12.07.2015 Views

(ed.). Gravitational waves (IOP, 2001)(422s).

(ed.). Gravitational waves (IOP, 2001)(422s).

(ed.). Gravitational waves (IOP, 2001)(422s).

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

GRAVITATIONAL WAVES


GRAVITATIONAL WAVESEdit<strong>ed</strong> byIgnazio CiufoliniDepartment of Engineering,University of LecceVittorio Gorini, Ugo MoschellaDepartment of Chemical, Mathematical and Physical Sciences,University of Insubria at ComoandPietro FréDepartment of Physics, University of TurinINSTITUTE OF PHYSICS PUBLISHINGBRISTOL AND PHILADELPHIA


c­ <strong>IOP</strong> Publishing Ltd <strong>2001</strong>All rights reserv<strong>ed</strong>. No part of this publication may be reproduc<strong>ed</strong>, stor<strong>ed</strong>in a retrieval system or transmitt<strong>ed</strong> in any form or by any means, electronic,mechanical, photocopying, recording or otherwise, without the prior permissionof the publisher. Multiple copying is permitt<strong>ed</strong> in accordance with the termsof licences issu<strong>ed</strong> by the Copyright Licensing Agency under the terms of itsagreement with the Committee of Vice-Chancellors and Principals.British Library Cataloguing-in-Publication DataA catalogue record for this book is available from the British Library.ISBN 0 7503 0741 2Library of Congress Cataloging-in-Publication Data are availableCommissioning Editor: James RevillProduction Editor: Simon LaurensonProduction Control: Sarah PlentyCover Design: Victoria Le BillonMarketing Executive: Colin FentonPublish<strong>ed</strong> by Institute of Physics Publishing, wholly own<strong>ed</strong> by The Institute ofPhysics, LondonInstitute of Physics Publishing, Dirac House, Temple Back, Bristol BS1 6BE, UKUS Office: Institute of Physics Publishing, The Public L<strong>ed</strong>ger Building, Suite1035, 150 South Independence Mall West, Philadelphia, PA 19106, USATypeset in TEX using the <strong>IOP</strong> Bookmaker MacrosPrint<strong>ed</strong> in the UK by MPG Books Ltd, Bodmin


ContentsPrefacexiii1 <strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview) 1References 9PART 1<strong>Gravitational</strong> Waves, Sources and DetectorsBernard F Schutz and Franco Ricci 11Synopsis 132 Elements of gravitational <strong>waves</strong> 152.1 Mathematics of lineariz<strong>ed</strong> theory 162.2 Using the TT gauge to understand gravitational <strong>waves</strong> 172.3 Interaction of gravitational <strong>waves</strong> with detectors 192.4 Analysis of beam detectors 212.4.1 Ranging to spacecraft 212.4.2 Pulsar timing 222.4.3 Interferometry 222.5 Exercises for chapter 2 223 <strong>Gravitational</strong>-wave detectors 243.1 <strong>Gravitational</strong>-wave observables 263.2 The physics of interferometers 283.2.1 New interferometers and their capabilities 323.3 The physics of resonant mass detectors 343.3.1 New bar detectors and their capabilities 373.4 A detector in space 383.4.1 LISA’s capabilities 393.5 <strong>Gravitational</strong> and electromagnetic <strong>waves</strong> compar<strong>ed</strong> and contrast<strong>ed</strong> 41


viContents4 Astrophysics of gravitational-wave sources 434.1 Sources detectable from ground and from space 434.1.1 Supernovae and gravitational collapse 434.1.2 Binary stars 444.1.3 Chirping binary systems 444.1.4 Pulsars and other spinning neutron stars 464.1.5 Random backgrounds 484.1.6 The unexpect<strong>ed</strong> 495 Waves and energy 505.1 Variational principle for general relativity 505.2 Variational principles and the energy in gravitational <strong>waves</strong> 515.2.1 Gauge transformation and invariance 525.2.2 <strong>Gravitational</strong>-wave action 525.3 Practical applications of the Isaacson energy 545.3.1 Curvature produc<strong>ed</strong> by <strong>waves</strong> 555.3.2 Cosmological background of radiation 555.3.3 Other approaches 565.4 Exercises for chapter 5 566 Mass- and current-quadrupole radiation 586.1 Expansion for the far field of a slow-motion source 586.2 Application of the TT gauge to the mass quadrupole field 606.2.1 The TT gauge transformations 606.2.2 Quadrupole field in the TT gauge 616.2.3 Radiation patterns relat<strong>ed</strong> to the motion of sources 626.3 Application of the TT gauge to the current-quadrupole field 646.3.1 The field at third order in slow-motion 646.3.2 Separating the current quadrupole from the mass octupole 656.3.3 A model system radiating current-quadrupole radiation 676.4 Energy radiat<strong>ed</strong> in gravitational <strong>waves</strong> 686.4.1 Mass-quadrupole radiation 696.4.2 Current-quadrupole radiation 696.5 Radiation in the Newtonian limit 707 Source calculations 717.1 Radiation from a binary system 717.1.1 Corrections 737.2 The r-modes 737.2.1 Linear growth of the r-modes 767.2.2 Nonlinear evolution of the star 777.2.3 Detection of r-mode radiation 797.3 Conclusion 80References 81Solutions to exercises 84


ContentsviiPART 2<strong>Gravitational</strong>-wave detectorsGuido Pizzella, Angela Di Virgilio, Peter Bender and Francesco Fucito 898 Resonant detectors for gravitational <strong>waves</strong> and their bandwidth 918.1 Sensitivity and bandwidth of resonant detectors 918.2 Sensitivity for various GW signals 958.3 Recent results obtain<strong>ed</strong> with the resonant detectors 998.4 Discussion and conclusions 101References 1029 The Earth-bas<strong>ed</strong> large interferometer Virgo and the Low FrequencyFacility 1039.1 Introduction 1039.1.1 Interferometer principles and Virgo parameters 1049.2 The SA suspension and requirements on the control 1089.3 A few words about the Low Frequency Facility 1119.4 Conclusion 112References 11410 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave mission 11510.1 Description of the LISA mission 11510.1.1 Introduction 11510.1.2 Overall antenna and spacecraft design 11610.1.3 Optics and interferometry system 12110.1.4 Free mass sensors 12510.1.5 Micronewton thrusters 12910.1.6 Mission scenario 13110.2 Expect<strong>ed</strong> gravitational-wave results from LISA 13210.2.1 LISA sensitivity and galactic sources 13210.2.2 Origin of massive black holes 13610.2.3 Massive black holes in normal galaxies 13810.2.4 Structure formation and massive black hole coalescence 14110.2.5 Fundamental physics tests with LISA 14310.2.6 Future prospects 146Acknowl<strong>ed</strong>gments 148References 14811 Detection of scalar gravitational <strong>waves</strong> 15211.1 Introduction 15211.2 Testing theories of gravity 15411.2.1 Free vibrations of an elastic sphere 15411.2.2 Interaction of a metric GW with the sphere vibrationalmodes 15511.2.3 Measurements of the sphere vibrations and wavepolarization states 157


