . (3.54)
2

The Einstein tensor and the energy-momentum tensor are divergenceless

G = T = 0 , (3.55)

where the zero divergence of the energy-momentum tensor comes from the conservation of

energy and momentum. Because the metric is symmetric, the field equations are ten

non-linear second order differential equations of the metric components.

It is instructive to examine the Newontian limit of the field equations. Using the

contraction of (3.53), we find

R = 8GT , (3.56)

where T = g T and G is Newtons constant (and not the contraction of the Einstein

24
Tensor G = g G ). This allows one to rewrite (3.53) as
 
1
R = 8G T g T . (3.57)
2

We establish the exact same conditions as earlier. The gravitational field will be a

small perturbation h about a Minkowski background

g = + h , |h |  1 (3.58)

and will be constant in time. Since h is sufficiently small, the first order terms of h in

R will dominate the contribution. R to the first order is written,

1
h h h + h

R = (3.59)
2

Since the system is non-relativistic, the energy density piece dominates, so we have

1 1
R00 = h00 = 2 h00 , (3.60)
2 2

where 2 is the standard, non-covariant Laplacian operator (x2 , y2 , z2 ). We also know from

(3.27),

h00 = 2 , (3.61)

so

R00 = 2 (3.62)

On the right hand side, we write T as a perfect fluid[1]

dx dx
T = ( + p) + pg , (3.63)
d d

25
with being the energy density and p being the momentum of spacetime. Again since this

is a non-relativistic system, the energy density of spacetime dominates

1
T00 g00 T = (3.64)
2 2

combining the left-hand and right-hand sides, one arrives at the Poisson equation for a

classical gravitational field,

2 = 4G . (3.65)

Once again, one finds GR reduces to the Newtonian limit of gravity.

26
CHAPTER 4: Gravitational Waves

There are many similarities between the Einstein equations and the Maxwell

equations of electromagnetism. From the Maxwell equations, one can derive a wave equation

that describes a fluctuation in the ambient electromagnetic field. One can do the same for

the Einstein equations. Since the Einstein equations describe the geometry of spacetime, a

wave equation will describe a fluctuation of spacetime itself. The Einstein equations in

principle are very difficult to solve because they are highly non-linear. The main reason for

this is that the charge of gravitation is energy and momentum, so a gravitational field will

interact with itself.

One way to make progress is to work with a field that is weak enough to not make

signficant interactions with itself. This is done when one looks at the weak field limit. From

equations (3.61) and (3.65) it is shown that the metric can be thought of classically as the

gravitational potential. In natural units this potential is dimensionless just like the metric.

The weak field limit is valid in familiar locales. For example, the gravitational potential is of

the order of 109 at the surface of Earth and 106 at the surface of the Sun.[1] GWs are

described as oscillations of the metric, and their amplitudes are expected to be of the order

1020 , considerably weaker than fields like Earths.[2, 3] For potentials of this magnitude, it

is evident that terms second order and higher in the metric will be negligible compared to

first order terms.

4.1. Basics of Gravitational Waves

This section reviews the work done in Chapter 10 of reference[1]. To arrive at a wave

equation from the field equations, one looks at a metric that is very close to the Minkowski

metric
g = + h , (4.1)

where |h |  1, and is so small that any term higher than first order in h will be

negligible. To write the contravariant form of the metric, g g = must hold, so to first

order

g = h . (4.2)

Starting with the field equations


 
1
R = 8G T g T , (4.3)
2

one assumes the wave is propagating in a vacuum far away from sources (T = 0). Then

the field equations become

R = 0 . (4.4)

To first order, R is written as

1
R = [ h + h h h ] = 0 . (4.5)
2

The solutions to these equations will vary with the coordinate system one uses. The most
0
general transformation from x to x that leaves the field weak is

0
x = x + (x) , (4.6)

where (x) is a coordinate shift based on the original coordinates. The Lorentz

transformation is then written

28
0
x
= + . (4.7)
x x

The metric in the new coordinate system is

0 0
0 0 x x
g = g , (4.8)
x x

so if h is a solution of (4.5), so will be

h0 0 = h . (4.9)

One can choose the form of (x), a property known as gauge invariance. To solve

(4.5), one must choose what gauge (what set of conditions) to work in. The most convenient

choice is the harmonic gauge

g = 0 , (4.10)

which to first order gives

1
h = h . (4.11)
2

The harmonic gauge turns (4.5) into a wave equation,

h = 0 , (4.12)

with a plane wave solution,

h (x) = e exp(ik x ) + e exp(ik x ) , (4.13)

29
where e is the wave amplitude and k the wave vector. From (4.11) and (4.13) it is easy to

see that

1
k e = k e (4.14)
2

and

k k = 0 . (4.15)

Since the metric is a symmetric 44 matrix, e must also be. This means there are 10

degrees of freedom; however, it will be shown that not all are physical.

Rotating the coodinate system to where the wave travels in the +x3 -direction (4.15)

gives the components of the wave vector to be

k1 = k2 = 0 k3 = k0 = . (4.16)

where is the waves radial frequency. Combining (4.16) with (4.14), one finds that e10 , e20 ,

e30 , and e22 can be written in terms of the other six, reducing the degrees of freedom from

10 to 6:

1
e10 = e31 , e20 = e32 , e30 = (e33 + e00 ), e22 = e11 . (4.17)
2

One can shift coordinates, so h goes to h0 0 given in (4.9). One then chooses

(x) = i exp(ik x ) i exp(ik x ) , (4.18)

which gives

30
e0 0 = e + k + k , (4.19)

and the remaining six components change to

e10 10 = e11 e10 20 = e12

e10 30 = e13 + 1 e20 30 = e23 + 2

e30 30 = e33 + 23 e00 00 = e00 20 .

Because one has the freedom to choose the components of , it is evident the only

components that have physical significance are e11 and e12 because the others can be made

zero by choosing appropriate components of . Thus in this gauge (known as the

transverse-traceless gauge), the wave amplitude e is the polarization tensor of GWs in

GR


0 0 0 0
0 e11 e12 0
e =
0 e12 e11
. (4.20)
0
0 0 0 0

With these components, one can see that a gravitational wave is quadrupolar in

nature. A wave passing perpendicular to a plane that contains a circular ring of test masses

will stretch and compress the ring as seen in Figure 4.1.

4.2. Energy and Momentum of Gravitational Waves

Gravitational waves carry energy and momentum. The previous section only used the

field equations to first order, but just as the energy and momentum of electromagnetic

31
Figure 4.1. A passing gravitational wave displaces a circular ring of test masses. The top ring
shows the perturbation caused by h11 . The bottom ring is perturbed by h12

waves are second order in their wave amplitude, one will find that the energy and

momentum of gravitational waves are second order in h . This section also follows the work

of reference[1]. Later in this document, the methods used to calculate the energy and

momentum of GWs for GR will be expanded to calculate the energy and momentum of a

non-GR gravitational wave.

Starting with the field equations

1
R g R = 8GT , (4.21)
2

one rewrites them as

(1) 1
R + R(1) = 8G[T + t ] (4.22)
2

where

1 (1) 1 1
t [R R + g R R(1) ] , (4.23)
8G 2 2
(1)
and R is the first order Ricci tensor

32
(1) 1
R = [ h h h + h ] . (4.24)
2

t can be interpreted as the energy-momentum tensor of the waves. It should be

noted that t is not generally covariant (hence the quotations about tensor), but it is

Lorentz covariant (meaning equations with t hold under Lorentz transformations that are

used in special relativity, but not under general coordinate transformations). If we break the

exact equations up in terms of a power series in h , the first terms that arise in t are the

second order terms of the Einstein tensor

1 1 1
t = G(2)
= h R
(1)
+ h R
(1) (2)
+ R + R(2) , (4.25)
2 2 2
(2)
where R is the second order Ricci tensor

(2) 1
R = h [ h h h + h ]
2
1
+ [2 h h ][ h + h h ] (4.26)
4
1
[ h + h h ][ h + h h ] .
4

(1) (2)
It is known from (4.5) that R = 0, so t is only dependent on R . When discussing

the energy and momentum carried by a wave, one is not concerned about a specific point in

the wave, but rather the average energy and momentum carried over several wavelengths.

(2)
Using the harmonic condition from (4.11) and the wave condition from (4.15), R can be

re-written

(2) 1 1
R = k k [e e |e |2 ] , (4.27)
2 2

33
which in the transverse-traceless gauge is

(2)
R = k k [|e11 |2 + |e12 |2 ] . (4.28)

Since k k = 0, R
(2)
is the only contributing term for t from (4.25). One now has a result

for the average momentum t03 carried by a GW

2
t03 = [|e11 |2 + |e12 |2 ] , (4.29)
8G

(4.29) will be useful later on when calculating the energy absorption cross section of

gravitational wave detectors.

4.3. Sources of Gravitational Waves

Just as electromagnetic radiation is emitted by accelerated charged particles,

gravitational radiation is emitted by accelerated mass-energy. Electromagnetic waves cannot

be emitted from spherically symmetric electric or magnetic sources; they need a changing

dipole moment in the electric/magnetic field to be emitted. It can be shown the first

moment in a gravitational field that can emit GWs is quadrupolar.[17] The zeroth moment

is the mass-energy distribution M0


Z
M0 = d3 x , (4.30)

and the field it produces is the Newtonian Potential

GM0
h (4.31)
r

where r is the distance from the source. The conservation of mass-energy forbids an isolated

source to vary dynamically; however in scalar-tensor modified gravity theories such as f (R)

a new scalar field is introduced that can radiate waves.

34
Next is the dipole moment of mass-energy M1

Z
M1 = xi d3 x . (4.32)

To get a field out of M1 , one must take a time derivative[17]

Gd
h M1 (4.33)
r dt

The time derivative of M1 is the total momentum of the source which must be conserved.

One can always boost to a frame where dM1 /dt = 0, so the dipole moment cannot

contribute to GW emissions. The same issue arises when one looks at the first moment of

the mass current ji = vi which is the spin angular momentum. Angular momentum of an

isolated source must be conserved.

The first moment that is not restricted to any conservation law is the quadrupole

moment M2
Z
M2 = xi xj d3 x . (4.34)

The weakness of gravity coupled with the fact that GWs need at least a quadrupole moment

in mass-energy to be emitted means that a source that can produce detectable GWs must

be incredibly massive. There are four categories of systems that emit detectable GWs, to

which we now turn.

The first is a compact binary star system that is inspiralling to an inevitable

coalescence. These binary systems are the canonical example of how GWs are emitted

because of their natural quadrupole moment and how simple it is to quantify expected

amplitudes and frequencies from these systems. These systems are typically neutron star -

neutron star systems, but can also be neutron star - black hole or black hole - black hole

35
systems. For two stars of equal mass M in a circular orbit of radius R with a period T , the

expected frequency of the emitted GW is simply 2/T , and the expected amplitude h is[3]

2GM 2 /R
h , (4.35)
r

where r is the distance from the source to the detector. Observing inspiralling binary

neutron star systems has lead to indirect detection of gravitational waves, which is to be

discussed later.

Another possible source would be a single rapidly spinning pulsar with an equatorial

ellipticity e.[1] With a moment of inertia I and frequency f , a pulsar can radiate GWs with

a characteristic amplitude[17]

GIf 2 e
h . (4.36)
r

The gravitational collapse of massive stars have also been regarded as a good

candidate for GW emission.[17] If the core of the star is rapidly spinning during its collapse,

the resulting supernova could have a strong enough quadrupole moment to produce GWs.

Spherically symmetric supernova are also a prime candidate to test scalar-tensor modified

gravity theories. If there exists a scalar mode predicted by modified theories of GR,

observing the GWs produced by a supernova would be a way to find out.

It is also expected that the universe has a background GW field that results from a

large number of random sources including even remnants of the early universe. The cosmic

microwave background allows us to study the universe back to about 380,000 years from the

beginning of the universe, and studies of Big Bang Nuceosynthesis allows us to better

understand conditions when the universe was only 3 minutes old. The weakness of gravity

36
allows gravitational waves to pass through matter almost entirely unaffected, and although

this makes GWs very challenging to detect, studying a primordial GW field would allow us

to look back earlier than 1024 s after the beginning of the universe[3], which would allow us

to study the laws of nature at much higher energies than humanity could ever hope to

produce in a laboratory.[18, 19, 20, 21, 22, 23, 24]

4.4. Detection of Gravitational Waves

4.4.1. Hulse-Taylor Observation

Hitherto, no experiment has yielded what most astrophysicists would consider direct

detection of gravitational waves, but there has been indirect detection through observing

binary pulsars.

The first binary pulsar system was observed by Joseph Taylor and his student Russell

Hulse in 1974. When first discovered, the single pulsar PSR 1913+16 was found to have a

rotational period of 59 ms. During Hulse and Taylors intial attempts to time the pulsar

down to the microsecond, they found the period would change up to 80 s throughout a

day, and as much as 8 s in 5 minutes. Pulsars are known to have very consistent rotational

periods. An irregular pulsar might vary 10s in its period over a year, so the change in

pulses were most extreme and peculiar. Through further recordings, they found the

variations in pulses followed an oscillatory pattern with a period of 7.75 hours. This

repetitive change in pulses led them to suspect they had found a binary system, and

observed Doppler shifts that can be explained through orbital motion have provided

sufficient corroborating evidence[25].

Since its discovery, the orbital period of PSR 1913+16 has been measured as a test of

GR. Emitted gravitational waves take energy out of their source. The energy loss causes

37
Figure 4.2. Hulse-Taylor Results[27]: graph of the change in time to reach periastron (the point
where orbiting bodies are closest to each other) of PSR 1913+16 vs time.

stars in a binary to inspiral toward each other leading to an eventual cohesion. As the stars

move toward each other, the radial velocities will increase, which means the orbital period

decreases. Hulse and Taylor recorded the orbital period of PSR 1913+16 over the next

several years after its discovery, and have found the period has been decreasing at a rate of

0.997 0.002 times the predicted rate of GR[26]. Figure 4.2 provides a visual of their

outstanding observations. Although this method of detection is not considered direct, the

observation provided results convincing enough to award Hulse and Taylor with the 1993

Nobel prize.

4.4.2. Detectors

The leading experiments on GW detection are the LIGO and VIRGO projects which

are ground based interferometers. An interferometer with two perpendicular arms measures

38
the phase change in the beam. The phase change occurs when an incident GW changes an

arm length L

L hL , (4.37)

where L is the change in length and h is the wave amplitude. Light takes about 105 s to

make a round trip in a 4 km long arm like the ones used in LIGO, which is much shorter

than the period of GWs the detector is trying to find. Optical cavities are built into the

interferometers to keep light in the arm for the entire period of an incident wave.[3].

The signal strength of an incident wave is dependent on the angle between the wave

vector and the plane of the detecors arms. From Figure 4.3 it is easy to see that a wave

vector perpendicular to the detector plane produces a maximum signal; on the other hand, a

wave traveling parallel with an arm will go unnoticed by that arm. For this reason, LIGO

has built detectors in Livingston, Louisana and Hanford, Washington. A wave incident with

Earth will hit the detectors at different angles which will produce different signals. This also

allows one to be able to determine the direction of the wave vector. Interferometers also

have the advantage of being able to detect GWs through a wide range of frequencies.