viiiContents11.3 <strong>Gravitational</strong> wave radiation in the JBD theory 15911.3.1 Scalar and Tensor GWs in the JBD Theory 16011.3.2 Power emitt<strong>ed</strong> in GWs 16111.3.3 Power emitt<strong>ed</strong> in scalar GWs 16211.3.4 Scalar GWs 16411.3.5 Detectability of the scalar GWs 16511.4 The hollow sphere 16811.5 Scalar–tensor cross sections 170Acknowl<strong>ed</strong>gments 176References 176PART 3The Stochastic <strong>Gravitational</strong>-Wave BackgroundD Babusci, S Foffa, G Losurdo, M Maggiore, G Matone and R Sturani 17912 Generalities on the stochastic GW background 18112.1 Introduction 18112.2 Definitions 18412.2.1 gw ( f ) and the optimal SNR 18412.2.2 The characteristic amplitude 18712.2.3 The characteristic noise level 18912.3 The overlap r<strong>ed</strong>uction function 19112.3.1 Two interferometers 19312.3.2 Interferometer—bar 19612.3.3 Interferometer—sphere 19612.4 Achievable sensitivities to the SGWB 19712.4.1 Single detectors 19712.4.2 Two detectors 19912.4.3 More than two detectors 20412.5 Observational bounds 20713 Sources of SGWB 21113.1 Topological defects 21113.1.1 Strings 21413.1.2 Hybrid defects 22113.2 Inflation 22313.2.1 Classical picture 22413.2.2 Calculation of the spectrum 22513.3 String cosmology 22913.3.1 The model 23013.3.2 Observational bounds to the spectrum 23413.4 First-order phase transitions 23513.5 Astrophysical sources 237References 239


ContentsixPART 4Theoretical developmentsHermann Nicolai, Alessandro Nagar, Donato Bini, Fernando De Felice,Maurizio Gasperini and Luc Blanchet 24314 Infinite-dimensional symmetries in gravity 24514.1 Einstein theory 24514.1.1 Introduction 24514.1.2 Mathematical conventions 24514.1.3 The Einstein–Hilbert action 24714.1.4 Dimensional r<strong>ed</strong>uction D = 4 → D = 3 24714.1.5 Dimensional r<strong>ed</strong>uction D = 3 → D = 2 24814.2 Nonlinear σ -models 25214.2.1 Ehlers Lagrangian as a nonlinear σ -model 25414.2.2 The Ernst equation 25514.2.3 The Matzner–Misner Lagrangian as a nonlinear σ -model 25514.3 Symmetries of nonlinear σ -models 25714.3.1 Nonlinear realization of SL(2, R) E 25714.3.2 Nonlinear realization of SL(2, R) MM 25814.4 The Geroch group 25914.4.1 Action of SL(2, R) E on ˜λ, B 2 25914.4.2 Action of SL(2, R) MM on λ, B 26014.4.3 The affine Kac–Moody SL(2, R) algebra 26014.5 The linear system 26114.5.1 Solving Einstein’s equations 26114.5.2 The linear system 26314.5.3 Derivation of the colliding plane metric by factorization 265Acknowl<strong>ed</strong>gments 267Further reading 26715 Gyroscopes and gravitational <strong>waves</strong> 26815.1 Introduction 26815.2 Splitting formalism and test particle motion: a short review 26915.3 The spacetime metric 27215.4 Searching for an operational frame 27415.5 Precession of a gyroscope in geodesic motion 27615.6 Conclusions 278References 27816 Elementary introduction to pre-big bang cosmology and to the relicgraviton background 28016.1 Introduction 28016.2 Motivations: duality symmetry 28316.3 Kinematics: shrinking horizons 28916.4 Open problems and phenomenological consequences 294


xContents16.5 Cosmological perturbation theory 29716.5.1 Choice of the frame 29716.5.2 Choice of the gauge 29916.5.3 Normalization of the amplitude 30216.5.4 Computation of the spectrum 30416.6 The relic graviton background 30916.7 Conclusion 316Acknowl<strong>ed</strong>gments 316Appendix A. The string effective action 317Appendix B. Duality symmetry 322Appendix C. The string cosmology equations 328References 33317 Post-Newtonian computation of binary inspiral waveforms 33817.1 Introduction 33817.2 Summary of optimal signal filtering 34017.3 Newtonian binary polarization waveforms 34317.4 Newtonian orbital phase evolution 34617.5 Post-Newtonian wave generation 34917.5.1 Field equations 34917.5.2 Source moments 35017.5.3 Radiative moments 35217.6 Inspiral binary waveform 354References 356PART 5Numerical relativityEdward Seidel 35918 Numerical relativity 36118.1 Overview 36118.2 Einstein equations for relativity 36318.2.1 Constraint equations 36518.2.2 Evolution equations 36718.3 Still newer formulations: towards a stable evolution system 36918.3.1 General relativistic hydrodynamics 37618.3.2 Boundary conditions 37818.3.3 Special difficulties with black holes 37918.4 Tools for analysing the numerical spacetimes 38218.4.1 Horizon finders 38218.4.2 Locating the apparent horizons 38318.4.3 Locating the event horizons 38518.4.4 Wave extraction 38618.5 Computational science, numerical relativity, and the ‘Cactus’ code 388


Contentsxi18.5.1 The computational challenges of numerical relativity 38818.6 Cactus computational toolkit 38918.6.1 Adaptive mesh refinement 39118.7 Recent applications and progress 39218.7.1 Evolving pure gravitational <strong>waves</strong> 39218.7.2 Black holes 39418.8 Summary 399Acknowl<strong>ed</strong>gments 40018.9 Further reading 40018.9.1 Overviews/formalisms of numerical relativity 40018.9.2 Numerical techniques 40118.9.3 Gauge conditions 40118.9.4 Black hole initial data 40118.9.5 Black hole evolution 40118.9.6 Black hole excision 40218.9.7 Perturbation theory and waveform extraction 40218.9.8 Event and apparent horizons 40218.9.9 Pure gravitational <strong>waves</strong> 40318.9.10 Numerical codes 403References 403Index 409


Preface<strong>Gravitational</strong> <strong>waves</strong> today represent a hot topic, which promises to play a centralrole in astrophysics, cosmology and theoretical physics.Technological developments have l<strong>ed</strong> us to the brink of their directobservation, which could become a reality in the coming years.The direct observation of gravitational <strong>waves</strong> will open an entirely newfield; gravitational wave astronomy. This is expect<strong>ed</strong> to bring a revolution inour knowl<strong>ed</strong>ge of the universe by allowing the observation of hitherto unseenphenomena such as coalescence of compact objects (neutron stars and blackholes), fall of stars into supermassive black holes, stellar core collapses, big-bangrelics and the new and unexpect<strong>ed</strong>.During Spring 1999, the SIGRAV—Società Italiana di Relatività eGravitazione (Italian Society of Relativity and Gravitation) sponsor<strong>ed</strong> theorganization of a doctoral school on ‘<strong>Gravitational</strong> Waves in Astrophysics,Cosmology and String Theory’, which took place at the Center for ScientificCulture ‘Alessandro Volta’ locat<strong>ed</strong> in the beautiful environment of Villa Olmoin Como, Italy.This book brings together the courses given at the school and providesa comprehensive review of gravitational <strong>waves</strong>. It includes a wide range ofcontributions by leading scientists in the field. Topics cover<strong>ed</strong> are: the basicsof GW with some recent advanc<strong>ed</strong> topics, GW detectors, the astrophysics of GWsources, numerical applications and several recent theoretical developments. Thematerial is written at a level suitable for postgraduate students entering the field.The main financial support for the School came from the Universityof Insubria at Como-Varese. Other contributors were the Department ofChemical, Physical and Mathematical Sciences of the same University, thePhysics Departments of the Universities of Milan and Turin, and the Instituteof Physics of Interplanetary Space—CNR, Frascati.We are grateful to all the members of the scientific organizing committeeand to the scientific coordinator of Centro Volta, Professor G Casati, for theirinvaluable help.We also acknowl<strong>ed</strong>ge the essential organizational support of the secretarialconference staff of Centro Volta, in particular of Chiara Stefanetti.I Ciufolini, V Gorini, U Moschella and P FréComo12 June 2000xiii


Chapter 1<strong>Gravitational</strong> <strong>waves</strong>, theory and experiment(an overview)Ignazio Ciufolini 1 and Vittorio Gorini 21 Dipartimento di Ingegneria dell’Innovazione, University ofLecce, ItalyE-mail: ciufoli@nero.ing.uniroma1.it2 Department of Chemical, Mathematical and Physical SciencesUniversity of Insubria at Como, ItalyE-mail: gorini@fis.unico.itGeneral relativity and electrodynamics display profound similarities and yetfundamental differences [1, 2]. In this connection, it may be interesting to pointout some historical analogies between the two fields.The enormous success of Maxwell’s equations did not rest only in thefact that they incorporat<strong>ed</strong>, together with the Lorentz force equation, all thelaws of electricity and magnetism, but also that on their basis James ClerkMaxwell (1831–1879) was able (in 1873) to pr<strong>ed</strong>ict the existence of a solutionconsisting of electric and magnetic fields changing in time, carrying energy andpropagating with spe<strong>ed</strong> c in vacuum: the electromagnetic <strong>waves</strong>. Nevertheless,some distinguish<strong>ed</strong> physicists, such as Lord Kelvin, had serious doubts about theexistence of such <strong>waves</strong>: ‘The so-call<strong>ed</strong> “electromagnetic theory of light” has nothelp<strong>ed</strong> us hitherto . . . it seems to me that it is rather a backward step . . . the onething about it that seems intelligible to me, I do not think is admissible . . . thatthere should be an electric displacement perpendicular to the line of propagation’.However, in 1887, eight years after Maxwell’s death, electromagnetic <strong>waves</strong>were both generat<strong>ed</strong> and detect<strong>ed</strong> by Heinrich Hertz (1857–1894); then, in 1901,Guglielmo Marconi transmitt<strong>ed</strong> and receiv<strong>ed</strong> signals across the Atlantic Ocean.In the twentieth century the detection and study of electromagnetic <strong>waves</strong>,other than visible light, open<strong>ed</strong> a new era of dramatic changes in the knowl<strong>ed</strong>ge ofour universe: cosmic radio <strong>waves</strong>, discover<strong>ed</strong> in the 1930s, reveal<strong>ed</strong> in subsequent1