The first attempts to directly detect gravitational waves were by using resonant mass

detectors. GWs couple very weakly to matter, so a wave travelling through matter will not

be significantly affected. This is what makes detection so difficult. The main idea behind a

resonant mass detector is to build the detector with a resonance frequency that matches the

incident wave, so some of the waves energy can be transferred to the detector which then

causes the detector to oscillate[1, 3].

In the early days the detectors where built as cylinders, but there has been recent

39
Figure 4.3. Incident GW perpendicular to the plane of the interferomter. The test masses are
low-transmissivity mirrors that provide optical cavities to increase the travel distance of the beam.

effort to make spherical resonant mass detectors[5]. The main advantage of a spherical

detector is that the signal strength is independent of the angle of the wave vector. Resonant

mass detectors have a very narrow range of detectable frequencies, so it is important to have

a good target frequency when building a detector. Once there are known frequencies of

detectable GWs, resonant mass detectors will provide a more affordable way to build a GW

observatory.

GW detectors are different from conventional telescopes in that they are

omni-directional. They function more like microphones than telescopes in that they monitor

their environment for disturbances in all directions.[3]

4.4.3. GW Strain

GW literature often mentions the strain of the wave.[2, 3, 4] This can refer to two

quantities; the first is simply the wave amplitude h, a dimensionless quantity. Detectors can

only detect waves that lie within their frequency bandwidth, so it is convenient to study a

strain that is dependent on frequency. Signals received by detectors running in the absence

40
of GWs are noise n(t), and a statistical analysis of the noise produces a detectors noise

correlation function ( )

( ) = hn(t1 )n(t2 )i (4.38)

where is the average time between noise incidents n(t1 ) and n(t2 ). The performance of a

detector is characterized by the power spectral density of its noise background[3]


Z
1
S(f ) ( )ei2f d, f 0 , (4.39)
2

where S(f ) is the power spectral density. The unit of S(f ) is Hz1 , and by combining

equations (4.38) and (4.39), S(f ) can be related to the detector noise[3]

1
he n (f 0 )i = S(f )(f f 0 ) ,
n(f )e (4.40)
2

with n
e(f ) being the Fourier transform of the noise n(t). Thus, the detectors noise amplitude
p
is the square root of its power spectral density S(f ). It is used to measure the detectors

strain sensitivity, and its unit is Hz1/2 . The second quantity that refers to a GWs strain is

used to compare the wave amplitude to the noise amplitude. This strain h is written as

p
h=e
h f (4.41)

where e
h is the wave amplitude of the Fourier transform of the wave h(x, t). e
h has units of

Hz1 , which gives h units of Hz1/2 . Throughout the rest of this document the wave strain

will refer to h.

Figure 4.4 is a graph of the lowest detectable strains of contemporary and future

detectors. In the top right corner of Figure 4.4 lies the range of the Mario Schenberg

spherical resonant mass detector. As one can see, its range is very limited compared to the

interferometers LIGO and VIRGO.

41
Figure 4.4. Detectable GW amplitude strain vs GW frequency[28]

4.5. Tidal Force on a Detector

In general, tidal forces are felt on an object by experiencing a difference in

gravitational fields between points in its volume. The tides of the ocean are caused by this

phenomenon. The gravitational fields of the Sun and the Moon create a small difference in

the gravitational field felt along Earths surface, and this causes Earths less viscous parts of

its surface (its oceans) to be made prolate in the directions of the Moon and Sun. Tidal

forces also affect satellites in orbit around Earth, causing them to experience a slight stress

resulting in a slight elongation towards the direction of Earth, a phenomenon called

microgravity. Since gravitational waves are oscillations of the ambient gravitational field, a

GW passing through an object will cause the object to experience oscillating tidal stresses

along the waves polarizations.

42
The effect of a gravitational wave passing through a detector on Earth (like a resonant

mass detector) has been studied[29]. This section reviews these effects. To begin, one first

shifts to the proper reference frame of the detector

ds2 = (1 + 2ai xi )dt2 + ij dxi dxj , (4.42)

where ai is the gravitational acceleration felt near the surface of Earth (|ai | = 9.81 m/s2 ).

In introductory physics courses, the gravitational potential near the surface of Earth is

described as a x. Recall in the Newtonian limit of GR that the 00 component of the metric

(g = + h ) includes the Newtonian potential

g00 = 00 + h00 = 1 2 , h00 = 2 (4.43)

while the spatial components of the field (hij ) were much closer to Minkowskian compared

to g00

gij ij . (4.44)

Therefore one should expect the spacetime interval to take the form of (4.42) by using the

Newtonian limit metric tensor and replacing with a x in the reference frame of a detector

on the surface of Earth.

Now in this reference frame, consider a mass element in the detector. The mass

element will be pushed and pulled by adjacent matter and any other external

non-gravitational fields. These forces can be expressed by a 4-force per mass f , which gives

the geodesic equation

43
d2 x
dx dx

f = + . (4.45)
d 2 d d

The mass element will experience very small velocities in the detector. Thus like in the

Newtonian cases before, the zeroth component of the 4-velocity will dominate, and the

spatial pieces can be neglected. It is known in special relativity that the dot product of an

objects 4-acceleration and 4-velocity u is zero ( u = 0). Since the spatial terms of the

mass elements 4-velocity is negligible, the temporal component of the force per mass should

also be negligible. Solving for the conventional acceleration of the mass element, one finds

2 2
d 2 xj d2 xj
  
dt d
j
f = 2
+ j00 2
= fj j00 , (4.46)
dt d dt dt

where d /dt can be found from Equation (4.42)

2 2
dxi dxj
 
i dt d
1 = (1 + 2ai x ) ij = 1 + 2ai xi . (4.47)
d d d dt

j00 is known to be simply aj to the zeroth order[29], which is expected from a

Newtonian point of view (because the geodesic equation is the GR equivalent of F = ma, so

on the surface of Earth, compensates for the deviation of an objects path moving

through a gravitational field). However to see the tidal effect of a GW, one needs to know

j00 to at least first order in x . In the transverse traceless gauge, the only non-zero

components of the Riemann tensor are of the form Rj 0i0 [1, 29], and general covariance

enforces that this is true in any gauge/coordinate system. So it is written

Rj 0i0 = 0 j0i i j00 + 0i j0 00 ji , (4.48)

where to zeroth order in x one has

44
j00 = aj

0i0 = ai

000 = 0ik = ji0 = jik = 0 ,

which gives

Rj 0i0 = i j00 + 00i j00 i j00 = ai aj Rj 0i0 . (4.49)

Integrating (4.49) throughout space gives the desired result

j00 = aj (1 + ai xi ) Rj 0i0 xi . (4.50)

Gathering terms one finds the conventional acceration of the detector is[29]

d 2 xj
= f j (1 + 2ai xi ) aj (1 + ai xi ) + Rj 0i0 xi , (4.51)
dt2

where the only contribution from GWs is in Rj 0i0

d 2 xj
   
= Rj 0i0 xi (4.52)
dt2 GW s GW s

This result will be used later in the context of GWs impinging upon spherical resonant

detectors.

45
CHAPTER 5: Modified Gravity

The equivalence principle forms a foundation for an entire class of gravity theories

known as metric theories. Metric theories of gravity are based on three assumptions: 1)

There exists a symmetric metric. 2) Test bodies follow geodesics of the metric. 3) In local

Lorentz frames, the non-gravitational laws of physics are those of special relativity[30].

These citeria are based on the equivalence principle discussed in section 3.2, and the work

done in chapter 6 is based on metric gravity theories. General Relativity and f (R) gravity

are examples of metric theories of gravity. This chapter reviews the properties of f (R)

gravitational waves, and expands these properties to calculate the energy-momentum of an

f (R) GW. The results of this chapter will be used in the next where the absorption

properties of a GW detector are dependent on metric theories of gravity.