2 <strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview)decades colliding galaxies; quasars with dimensions of the order of the solarsystem but having luminosities orders of magnitude larger than our galaxy;enormous jets from galactic nuclei and quasars reaching lengths of hundr<strong>ed</strong>s ofthousands of light years, rapidly rotating pulsars with rotational periods of a fewmilliseconds and, not least, the cosmic microwave background, a relic of the hotbig bang. X-rays reveal<strong>ed</strong> accretion disks about black holes and neutron stars.Similarly, millimetre, infrar<strong>ed</strong> and ultraviolet radiation, and gamma rays open<strong>ed</strong>other dramatic windows of knowl<strong>ed</strong>ge on our universe.In the same way, in general relativity [1, 2], Einstein’s field equations(1915) not only describ<strong>ed</strong> the gravitational interaction via the spacetime curvaturegenerat<strong>ed</strong> by mass-energy, but also contain<strong>ed</strong>, through the Bianchi identities, theequations of motion of matter and fields, and on their basis Albert Einstein,in 1916, a few months after the formulation of the theory, pr<strong>ed</strong>ict<strong>ed</strong> theexistence of curvature perturbations propagating with spe<strong>ed</strong> c on a flat and emptyspacetime; the gravitational <strong>waves</strong> [4]. Einstein’s gravitational-wave theory was alineariz<strong>ed</strong> theory treating weak <strong>waves</strong> as weak perturbations of a flat background[1, 3, 5]. Similarly to what happen<strong>ed</strong> when electromagnetic <strong>waves</strong> were firstpr<strong>ed</strong>ict<strong>ed</strong>, some distinguish<strong>ed</strong> physicists had serious doubts about their existence.Arthur Eddington thought that these weak-field solutions of the wave equationobtain<strong>ed</strong> from Einstein’s field equations were just coordinate changes which were‘propagating . . . with the spe<strong>ed</strong> of thought’ [6].The lineariz<strong>ed</strong> theory of gravitational <strong>waves</strong> had its limits because the linearapproximation is not valid for sources where gravitational self-energy is notnegligible. It was only in 1941 that Landau and Lifshitz [7] describ<strong>ed</strong> the emissionof gravitational <strong>waves</strong> by a self-gravitating system of slowly moving bodies.However, in the following years there were serious doubts about the reality ofgravitational <strong>waves</strong> and not until 1957 did a g<strong>ed</strong>anken experiment by HermannBondi show that gravitational <strong>waves</strong> do inde<strong>ed</strong> carry energy [8].This thought experiment was bas<strong>ed</strong> on a system of two beads sliding ona stick with only a slight friction opposing their motion. If a plane gravitationalwave impinges on this system, the beads move back and forth on the stick becauseof the change in the proper distance between them due to the change of themetric, i.e. to the gravitational-wave perturbation; this change is govern<strong>ed</strong> by thegeodesic deviation equation and the proper dispacement between the two beadsis a function of the gravitational-wave metric perturbation. Thus, the frictionbetween beads and stick heats the system and thus increases the temperature ofthe stick. Therefore, since there is an energy transfer from gravitational <strong>waves</strong>to the system in the form of increas<strong>ed</strong> temperature of the system, this thoughtexperiment show<strong>ed</strong> that gravitational <strong>waves</strong> do inde<strong>ed</strong> carry energy and are a realphysical entity [1, 8].It is interesting to note that in 1955 John Archibald Wheeler had devis<strong>ed</strong>the conceivable existence of a body with no ‘mass’ built up by gravitationalor electromagnetic, radiation alone [9]. Inde<strong>ed</strong>, an object can, in principle,be construct<strong>ed</strong> out of gravitational radiation or electromagnetic radiation, or


<strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview) 3a mixture of the two, and may hold itself together by its own gravitationalattraction. A collection of radiation held together in this way, is call<strong>ed</strong> a geon(gravitational electromagnetic entity) and studi<strong>ed</strong> from a distance, such an objectwould present the same kind of gravitational attraction as any other mass. Yet,nowhere inside the geon is there a place where there is ‘mass’ in the conventionalsense of the term. In particular, for a geon made of pure gravitational radiation—a gravitational geon—there is no local measure of energy, yet there is globalenergy. The gravitational geon owes its existence to a dynamical localiz<strong>ed</strong>—buteverywhere regular—curvature of spacetime, and to nothing more. Thus, a geonis a collection of electromagnetic or gravitational-wave energy, or a mixture ofthe two, held together by its own gravitational attraction, that was describ<strong>ed</strong> byWheeler as ‘mass without mass’.In the 1960s, Joseph Weber began the experimental work to detectgravitational <strong>waves</strong>. He was essentially alone in this field of research [10]. Then,the theoretical work of Wheeler, Bondi, Landau and Lifshitz, Isaacson, Thorneand others and the experimental work of Weber, Braginski, Amaldi and othersopen<strong>ed</strong> a new era of research in this field. In 1972 Steven Weinberg wrote ‘. . .gravitational radiation would be interesting even if there were no chance of everdetecting any, for the theory of gravitational radiation provides a crucial linkbetween general relativity and the microscopic frontiers of physics’ [11].Today gravitational <strong>waves</strong>, both theory and experiment, are one of the maintopics of research in general relativity and gravitation [3].In the same way as electromagnetic <strong>waves</strong> other than visible light, that isradio, millimetre, infrar<strong>ed</strong>, ultraviolet, x-ray and gamma-ray astronomy open<strong>ed</strong>new windows and brought radical changes in our knowl<strong>ed</strong>ge of the universe,gravitational-wave astronomy is expect<strong>ed</strong> to bring a revolution in our knowl<strong>ed</strong>geof the universe by observing new exotic phenomena such as formation andcollision of black holes, fall of stars into supermassive black holes, primordialgravitational <strong>waves</strong> emitt<strong>ed</strong> just after the big bang .... Nevertheless, today,about 85 years after the pr<strong>ed</strong>iction of gravitational <strong>waves</strong> by Einstein, the onlyevidence for their actual existence is indirect and comes from the observationof the energy loss from the binary pulsar system PSR 1913+16, discover<strong>ed</strong> in1974 by Hulse and Taylor [12]. Quite remarkably, though of no surprise, theobserv<strong>ed</strong> energy loss of the binary pulsar is in agreement with the theoreticalpr<strong>ed</strong>iction by general relativity for the energy loss by gravitational radiationemitt<strong>ed</strong> by a binary system, to within less than 0.3% error (in this respect, itmight be interesting to note here that, in regard to the field that in generalrelativity is formally analogous to the magnetic field in electrodynamics, i.e. theso-call<strong>ed</strong> gravitomagnetic field, pr<strong>ed</strong>ict<strong>ed</strong> by Lense and Thirring in 1916, thefirst evidence and measurement of the existence of such an effect on Earth’ssatellites, due to the Earth’s rotation, was publish<strong>ed</strong> only in 1996, that is 80years after the derivation of the effect [13]). Thus, today, together with theenormous experimental efforts to detect gravitational <strong>waves</strong>, from bar detectorsto laser interferometers on Earth, GEO-600, LIGO, VIRGO, ..., and from laser