As gravitational waves have not been directly detected, they provide a natural

environment within which to examine possible modifications of general relativity. For

example it is known that in modified gravity theories, GWs have their standard tensor

polarization as in GR (from equation (4.20)) plus an addition scalar polarization. Where

such a scalar mode arises, and how it it alters the energy and momentum of GWs will be

shown in this chapter.

5.1. GWs in f (R) Gravity

This section follows the work done in reference[14]. The field equations for f (R)

gravity are derived from the generalization of the Einstein-Hilbert action in equation (3.46)


Z
IG = d4 x gf (R) . (5.1)

Following the same calculus of variations procedure in Section 3.4, one finds the field
equations to be

1
F (R)R g f (R) + (g )F (R) = 8GT , (5.2)
2

where F (R) = df (R)/dR and is the covariant derivative.

The trace of the field equation gives

F (R)R 2f (R) + 3 F (R) = 8GT . (5.3)

Perturbing (5.3) in a vacuum with a non-zero background curvature R0 yields

3 F + R0 F + F (R0 )R 2f = 0 . (5.4)

Using the relations f = F (R0 )R and F = R[dF (R0 )/dR], one obtains a wave equation

for a massive scalar mode

hf = m2s hf , (5.5)

where hf is the field of the scalar mode[14]

F
hf , (5.6)
F (R0 )

and the mass ms is written as[14]

 
1 F (R0 )
m2s = R0 , (5.7)
3 dF (R0 )/dR

If the background curvature of the vacuum is taken to be Minkowskian (R0 = 0), the

47
solution of the wave equation (5.5) is written

hf = A exp(iq x ) + A exp(iq x ) (5.8)

where q is the scalar wave vector with components (s , qi ) and A is the wave amplitude.

From equation (5.5), one finds

q q = s2 + q 2 = m2s , s2 = q 2 + m2s . (5.9)

Thus, the metric perturbation h from a gravitational wave is decomposed into scalar and

tensor pieces[14]

h = hT + hS (5.10)

where hT is the perturbation one finds in non-modified GR and


hf 0 0 0
0 hf 0 0
hS = hf = . (5.11)
0 0 hf 0
0 0 0 hf

5.2. Energy-Momentum of GWs in f (R) gravity

Now we turn to our calculation of the energy-momentum tensor for an f (R)

gravitational wave. This section expands on the methods used in Section 4.2 by using f (R)

gravity to generalize the energy and momentum carried by a GW, similarly to how the f (R)

action generalizes the field equations. This result will be used later to calculate the energy

flux for an f (R) GW incident with a spherical resonant mass detector.

To begin, the functions f (R) and F (R) are decomposed in powers of h :

48
f (R) = f (0) + f (1) + f (2) + ... , (5.12)

F (R) = F (0) + F (1) + F (2) + ... , (5.13)

In the GR case, the energy momentum tensor for a GW t is found to be the second order

terms of the Einstein tensor G . The energy-momentum tensor for an f (R) GW is similarly

found to be the second order in terms of the f (R) field equations

1 h (0) (2) 1
t = F R + F (1) R
(1)
( f (2) + h f (1) )
8G 2
(1)
+ ( F (2) F (1) ) + h ( F (1) ) (5.14)
i
F (2) + (1)
F
(1)
.

From equation (5.6), F (1) can be rewritten as

F (1) = F (0) hf . (5.15)

For a gravitational wave incident on a detector in the x3 -direction, the energy flux that

passes through the detector is determined from the average t03 component of t . The

average of t03 is written from (5.14):

1 h (0) (2) (1) (1)


i
t03 = F hR03 i + F (0) hhf R03 i h0 3 F (2) i + F (0) h03 hf i , (5.16)
8G

where the brackets hi denote the average over the wave. To calculate the values of these

terms, one finds the following identities from equations (5.8) and (5.11) useful:

49
1
hS = hS = hf , (5.17)
4
1
q A = q A = q A , (5.18)
4

where A is the polarization tensor of the scalar wave


+A 0 0 0
0 A 0 0
A = (5.19)
0 0 A 0
0 0 0 A
(2)
The first term in (5.16) contains the average of the second order Ricci tensor hR i. Just

like the metric tensor can be split into scalar and tensor pieces in equation (5.10), the Ricci

tensor can be decomposed into its tensor and scalar pieces. The average of the tensor piece

(2)
has been calculated in (4.28). The average of the scalar pice of R is found to be

n
(2)
hR i = < A [q q A q q A q q A + q q A ]
1
+ [q A q A ] [q A + q A q A ] (5.20)
2
1 o
[q A + q A q A ] [q A + q A q A ] .
2
Using (5.18), equation (5.20) can be reduced to

(2)
hR i = q q |A|2 m2s |A|2 , (5.21)

and
(2)
hR03 i = q0 q3 |A|2 . (5.22)

The second term in (5.16) is found to be

(1)
hhf R03 i = 2q0 q3 |A|2 , (5.23)

50
which cancels the final term

(1)
h03 hf i = 2q0 q3 |A|2 . (5.24)

The only term left from (5.16) is h0 3 F (2) i. To calculate this term, one first needs to

calculate F (2) . One begins by performing a Taylor expansion of f (R),

c2 2 c3 3
f (R) = c0 + c1 R + R R + ... , (5.25)
2 3

where cn s are expansion coefficients. Since F (R) = df (R)/dR, it follows that

F (R) = c1 + c2 R + c3 R2 + ... , (5.26)

and

F (2) = c2 R(2) + c3 R(1) R(1) , (5.27)

where

R(2) = (g R )(2) = R
(2)
h R
(1)
, R(1) = R
(1)
. (5.28)

Through a long an rigorous calculation, it is found that F (2) has terms of h2f and hf hf

with coefficients in multiples of c2 , c3 , and ms . It is found that

h0 3 h2f i = 0 (5.29)

and

h0 3 [ hf hf ]i = 0 , (5.30)

51
so h0 3 F (2) i does not contribute to t03 . The final result is then

F (0)
t03 = [k0 k3 (|e11 |2 + |e12 |2 ) q0 q3 |A|2 ] , (5.31)
8G

or in terms of the wave frequencies

" s #
F (0) m 2
t03 = 2 (|e11 |2 + |e12 |2 ) + S2 |A|2 1 2S . (5.32)
8G S

The significance of this result is that it is independent of the chosen model of f (R).

One can see that (5.32) reduces back to the GR result with f (R) = R and S = A = 0. To

summarize, we have calculated the energy-momentum for a GW in f (R), and we will now

see what effect Equation (5.32) will produce in next chapter.

52
CHAPTER 6: Acoustic Response of a Spherical Detector
6.1. Mathematical Framework

We would like to determine the observational effect produced by GWs in f (R)

gravity. For this, we now turn to a disscussion of the acoustic response of an isotropic elastic

solid caused by an incident gravitational wave. Since gravitational wave induced

perturbations will be very small compared to the size of the detector, the perturbations of

the solid will be modelled according to classical non-relativistic linear Elastic Theory.[31]

This section provides in depth details of the work done in reference[5].

Consider a mass element at coordinate x from an isotropic elastic solids center of

mass. An incident GW will displace the mass element described by a displacement vector u

(Figure 6.1). This displacement vector is a solution to:

2u
2
2 u ( + )( u) = f(x, t) , (6.1)
t

where is the mass density of the solid, and are the solids Lame coefficients, and is

the conventional differential operator. The function f(x, t) is the driving force of the

perturbations and will be separated into the product spatial and temporal functions

f(x, t) = f(x)g(t) . (6.2)

In the event of detection the body is inintially at rest, and then the driving force (a

gravitational wave) begins to perturb the solid. Thus, the following initial conditions are

adopted:

u(x, 0)
u(x, 0) = 0 , =0 . (6.3)
t
Figure 6.1. Displacement u of a mass element on the surface of a resonating sphere. The dark
circle represents the cross section of the sphere in equilibrium, while the gray represents the sphere
in a perturbed state.