4 <strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview)interferometers in space, LISA, to Doppler tracking of interplanetary spacecrafts,there is, aim<strong>ed</strong> at increasing the chances of future detections, a strongly relat<strong>ed</strong>theoretical and computational work to understand and pr<strong>ed</strong>ict the emission orgravitational <strong>waves</strong> from astrophysical systems in strong field conditions [3]. Inthis book contributions of leading experts in the field of gravitational <strong>waves</strong>, boththeoretical and experimental, are present<strong>ed</strong>.The basic contribution by Bernard Schutz and Franco Ricci deals with themain features of gravitational <strong>waves</strong>, sources and detectors. The contributionis divid<strong>ed</strong> into six chapters and some chapters are follow<strong>ed</strong> by a few exercises.The first chapter describes the lineariz<strong>ed</strong> theory and the fundamental propertiesof weak gravitational <strong>waves</strong>, perturbations of a flat background, analys<strong>ed</strong> inthe so-call<strong>ed</strong> transverse-traceless gauge. The second and third chapters dealwith detectors and astrophysical sources; in particular an overview is present<strong>ed</strong>of the most important detectors under construction (their physics, sensitivityand opportunity for the future) and the main expect<strong>ed</strong> sources of gravitational<strong>waves</strong>, such as binary systems, neutron stars, pulsars, γ -ray bursts, etc. Thefourth chapter deals with the mathematical theory of <strong>waves</strong> in general, stressenergytensor and energy carri<strong>ed</strong> by gravitational <strong>waves</strong>. The subsequent chapterdescribes radiation generation in lineariz<strong>ed</strong> theory: mass- and current-quadrupoleradiation, i.e. the quadrupole formulae for the outgoing flux of gravitational-waveenergy emitt<strong>ed</strong> by a system characteriz<strong>ed</strong> by slow motion. Finally, the last chapterdescribes some applications of radiation theory to some sources: binary systemsand especially r-modes of neutron stars.The contribution by Guido Pizzella deals with bar detectors of gravitational<strong>waves</strong>. A gravitational-wave resonant detector is usually a cylindrical barof length L. The small change δL in the length of the whole bar at thefundamental resonance angular frequency, ω 0 , can be describ<strong>ed</strong> by the solution ofthe equation of a harmonic oscillator, with resonance angular frequency ω 0 (witha supplementary 4/π 2 factor obtain<strong>ed</strong> by solving the problem of a continuousbar). In a gravitational-wave resonant detector the mechanical oscillations ofthe bar induc<strong>ed</strong> by a gravitational wave are convert<strong>ed</strong> by an electromechanicaltransducer into electric signals which are amplifi<strong>ed</strong> with a low noise amplifier,such as a dc SQUID. Then the data analysis is perform<strong>ed</strong>. Using a resonantantenna one measures the Fourier component of the metric perturbation nearthe antenna resonance frequency ω 0 . The typical damping time of the resonantdetector is 2Q/ω 0 , where Q is the so-call<strong>ed</strong> quality factor of the resonant detector.The ultimate sensitivity of bar antennae to a fractional change in dimension dueto a short burst of gravitational radiation has been estimat<strong>ed</strong> to be of the orderof 10 −20 or 10 −21 . Bar detectors, usually 3 m long aluminum bars, work at atypical frequency of about 10 3 Hz. Resonant antennae were first built by J Weber,around 1960, at the University of Maryland. Subsequently, gravitational-waveresonant detectors have been operat<strong>ed</strong> by the following universities: Beijing,Guangzhou, Louisiana, Maryland, Moscow, Rome, Padua, Stanford, Tokyo andWestern Australia at Perth. The contribution of Pizzella deals with the bandwidth


<strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview) 5and the sensitivities of resonant detectors. It is shown that it might be possible toreach a frequency bandwidth up to 50 Hz. The sensitivity of five cryogenic bardetectors in operation, ALLEGRO, AURIGA, EXPLORER, NAUTILUS ANDNIOBE is then discuss<strong>ed</strong>.The paper by Angela Di Virgilio treats laser interferometers on Earthand in particular the Italian–French antenna VIRGO. <strong>Gravitational</strong>-wave laserinterferometerson Earth will operate in the frequency range between 10 4 Hz anda few tens of hertz. Various types of gravitational-wave laser interferometers havebeen propos<strong>ed</strong>, among which are the standard Michelson and Fabry–Perot types.A Michelson-type gravitational-wave laser interferometer is essentially made ofthree masses suspend<strong>ed</strong> with wires at the ends of two orthogonal arms of lengthl. When a gravitational-wave with r<strong>ed</strong>uc<strong>ed</strong> wavelength λ GW≫ l is impinging,for example, perpendicularly to this system, variations in the metric perturbationh due to the gravitational wave will, in turn, produce oscillations in the differencebetween the proper lengths of the two arms δl(t) and therefore oscillations in therelative phase of the laser light at the beamsplitter; thus, they will finally produceoscillations in the intensity of the laser light measur<strong>ed</strong> by the photodetector. Ifthe laser light will travel back and forth between the test masses 2N times (N =number of round trips), then the variation of the difference between the properlengths of the two arms will be (assuming Nl ≪ λ GW): l = 2Nlh(t), andtherefore, the relative phase delay due to the variations in δl will be:φ = lλ ¯L= 2Nl h(t),λ ¯Lwhere λ ¯L is the r<strong>ed</strong>uc<strong>ed</strong> wavelength of the laser light.For most of the fundamental limiting factors of these Earth-bas<strong>ed</strong> detectors,such as seismic noise, photon shot noise, etc ..., the displacement noise isessentially independent from the arms length l. Therefore, by increasing l oneincreases the sensitivity of the detectors.Two antennae with 4 km arm lengths in the USA, the MIT and CaltechLIGOs, should reach sensitivities to bursts of gravitational radiation of the orderof h ∼ 10 −20 –10 −21 between 1000 and 100 Hz. GEO-600 is an underground600 m laser interferometer built by the University of Glasgow and the Max-Planck-Institutes for Quantum Optics and for Astrophysics at Garching. TAMAis a 300 m antenna in Japan and ACIGA is a 3 km antenna plann<strong>ed</strong> in Australia.The paper of Di Virgilio describes the 3 km laser interferometer VIRGO, builtby INFN of Pisa together with the University of Paris-Sud at Orsay, that shouldreach frequencies of operation as low as a few tens of hertz, using special filtersto eliminate the seismic noise at these lower frequencies. The ultimate burstsensitivity for all of the above large interferometers is currently estimat<strong>ed</strong> to be ofthe order of 10 −22 or 10 −23 at frequencies near 100 Hz.The paper of Peter Bender describes the space gravitational-wave detectorLISA (Laser Interferometer Space Antenna). Below about 10 Hz the sensitivityof Earth-bas<strong>ed</strong> gravitational-wave detectors is limit<strong>ed</strong> by gravity gradients


6 <strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview)variations. Even for perfect isolation of a detector from seismic and ground noise,an Earth-detector would still be affect<strong>ed</strong> by the time changes in the gravity fielddue to density variations in the Earth and its atmosphere. Due to this source ofnoise the sensitivity has been calculat<strong>ed</strong> to worsen as roughly the inverse fourthpower of the frequency.Therefore, to avoid this type of noise and to r<strong>ed</strong>uce noise from other sources,one should use an interferometer far from Earth and with very long arms. Inde<strong>ed</strong>,to detect gravitational <strong>waves</strong> in the range of frequencies between about 10 −4 and1 Hz, Bender proposes to orbit in the solar system a space interferometer made ofthree spacecraft at a typical distance from each other of 5000 000 km.Although the phase measurement system and the thermal stability areessential requirements, it is the main technological challenge of this experimentto keep very small the spurious accelerations of the test masses. A dragcompensating system will be able to largely r<strong>ed</strong>uce these spurious accelerations.Considering all the error sources, it has been calculat<strong>ed</strong> that, for periodicgravitational <strong>waves</strong>, with an integration time of about one year, LISA should reacha sensitivity able to detect amplitudes of h ∼ 10 −23 , in the range of frequenciesbetween 10 −3 and 10 −2 Hz, amplitudes from h ∼ 10 −20 to about 10 −23 between10 −4 and 10 −3 Hz, and amplitudes from h ∼ 10 −22 to about 10 −23 between 1and 10 −2 Hz.Therefore, comparing the LISA sensitivity to the pr<strong>ed</strong>ict<strong>ed</strong> theoreticalamplitudes of gravitational radiation at these frequencies, LISA should be ableto detect gravitational <strong>waves</strong> from galactic binaries, including ordinary mainsequencebinaries, contact binaries, cataclysmic variables, close white dwarfbinaries, neutron star and black hole binaries. The LISA sensitivity should alsoallow detection of possible gravitational pulses from distant galaxies from theinspiral of compact objects into supermassive black holes in Active GalacticNuclei and from collapse of very massive objects to form black holes. LISAshould also allow us to detect the stochastic background due to unresolv<strong>ed</strong> binarysystems.The contribution by Francesco Fucito treats spherical shape antennae andthe detection of scalar gravitational <strong>waves</strong>. General relativity pr<strong>ed</strong>icts only twoindependent states of polarization of a weak gravitational wave, the so-call<strong>ed</strong>‘×’ and ‘+’ ones. Nevertheless, metric theories of gravity alternative to generalrelativity and non-metric theories of gravity pr<strong>ed</strong>ict different polarization states(up to six components in metric theories [14]). For example, the Jordan–Brans–Dicke theory pr<strong>ed</strong>icts also an additional scalar component of a gravitationalwave and, as the author explains, string theory could also imply the existenceof other components. In this paper the possibility is describ<strong>ed</strong> of placing limitson, or detecting, these additional polarizations of a gravitational wave, thustesting theories of gravity alternative to general relativity, by using spherical shap<strong>ed</strong>etectors. Spheroidal detectors of gravitational <strong>waves</strong> of two types are discuss<strong>ed</strong>,standard and hollow spherical ones.The paper by Babusci, Foffa, Losurdo, Maggiore, Matone and Sturani


<strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview) 7treats stochastic gravitational <strong>waves</strong>. As the authors explain, the stochasticgravitational-wave background (SGWB) is a random background of gravitational<strong>waves</strong> without any specific sharp frequency component that might giveinformation about the very early stages of our universe. It is important to note thatrelic cosmological gravitational <strong>waves</strong> emitt<strong>ed</strong> near the big bang might provideunique information on our universe at a very early stage. Inde<strong>ed</strong>, as regards thecosmic microwave background radiation, electromagnetic <strong>waves</strong> decoupl<strong>ed</strong> a few10 5 years after the big bang, whereas relic cosmological gravitational <strong>waves</strong>, theauthors explain, might come from times as early as a few 10 −44 s. The authorsdiscuss that, in order to increase the chances of detecting a stochastic backgroundof gravitational <strong>waves</strong>, the correlation of the outputs between two, or more,detectors would be convenient. Thus, after discussing three different detectors:laser interferometers, cylindrical bars and spherical antennae, the authors presentvarious possibilities of correlation, between two laser interferometers (VIRGO,LIGOs, GEO-600 and TAMA-300), and between a laser interferometer and acylindrical bar (AURIGA, NAUTILUS, EXPLORER) or a spherical antenna; theyalso discuss correlation between more than two detectors.In the second part of this paper they discuss sources of the background ofstochastic gravitational <strong>waves</strong>: topological defects in the form of points, linesor surfaces, call<strong>ed</strong> monopoles, cosmic strings and domain walls. In particular,they discuss cosmic strings and hybrid defects; inflationary cosmological models;string cosmology; and first-order phase transitions which occurr<strong>ed</strong> in the earlystage of the expansion of the universe, for example in GUT-symmetry breakingand electroweak-symmetry breaking. Finally, they discuss astrophysical sourcesof stochastic gravitational <strong>waves</strong>. The conclusion is that the frequency domainof cosmological and astrophysical sources of stochastic gravitational <strong>waves</strong>might be very different and thus, the authors conclude, the astrophysicalbackgrounds might not mask the detection of a relic cosmological gravitationalwavebackground at the frequencies of the laser interferometers on Earth.The contribution by Nicolai and Nagar deals with the symmetry propertiesof Einstein’s vacuum field equations when the theory is r<strong>ed</strong>uc<strong>ed</strong> from four to twodimensions, namely in the presence of two independent spacelike commutingKilling vectors. Under these conditions, and using the vierbein formalism, theauthors show that one can use a Kaluza–Klein ansatz to rewrite the Einstein–Hilbert Lagrangian in the form of two different two-dimensionally r<strong>ed</strong>uc<strong>ed</strong>Lagrangians nam<strong>ed</strong> the Ehlers and Matzner–Misner ones, respectively, after thepeople who first introduc<strong>ed</strong> them. Each of these two Lagrangians representstwo-dimensional r<strong>ed</strong>uc<strong>ed</strong> gravity in the conformal gauge as given by a part ofpure two-dimensional gravity, characteriz<strong>ed</strong> by a conformal factor and a dilatonfield plus a ‘matter part’ given by two suitable bosonic fields. In either case, thematter part has a structure of a nonlinear sigma model with an SL(2,R)/SO(2)symmetry. These two different nonlinear symmetries can be combin<strong>ed</strong> into aunifi<strong>ed</strong> infinite-dimensional symmetry group of the theory, call<strong>ed</strong> the Gerochgroup, whose Lie algebra is an affine Kac–Moody algebra, and whose action on