The homogeneous equation

t2 u Lu = 0 , (6.4)

has eigensolutions

2
LuN = N uN , (6.5)

where L is a differential operator defined as

L 2 + ( + )() . (6.6)

u(x, t) is separated into functions of x and t

u(x, t) = u(x)s(t) , (6.7)

The eigensolutions uN (x) have an orthogonality condition

54
Z
uM (x) uN (x)(x)d3 x = M M N , (6.8)

and the solution to the spatial dependent part of u is found using this condition

X
u(x) = fN uN (x) , (6.9)
N

where

Z
1
fN uN (x) f (x)d3 x , (6.10)
M

and M is the mass of the solid.

Using equations (6.5), (6.7), and (6.9), equation (6.1) takes the form

X
fN uN (x) t2 s(t) + N
2
 
s(t) = f (x)g(t) (6.11)
N

The driving force is structured in such a way that the temporal piece of the displacement

vector s(t) can be expressed in terms of a Green function integral

Z
s(t) = S(t, t0 )g(t0 )dt0 , (6.12)
0

where S(t, t0 ) is the Green function. The Green function is the resonse of the system to a

unit impulse at t = t0 .[32] g(t) is also rewritten

Z
g(t) = g(t0 )(t t0 )dt0 . (6.13)
0

Putting (6.12) and (6.13) in (6.11) one arrives at

55
X Z Z
t2 S(t, t0 ) 2
S(t, t0 ) 0 0
g(t0 )(t t0 )dt0

fN uN (x) + N g(t )dt = f (x) , (6.14)
N 0 0

where the solution to S(t, t0 ) lies in the equation

t2 S(t, t0 ) + N
2
S(t, t0 ) = (t t0 ) . (6.15)

Equation (6.15) is solved with a Laplace transform L(f (t))

Z
L(f (t)) = f (t)ept dt , (6.16)
0

where p is the transform parameter. Transforming equation (6.15) gives

Z Z Z
0 pt 0 pt
t2 S(t, t )e dt + 2
N S(t, t )e dt = (t t0 )ept dt , (6.17)
0 0 0

which is simplified to

(p2 + N
2
)L(S(t, t0 )) = L((t t0 )) . (6.18)

S(t, t0 ) and (t t0 ) transform

0
L(S(t, t0 )) = ept L(S(p)) , (6.19)

and

0
L((t t0 )) = ept , (6.20)

which allows (6.18) to take the form

56
0 0 1
(p2 + N
2
)ept L(S(p)) = ept L(S(p)) = 2
. (6.21)
p2 + N

The inverse transform of /(p2 + 2 ) is sin(t). Using this in (6.21) gives the solution for

S(t, t0 )

0, 0 < t < t0
(
0 1
S(t, t ) = (6.22)
sin[N (t t0 )], 0 < t0 < t
N
and s(t)

Z
1 1
s(t) = gN (t) = g(t0 )sin[N (t t0 )]dt0 . (6.23)
N N 0

Therefore the general solution to (6.1) is

X fN
u(x, t) = uN (x)gN (t) (6.24)
N
N

6.2. GW Tidal Forces

Gravitational waves interact with matter through tidal forces. From the conventional

acceleration in Equation (4.52), one can write a GW tidal force density as

fi (x, t) = R0i0j (t)xj . (6.25)

The Riemann tensor is evaluated at the solids center, so there is no dependence of x. The

Riemann tensor is symmetric in i and j, and any arbitrary symmetric tensor Sij can be

decomposed as[5]

2
X
(S) (S) (m)
Sij (t) = S (t)Eij + S (m) (t)Eij , (6.26)
m=2

57
Figure 6.2. Quadrupolar modes of a sphere[33]. The dark regions signify the maximal radial
perturbation of the sphere from a GW incident with the zenith direction of the sphere.

(m) (S)
where Eij are 5 linearly independent symmetric and traceless tensors, and Eij is a

multiple of the identity matrix[5],


 1/2 1 0 0
(S) 1 0 1
Eij = 0 (6.27)
4
0 0 1

 1/2 1 0 0
(m=0) 5
Eij = 0 1 0 (6.28)
16
0 0 2

 1/2 0 0 1
(m=1) 15
Eij = 0 0 i (6.29)
32
1 i 0

 1/2 1 i 0
(m=2) 15 i
Eij = 1 0 . (6.30)
32
0 0 0
(S) (m)
The matrices Eij and Eij are chosen to satisfy the relations:

(S) (m)
Eij ni nj = Y00 (, ) ; Eij ni nj = Y2m (, ) , (6.31)

where ni is the radial unit vector ni = xi /|x|, and the Y`m are spherical harmonics that are

58
chosen to represent the scalar (l = 0) and quadrupolar (l = 2) resonance modes of a sphere.

Figure 6.2 gives a visual of the quadrupolar resonances.

Additionally these matrices obey orthogonality relations given by

X (m0 ) (m) 15 X (S) (m)


X (S) (S) 3
Eij Eij = m0 m , Eij Eij =0 , Eij Eij = . (6.32)
i,j
8 i,j i,j
4

These relations can be used to invert (6.26):

4 (S)
S (S) (t) = E Sij (t) , (6.33)
3 ij

8 (m)
S (m) (t) = E Sij (t) . (6.34)
15 ij

Replacing Sij with R0i0j , the tidal force density can be decomposed into

2
X
f (x, t) = f S (x)g (S) (t) + f (m) (x)g (m) (t) . (6.35)
m=2

The scalar S and quadrupolar m components can be written in terms of matrices Eij as

(S) (S) 4 (S)


fi (x) = Eij xj , g (S) (t) = E R0i0j (t) (6.36)
3 ij
(m) (m) 8 (m)
fi (x) = Eij xj , g (m) (t) = E R0i0j (t) (6.37)
15 ij

(6.24) yields the formal solution of a solids response to a GW perturbation using (6.36) and

(6.37)

X uN (x) h 2
X i
(S) (S) (m) (m)
u(x, t) = fN gN (t) + fN gN (t) (6.38)
N
N m=2

59
So far, there has been no reference to the field equations of the gravitational model one

is using. However, the fact that the Riemann tensor is used means that the model must be a

metric theory of gravity. This means that Equation (6.38) is independent of the underlying

gravity theory as long it is a metric theory[5].

6.3. Spherical Detector Response

The work from the previous two sections can be applied to a spherical detector. The

integrals in (6.10) need to be solved in the context of a homogeneous sphere. This gives[5]:

Z
(S) 1
fnlm unlm (x) f(S) (x)d3 x = an l0 m0 , (6.39)
M Sphere

Z
(m0 ) 1 0
fnlm unlm (x) f((m ) (x)d3 x = bn l2 m0 m , (6.40)
M Sphere

where an and bn are constants measured in length that are dependent on the spheres

mechanical properties. The n index indicates the nth eigensolution, l = 0 and l = 2 signifies

the monopole and quadrupole modes respectively, and m signifies the particular mode.

Substituting Equations (6.39) and (6.40) into (6.38), one arrives at:


" 2
#
X an (S)
X bn X (m)
u(x, t) = un00 (x)gn0 (t) + un2m (x)gn2 (t) (6.41)

n=1 n0

n=1 n2 m=2

Equation (6.41) can be Fourier transformed to determine the displacement according to the

frequency of the wave. That is,

Z
U(x, ) u(x, t)eit dt (6.42)

which gives

60

X an
U(x, ) = un00 (x)G(S) ()[( n0 ) ( + n0 )]
i n=1 n0

X bn
+ un2m (x)G(m) ()[( n2 ) ( + n2 )] , (6.43)
i n=1 n2

where G(S) () and G(m) () are the Fourier transforms of g (S) (t) and g (m) (t).

Six transducers appropriately placed on the sphere would allow scientists to examine

the spheres oscillations of its monopole and quadruploe modes.[5, 33] If the location of the

GW source is known, a rotation of coordinates can be made to align the laboratory

+x3 -direction with the wave vector. Once the rotation is made, one determine the functions

G(S,m) () by studying the signals from the transducers, and then inverse Fourier transform

them for g (S,m) (t) to solve for the components of the Riemann tensor R0i0j (t). This is the

ideal case for testing GR because one would be able to directly determine the polarizations

of an incident GW without making any assumptions about the underlying gravity theory.[5]

If the GW source is not known, one cannot make the appropriate rotation of

coordinates. This means that one must make some assumption about the GW theory. An

assumption one can make is that the m = 1 quadrupolar modes cannot be excited by a

GW, which is the case for gravity theories like GR and f (R). With this assumption, one can

rotate coordinates so g (1) (t) = 0, and then solve for the components of the Riemann

tensor.[5]

6.4. Absorption Cross Section

We extend the work in the previous section and calculate the absorption cross section

of an incident gravitational wave passing through a spherical detector. Unlike the work done

in the previous two sections, the absorption cross section is dependent on the underlying

61
GW theory. Later subsections go over the calculation for GR and f (R) gravity.

Following the work in references[1, 5] , the absorption cross section of a detector is

given by the energy absorbed by the detector E() per unit incident GW flux ()

E()
abs () = . (6.44)
()

The energy absorbed is calculated by the integration of the spectral energy density W ()

over a small interval around = nl

Z + Z +
d d
E(nl ) = W () + W () , (6.45)
 2  2

where the spectral energy density W () is defined to be

Z
1 1 2
W () |U(x, )|2 d3 x . (6.46)
T Solid 2

The energy absorbed is then

1
E(n0 ) = M a2n |G(S) (n0 )|2 , (6.47)
2

and

2
1 2
X
E(n2 ) = M bn |G(m) (n2 )|2 , (6.48)
2 m=2

where M is the detector mass. The GW flux is calculated from the energy-momentum

tensor t of the wave that travels in the +x3 -direction by the formula

Z Z
d
() = (t)dt (6.49)
0 2

62
where

(t) = xi t0i (6.50)

In order to identify (), it is useful to further use the Fourier transform:

Z
(t) = deit () , (6.51)

which leads to
Z Z
()d = 4 2 ()d , (6.52)
0

where one uses the identity

Z
0
dtei( )t = 2( 0 ) . (6.53)

Finally, if () is an even function of then the GW flux () is

() = 8 2 () . (6.54)

6.4.1. GR

The previously known absorption cross section for GR is now derived. For a wave

traveling in the +x3 -direction, using Equation (4.29) and (6.54) yields

2
() = e11 |2 + |e
[|e e12 |2 ] . (6.55)
G

where ee11 and ee12 are the amplitudes of the Fourier transform of the wave. The energy

absorbed is determined by first calculating the Riemann tensor which to first order is

1
R0i0j = [0 0 hij 0 j h0i 0 i h0j + i j h00 ] , (6.56)
2

63
and the transverse-traceless gauge gives

1 1
R0i0j = 0 0 hij = 2 hij . (6.57)
2 2

Using this result to calculate g (S) and g (m) from (6.36) and (6.37), one finds the only

non-zero result comes from m = 2, as one would expect for a GR gravitational wave. g (2)

is

r
2 2
g (2) (t) = [h11 + ih12 ] , (6.58)
15

and

r
2 2
g (2) (t) = [h11 ih12 ] , (6.59)
15

Fourier transforming g (2) yields

r
(2) 8 3 2 e
G = [h11 + ie
h12 ] , (6.60)
15 n2

and

r
8 3 2 e
G(2) = [h11 ie
h12 ] , (6.61)
15 n2

where n2 is the detectors quadrupolar resonance frequency. One finds the energy absorbed

to be

8 3
E (2)( (n2 ) = 4 2 2
M n2 e11 + ee212 )
bn (e (6.62)
15

64
Summing E (2) and E (2) together and dividing by equation (6.55) yields the absorption

cross section for GR which is a previously known result [5]

16 2 2 2
m=2 (n2 ) = GM n2 bn . (6.63)
15

6.4.2. f (R)

In this section we show our calculation for the response of an f (R) GW. This is a

novel calculation, but the methods used follow the work of calculating the response of a GW

in the Brans-Dicke scalar-tensor gravity theory[5, 34] and produce an independent result for

f (R). The components of the Riemman tensor in f (R) are derived to calculate the energy

absorbed, and the t03 result from Equation (5.32) must be used to calculate the flux.

In f (R) gravity, a scalar mode exists in addition to the quadrupolar modes. The

scalar mode will affect the detector through the monopole and m = 0 quadrupole modes.

The flux from the scalar mode is found from combining (5.32) and (6.54)

F (0) e2 2 p
= A 1 m2s / 2 . (6.64)
G

where A
e is the amplitude of the Fourier transform of the scalar wave hf . The f (R) metric

perturbation is written


hf 0 0 0
0 h+ hf h 0
h = , (6.65)
0 h h+ hf 0
0 0 0 hf
where

h+ = e11 exp(ik x ) + e11 exp(ik x ) , (6.66)

h = e12 exp(ik x ) + e12 exp(ik x ) , (6.67)

65
and

hf = A exp(iq x ) + A exp(iq x ) . (6.68)

The Riemann tensor in this metric is

1
R0i0j = [0 0 hij + i j h00 ] . (6.69)
2

which, according to Equation (6.25), means that an f (R) GW will also produce a

longitudinal force. The functions G1 () are still zero just like in GR, but the monopole

mode and the zero quadrupole mode are no longer zero. They are respectively

r
m2s e
 
(S) 16 3 2
G (n0 ) = n0 + hf (6.70)
9 2

r
(0) 16 3 2 2 e
G (n2 ) = [ms n2 ]hf (6.71)
45

with ms being the mass of the scalar mode, and G(2) are the same as in GR. The energy

absorbed by the scalar and m = 0 modes of the sphere are

2
8 3 m2s e2

(S) 2 2
E = M an n0 + A , (6.72)
9 2

and

8 3
E (0) = M b2n [m2s n2
2 2 e2
]A . (6.73)
45

Dividing Equations (6.72) and (6.73) by Equation (6.64) yields the absorption cross sections

for the monopole mode and the m = 0 quadrupolar mode

66
8 2 G 2 2
2 (n0 + ms /2)
2
S (n0 ) = M a n p (6.74)
9F (0) 2
1 m2s /n0

8 2 G 2 23
m=0 (n2 ) = (0)
M b2n n2
2
(1 m2s /n2 ) (6.75)
45F

This result is complementary to the absorption cross section for a Brans-Dicke GW.

Like f (R), Brans-Dicke theory is a modified version of GR that includes a scalar field[1].

For a Brans-Dicke GW, the scalar field is massless and travels at the speed of light[34], and

the scalar wave distributes 5/6 of its energy into the scalar mode of the sphere and 1/6 into

the m = 0 mode[5]. In the massless limit of Equations (6.74) and (6.75) (ms = 0), the scalar

wave is found to have the same energy distribution to the scalar and m = 0 modes as in

Brans-Dicke theory.