8 <strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview)the matter fields is both nonlinear and non-local. The existence of such an infinit<strong>ed</strong>imensionalsymmetry guarantees that the two-dimensionally r<strong>ed</strong>uc<strong>ed</strong> nonlinearfield equations are integrable. This can be shown in a standard way by exploitingthe symmetry to prove the equivalence of the theory to a system of lineardifferential equations whose compatibility conditions yield just the nonlinearequations that one wants to solve. As an example of the application of themethod to the construction of exact solutions of the two-dimensionally r<strong>ed</strong>uc<strong>ed</strong>Einstein’s equations, the results are employ<strong>ed</strong> to derive the exact expression ofthe metric which describe colliding plane gravitational <strong>waves</strong> with collinear andnon-collinear polarization.Gasperini’s contribution deals with string cosmology and with the basic ideasof the so-call<strong>ed</strong> pre-big bang scenario of string cosmology. Then it treats theinteresting problem of observable effects in different cosmological models, andin particular the so-call<strong>ed</strong> background of relic gravitational <strong>waves</strong>, comparing itwith the expect<strong>ed</strong> sensitivities of the gravitational-wave detectors. The conclusionis that the sensitivity of the future advanc<strong>ed</strong> detectors of gravitational <strong>waves</strong> maybe capable of detecting the background of gravitational <strong>waves</strong> pr<strong>ed</strong>ict<strong>ed</strong> in thepre-big bang scenario of string cosmology and thus these detectors might testdifferent cosmological models and also string theory models.The paper by Bini and De Felice studies the problem of the behaviourof a test gyroscope on which a plane gravitational wave is impinging. Theauthors analyse whether there might be observable effects, i.e. a precession ofthe gyroscope with respect to a suitably defin<strong>ed</strong> frame of reference that is notFermi–Walker transport<strong>ed</strong>.The contribution by Luc Blanchet deals with the post-Newtoniancomputation of binary inspiral waveforms. In general relativity, the orbital phaseof compact binaries, when gravitational radiation emitt<strong>ed</strong> is consider<strong>ed</strong>, is notconstant as it is in the Newtonian calculation, but is a complex, nonlinear functionof time, depending on small post-Newtonian corrections. For the data analysison detectors, a formula containing at least the 3PN (third-post-Newtonian) orderbeyond the quadrupole formalism (see the contribution by Schutz and Ricci) isne<strong>ed</strong><strong>ed</strong>, that is a formula including terms of the order of (v/c) 6 (where v isa typical velocity in the source and c is the spe<strong>ed</strong> of light). Blanchet’s paperthus treats the derivation of the third-post-Newtonian formula for the emission ofgravitational radiation from a self-gravitating binary system.The paper by Ed Seidel deals with numerical relativity. Among theastrophysical sources of gravitational radiation that might be detect<strong>ed</strong> by laserinterferometers on Earth there is the spiralling coalescence of two black holesor neutron stars. However, gravitational <strong>waves</strong> are so weak at the detectors onEarth that, as Seidel explains in his paper, one ne<strong>ed</strong>s to know the waveform inorder to reliably detect them, in other words gravitational-wave signals can beinterpret<strong>ed</strong> and detect<strong>ed</strong> only by comparing the observational data with a set oftheoretically determin<strong>ed</strong> ‘waveform templates’. Unfortunately, we can solve theEinstein’s field equations (coupl<strong>ed</strong>, nonlinear partial differential equations) only


References 9in especially simple cases. Thus, to find solutions of the Einstein’s equations,for example in a system with emission of gravitational radiation, we ne<strong>ed</strong> to findnumerical solutions of these field equations, i.e. we ne<strong>ed</strong> numerical relativity.Nevertheless, even the numerical approach to the emission of gravitational <strong>waves</strong>in strong field is extremely difficult and computer-time consuming. For example,as Seidel explains, the computer simulation of the coalescence of a compact objectbinary will require several years of super-computer time. However, special codesto solve the complete set of Einstein’s equations have been design<strong>ed</strong> that run veryefficiently on large-scale parallel computers, in particular, one of these codes, theCactus Computational Toolkit is present<strong>ed</strong> in this paper. Then, after a descriptionof the numerical formulation of the theory of general relativity, constraintequations and evolution equations, the numerical techniques for solving theevolution equations are report<strong>ed</strong> and finally some recent applications, includinggravitational <strong>waves</strong> and the evolution and collisions of black holes, are present<strong>ed</strong>.It is important to note that there have been and there are large collaborations innumerical relativity, including: the NSF Black Hole Grand Challenge Project, theNASA Neutron Star Grand Challenge Project, the NCSA/Potsdam/WashingtonUniversity numerical relativity collaboration and a EU European collaboration often institutions.References[1] Misner C W, Thorne K S and Wheeler J A 1973 Gravitation (New York: Freeman)[2] Ciufolini I and Wheeler J A 1995 Gravitation and Inertia (Princeton, NJ: PrincetonUniversity Press)[3] Thorne K S 1987 Three Hundr<strong>ed</strong> Years of Gravitation <strong>ed</strong> S W Hawking and W Israel(Cambridge: Cambridge University Press)[4] Einstein A 1916 Preuss. Akad. Wiss. Berlin, Sitzber. 688[5] Grishchuk L P and Polnarev A G 1980 General Relativity and Gravitation, OneHundr<strong>ed</strong> Years After the Birth of Albert Einstein <strong>ed</strong> A Held (New York: Plenum)[6] Douglass D H and Braginsky V B 1979 General Relativity, An Einstein CentenarySurvey <strong>ed</strong> S W Hawking and W Israel (Cambridge: Cambridge University Press)[7] Landau L D and Lifshitz E M 1962 The Classical Theory of Fields (Reading, MA:Addison-Wesley)[8] Bondi H 1957 Nature 179 1072See also Bondi H and McCrea W H 1960 Proc. Camb. Phil. Soc. 56 410[9] Wheeler J A 1955 Phys. Rev. 97 511[10] Weber J 1961 General Relativity and <strong>Gravitational</strong> Waves (New York: Wiley–Interscience)[11] Weinberg S 1972 Gravitation and Cosmology (New York: Wiley)[12] Hulse R A and Taylor J H 1975 Astrophys. J. 195 L51