6.5. Observational Possibilities

This section discusses briefly, the minimal detectable amplitude strain of the scalar

wave |A|min . The detectors signal to noise ratio SN R is defined[34]

A En
SN R = = (6.76)
S Emin

with S being the noise amplitude, En the energy absorbed by the detector given by

equation (6.45), Emin the minimum detectable energy that is dependent on the detectors

thermal and electronic noises, and A the wave strain given by equation (4.41)

p
A= fA
e . (6.77)

The theoretical lower bound on Emin is fixed by quantum mechanics to be ~ in S.I.

67
units or in natural units. Using Equations (6.72), (6.73), and (6.76), one can determine

the minimal detectable scalar amplitude for an f (R) GW

  21
9 (SN R)Emin
|A
eS |min =
2 , (6.78)
8 2 M a2n [n0
2
+ m2s /2]

  12
45 (SN R)Emin
|A
em=0 |min =
2 2
. (6.79)
8 2 M b2n [m2s n2 ]

If Emin is taken to the quantum limit, the signal to noise ratio cannot be less than one.

Emin = and SN R = 1 produce the theoretically lowest wave strains a detector can see

according to this mathematical model; however, their are methods of quantum squeezing

that are used in quantum optics that may be viable for resonant mass detectors.[3, 35] A

detector with a degenerate quadrupolar eigenfrequency fn operating at quantum efficiency

can detect minimum strains given by


  12
p 45 2fn
Am=0 = fn . (6.80)
8 M b2n [m2s (2fn )2 ]2
2

6.5.1. Minimum Strain for the Mario Schenberg Detector

This section uses Equation (6.80) to calculate the quantum limit minimum

quadrupolar strain produced by the scalar wave that the Mario Schenberg detector could

detect. Table 6.1 shows the parameters used.

The parameter bn in Equation (6.80) has been calculated by using the relation[5]

Z R
1
bn = [An2 (r) + 3Bn2 (r)]r3 dr , (6.81)
M 0

where the functions An2 (r) and Bn2 (r) are the amplitudes of quadrupolar (l = 2) spheroidal

displacements in a solid elastic homogeneous sphere[5] [33]

68
Table 6.1. Mario Schenberg detector parameters [33]

S.I. units Natural Units


mass M 1147.85 kg 6.4486 1038 eV
radius R 0.3239 m -
degenerate eigenfrequency fn 3206.3 Hz 2.1097 1012 eV
Poisson Ratio 0.364 0.364
11
Young Modulus Y 1.33 10 Pa -
4.88 1010 Pa -
1.30 1011 Pa -
p1 -5.57 -5.57
p2 2.28 2.28

un2m (r, , ) = [An2 (r)r + Bn2 (r)R]Y2m (, ) . (6.82)

An2 (r) and Bn2 have values following[33]

R
An2 = p1 R j2 (qn2 r) + 6p2 j2 (kn2 r) , (6.83)
r r


Bn2 = p1 j2 (qn2 r) + 6p2 [rj2 (kn2 r)] , (6.84)
r

where p1 and p2 are normalization parameters determined from Equations (6.8) and (6.82),

and j2 (x) is the second order spherical bessel function

 
3 1 3 cos x
j2 (x) = 3
sin x . (6.85)
x x x2

The wave numbers qn2 and kn2 in (6.84) are the curl-free and divergenence-free wave

numbers of the resonances of the sphere, respectively. Their values are dependent on the

resonant eigenfrequencies n2 and mechanical properties of the detector [5]

69
2 2
2 n2 2 n2
kn2 , qn2 , (6.86)
+ 2

with being the density of the detector and and the Lame coefficients, which can be

described by the detectors Young modulus Y and Poisson ration

Y 2
= , = . (6.87)
2(1 + ) 1 2

Using the degenerate eigenfrequency fn = 3206.3 and the parameters used in Table

6.1, bn is found to be

bn = 0.402m = 2.04 106 eV , (6.88)

and using a negligible wave mass ms gives the quantum limit strain produced by the scalar

wave on the Mario Schenberg detector to be

Am=0 = 6.34 1024 Hz1/2 . (6.89)

This quantum limit is two orders of magnitude lower than the minimum strain of the

Schenberg detector shown in Figure 4.4.

6.5.2. Scalar Wave Mass Constraints

The mass of the scalar wave is limited by equation (5.9), that is

p
2 m2s = qs . (6.90)

For the scalar wave to have a real wave number qs , the mass cannot exceed the wave

frequency , so the mass in the previous section has natural limits of

70
0eV ms < 1.33 1011 eV . (6.91)

If one looks at cosmological constraints, the mass is further limited to be[14]

0eV ms 1033 eV , (6.92)

which is negligible compared to the frequency of detectable waves.

According to equations (6.78) and (6.79), a non-negligible mass would reduce the

minimum strain required to excite the monopole mode and increase the strain required to

excite the quadrupole. Thus, the difference between the resonant strengths in the monopole

and quadrupolar modes of the detector would provide an excellent test of contemporary

cosmological models.

71
CHAPTER 7: Conclusion

To review what has been covered, we have introduced the basic ideas of general

relativity and gravitational waves. We have also discussed the importance of GW astronomy

and how it can be used as a test of our understanding of the universe. For the first time, the

absorption cross section of GWs in f (R) gravity on spherical detectors has been calculated.

To calculate the absorption cross section for an f (R) GW, one had to first determine

the energy flux of the GW. The result of section 5.2 yielded the energy and momentum of a

GW in f (R) gravity. Knowing the energy-momentum tensor of a GW is useful for

determining how the wave affects an object, and it also has its uses in attempting to

quantize gravity[1]. Equation (5.32) is an independent calculation that has the versatility to

be applied to any f (R) model.

Chapter 6 has used the framework in [5, 34] to derive the f (R) GW physics on a

spherical detector. At the end of the chapter, we were able to apply these physics to

calculate the quantum limit minimum strain that could be registered by the Mario

Schenberg detector, and we were able to briefly study how the mass of a scalar wave would

affect its detectablility.

Future work includes considering the Sun as a spherical detector. Current work in

helioseismology allows the surface velocity amplitudes and frequencies of solar acoustic

modes to be determine with high precision[36]. Stars near massive black hole binaries can

act as GW-charged batteries, absorbing and emitting GWs produced by the binary

system, and GWs emitted by galactic white dwarf binary systems may be detectable of

observing the acoustic resonances of the Sun.[37] The work in chapter 6 is based on a GW

incident with a homogeneous sphere, so special considerations will have to be made for the
inhomogeneous structure of the Sun.

73
BIBLIOGRAPHY

[1] Steven Weinberg, Gravitation and Cosmology: The Principles and Applications of the

General Theory of Relativity, John Wiley & Sons, Inc. (1972)

[2] K. Riles, Prog.Part.Nucl.Phys. 68 (2013) 1-54, arXiv:1209.0667v3

[3] B. S. Sathyaprakash & B. F. Schutz, Living Rev. Relativity 12 (2009)

[4] Odylio Denys Aguiar, Res.Astron.Astrophys. 11 (2011) 1-42, arXiv:1009.1138

[5] J. A. Lobo, Phys.Rev. D52 (1995) 591, gr-qc/0006102

[6] T. Sotiriou & V. Faraoni, Rev.Mod.Phys. 82 (2010) 451-497, arXiv:0805.1726

[7] A. Riess et al., Astron.J. 116 (1998) 1009-1038, astro-ph/9805201

[8] S. Perlmutter et al., Astrophys.J. 517 (1999) 565-586, astro-ph/9812133

[9] P.A.R. Ade et al., Astron.Astrophys. 571 (2014) A16, 1303.5076

[10] T. Sotiriou, Class.Quant.Grav. 23 (2006) 5117-5128, gr-qc/0604028

[11] T. Sotiriou, arXiv:0710.4438

[12] A. De Felice & S. Tsujikawa, Living Rev.Rel. 13 (2010) 3, arXiv:1002.4928

[13] Salvatore Capozziello & Mariafelicia De Laurentis, Phys.Rept. 509 (2011) 167-321,

1108.6266

[14] L. Yang, C.-C. Lee, & C.-Q. Geng, JCAP 1108 (2011) 029, arXiv:1106.5582

[15] Sean Carroll, A No-Nonsense Introduction to General Relativity (2001)

[16] Sean Carroll, Lecture Notes on General Relativity (1997)

[17] Eanna E. Flanagan & Scott A. Hughes, New J.Phys. 7 (2005) 204, gr-qc/0501041

[18] L. P. Grishchuk, Sov. J. Exp. Theor. Phys. 40, 409 (1975)

[19] A. A. Starobinsky, Zh. Eksp. Teor. Fiz. Pisma Red. 30, 719 (1979).

[20] V. A. Rubakov, M. V. Sazhin, and A. V. Veryaskin, Phys. Lett. B 115, 189 (1982).
[21] R. Fabbri and M. D. Pollock, Phys. Lett. B 125, 445 (1983).

[22] L. F. Abbott and M. B. Wise, Nucl. Phys. B 244, 541 (1984).

[23] A. Ashoorioon, P. S. Bhupal Dev, and A. Mazumdar, Mod.Phys.Lett. A29 (2014) 30,

1450163, arXiv:1211.4678

[24] L. M. Krauss and F. Wilczek, Phys. Rev. D 89, 047501 (2014), arXiv:1309.5343

[25] R. Hulse & J. Taylor, 1975 APJ 195 51H

[26] J. Weisberg & J. Taylor, 2010 APJ 722 1030

[27] J. Weisberg & J. Taylor, ASP Conf.Ser. 328 (2005) 25, astro-ph/0407149

[28] S. Hild, Class Quantum Grav. 29 (2012) 124006

[29] Misner, Thorne, Wheeler, Gravitation, W. H. Freeman and Company (1973)

[30] Clifford M. Will, Living Rev.Rel. 17 (2014) 4, arXiv:1403.7377

[31] L. D. Landeau & E. M. Lifshitz Theory of Elasticity, Pergamon (1970)

[32] M. Boas, Mathmatical Methods in the Physicsal Sciences, John Wiley & Sons, Inc.

(1966)

[33] Cesar A. Costa, Odylio D. Aguiar, & Nadja S. Magalhaes, gr-qc/0312035

[34] M. Bianchi, M. Brunnetti, E. Coccia, F. Fucito, & J. A. Lobo, Phys.Rev. D57 (1998)

4525-4534, gr-qc/9709045

[35] C.M. Caves, K.S. Thorne, R.W.P. Drever, V.D. Sandberg, & M. Zimmermann,

Rev.Mod.Phys. 52 (1980) 341-392

[36] Ildio Lopes, Joseph Silk, Astrophys.J. 794 (2014) 32, arXiv:1406.1147

[37] B. McKernan, K.E.S. Ford, B. Kocsis, Z. Haiman, Mon.Not.Roy.Astron.Soc. 445 (2014)

74, arXiv:1405.1414

[38] Shinji Tsujikawa, Phys.Rev. D77 (2008) 023507, arXiv:0709.1391

75
[39] Alexei A. Starobinsky, JETP Lett. 86 (2007) 157-163 , arXiv:0706.2041

[40] Saeed Mirshekari, arXiv:1308.5240

[41] Ken-ichi Nakao, Tomohiro Harada, Masaru Shibata, Seiji Kawamura, & Takashi

Nakamura, Phys.Rev. D63 (2001) 082001, gr-qc/0006079

[42] Ch. Corda & S.A. Ali, C. Cafaro, Int.J.Mod.Phys. D19 (2010) 2095-2109,

arXiv:0902.0093

[43] A.Emir Gumrukcuoglu, Sachiko Kuroyanagi, Chunshan Lin, Shinji Mukohyama, &

Norihiro Tanahashi, Class.Quant.Grav. 29 (2012) 235026, arXiv:1208.5975

[44] J. A. Lobo, Lect.Notes Phys. 617 (2003) 210-241, gr-qc/0202063

[45] Danilo Babusci, Luca Baiotti, Francesco Fucito, & Alessandro Nagar, Phys.Rev. D64

(2001) 062001, gr-qc/0105028

[46] Christian Corda, Int.J.Mod.Phys. D18 (2009) 2275-2282, arXiv:0905.2502

[47] Christian Corda, Int.J.Mod.Phys. A23 (2008) 1521-1535, arXiv:0711.4917

[48] Justin Alsing, Emanuele Berti, Clifford M. Will, Helmut Zaglauer, Phys.Rev. D85

(2012) 064041, arXiv:1112.4903

[49] Lam Hui, Sean T. McWilliams, I-Sheng Yang, Phys.Rev. D87 (2013) 8, 084009,

arXiv:1212.2623

[50] E.O. Kahya, Phys.Lett. B701 (2011) 291-295, arXiv:1001.0725

[51] Michele Maggiore, Phys.Rept. 331 (2000) 283-367, gr-qc/9909001

[52] Nicols Yunes & Xavier Siemens, Living Rev.Rel. 16 (2013) 9, arXiv:1304.3473

[53] E. Coccia, J.A. Lobo, & J.A. Ortega, Phys.Rev. D52 (1995) 3735-3738

[54] E. Coccia, V. Fafone, G. Frossati, J.A. Lobo, & J.A. Ortega, Phys.Rev. D57 (1998)

2051-2060, gr-qc/9707059

76
[55] M. Bianchi, E. Coccia, C.N. Colacino, V. Fafone, & F. Fucito, Class.Quant.Grav. 13

(1996) 2865-2874, gr-qc/9604026

[56] C.F. Da Silva Costa & O.D. Aguiar, J.Phys.Conf.Ser. 484 (2014) 012012

[57] Michele Maggiore & Alberto Nicolis, Phys.Rev. D62 (2000) 024004, gr-qc/9907055

[58] J. A. Lobo, gr-qc/0006055

[59] Christian Corda, Mod.Phys.Lett. A22 (2007) 1727-1735, arXiv:0706.3782

[60] H.G. Khosroshahi & Y. Sobouti, Astron.Astrophys. 321, 1024-1025 (1997)

77
Sylvester, Alex J. Bachelor of Science, University of Louisiana at Lafayette, Spring 2013;
Master of Science, University of Louisiana at Lafayette, Fall 2015
Major: Physics
Title of Thesis: Testing General Relativity with Spherical Resonant Mass Detectors
Thesis Director: Dr. James B. Dent
Pages in Thesis: 87; Words in Abstract: 91

ABSTRACT

Gravitational waves in f (R) gravity excite monopole and m = 0, 2 quadrupole

resonance modes of a spherical detector. This document reviews the basic ideas of general

relativity and gravitational waves, and then applies those concepts to an f (R) gravitational

wave. The acoustic responce of a GW incident with a spherical detector is reviewed in

detail, and the absorption cross section for an f (R) GW impinging on the spherical detector

is calculated. Minimum detectable scalar wave amplitudes are explored for the Mario

Schenberg detector. The mass of the scalar mode affects its detectability.
BIOGRAPHICAL SKETCH

Alex J. Sylvester was born in Opelousas, Louisiana on October 2, 1989. He was born

into a family of rice, soybean, and crawfish farmers from the rural community of Whiteville,

Louisiana and is the third of six children.

After graduating from Sacred Heart High School in Ville Platte, Louisiana, Sylvester

started his undergraduate studies in June of 2008 at LSU-Eunice. He transferred to the

University of Louisiana at Lafayette in August of 2009, graduating with a Bachelor of

Science in Physics in May of 2013. He began pursuing a Master of Science in Physics at the

University of Louisiana at Lafayette in August of 2013 and completed the degree in

December of 2015.

S-ar putea să vă placă și