10 <strong>Gravitational</strong> <strong>waves</strong>, theory and experiment (an overview)[13] Ciufolini I et al 1996 Nuovo Cimento A 109 575See also Ciufolini I et al 1997 Class. Quantum Grav. 14 2701Ciufolini I et al 1998 Science 279 2100[14] Will C M 1993 Theory and Experiment in <strong>Gravitational</strong> Physics (Cambridge:Cambridge University Press)


PART 1GRAVITATIONAL WAVES, SOURCESAND DETECTORSBernard F Schutz 1 and Franco Ricci 21 Max Planck Institute for <strong>Gravitational</strong> Physics (Albert EinsteinInstitute), Golm bei Potsdam, Germany and Department of Physicsand Astronomy, Cardiff University, Wales2 University ‘La Sapienza’, Rome, Italy


Synopsis<strong>Gravitational</strong> <strong>waves</strong> and their detection are becoming increasingly importantboth for the theoretical physicist and the astrophysicist. In fact, technologicaldevelopments have enabl<strong>ed</strong> the construction of such sensitive detectors (barsand interferometers) that the detection of gravitational radiation could becomea reality during the next few years. In these lectures we give a brief overview ofthis interesting and challenging field of modern physics.The topics cover<strong>ed</strong> are divid<strong>ed</strong> into six lectures. We begin (chapter 2) bydescribing gravitational <strong>waves</strong> in lineariz<strong>ed</strong> general relativity, where one canexamine most of the basic properties of gravitational radiation itself; propagation,gauge invariance and interactions with matter (and in particular with detectors).The second lecture (chapter 3) deals with gravitational-wave detectors:how they operate, what their most important sources of noise are, and whatmechanisms are us<strong>ed</strong> to overcome noise. We report here on the most importantdetectors plann<strong>ed</strong> or under construction (both ground-bas<strong>ed</strong> and space-bas<strong>ed</strong>ones), their likely sensitivity and their prospects for making detections. Otherspeakers will go into much more detail on specific detectors, such as LISA.The third lecture (chapter 4) deals with the astrophysics of likely sourcesof gravitational <strong>waves</strong>: binary systems, neutron stars, pulsars, x-ray sources,supernovae/hypernovae, γ -ray bursts and the big bang. We estimate the expect<strong>ed</strong>wave amplitude h and the suitability of specific detectors for seeing <strong>waves</strong> fromeach source.The fourth lecture (chapter 5) is much more theoretical. Here we developthe mathematical theory of gravitational <strong>waves</strong> in general, their effective stressenergytensor, the energy carri<strong>ed</strong> by gravitational <strong>waves</strong>, and the energy in arandom wave field (gravitational background generat<strong>ed</strong> by the big bang).The fifth lecture (chapter 6) takes the theory further and examines thegeneration of gravitational radiation in lineariz<strong>ed</strong> theory. We show in som<strong>ed</strong>etail how both mass-quadrupole and current-quadrupole radiation is generat<strong>ed</strong>,including how characteristics of the radiation such as its polarization are relat<strong>ed</strong>to the motion of the source. Current-quadrupole radiation has become importantvery recently and may inde<strong>ed</strong> be one of the first forms of gravitational radiation tobe detect<strong>ed</strong>. We attempt to give a physical description of the way it is generat<strong>ed</strong>.13


14 SynopsisThe final lecture (chapter 7) explores applications of the theory we hav<strong>ed</strong>evelop<strong>ed</strong> to various sources. We calculate the quadrupole moment of a binarysystem, the energy radiat<strong>ed</strong> in the Newtonian approximation and the back-reactionon the orbit. We conclude with a brief introduction to the current-quadrupol<strong>ed</strong>riveninstability in the r-modes of neutron stars.Chapters 2 and 5 are follow<strong>ed</strong> by a few exercises to assist students.We presume the reader has some background in general relativity and itsmathematical tools in differential geometry, at the level of the introductorychapters of Schutz (1985). A list of references is present<strong>ed</strong> at the end of theselectures of sources suitable for further and background reading.


Chapter 2Elements of gravitational <strong>waves</strong>General relativity is a theory of gravity that is consistent with special relativity inmany respects, and in particular with the principle that nothing travels faster thanlight. This means that changes in the gravitational field cannot be felt everywhereinstantaneously: they must propagate. In general relativity they propagate atexactly the same spe<strong>ed</strong> as vacuum electromagnetic <strong>waves</strong>: the spe<strong>ed</strong> of light.These propagating changes are call<strong>ed</strong> gravitational <strong>waves</strong>.However, general relativity is a nonlinear theory and there is, in general, nosharp distinction between the part of the metric that represents the <strong>waves</strong> andthe rest of the metric. Only in certain approximations can we clearly definegravitational radiation. Three interesting approximations in which it is possibleto make this distinction are:• lineariz<strong>ed</strong> theory;• small perturbations of a smooth, time-independent background metric;• post-Newtonian theory.The simplest starting point for our discussion is certainly lineariz<strong>ed</strong> theory,which is a weak-field approximation to general relativity, where the equations arewritten and solv<strong>ed</strong> in a nearly flat spacetime. The static and wave parts of thefield cleanly separate. We idealize gravitational <strong>waves</strong> as a ‘ripple’ propagatingthrough a flat and empty universe.This picture is a simple case of the more general ‘short-wave approximation’,in which <strong>waves</strong> appear as small perturbations of a smooth background that is tim<strong>ed</strong>ependent and whose radius of curvature is much larger than the wavelength of the<strong>waves</strong>. We will describe this in detail in chapter 5. This approximation describeswave propagation well, but it is inadequate for wave generation. The most usefulapproximation for sources is the post-Newtonian approximation, where <strong>waves</strong>arise at a high order in corrections that carry general relativity away from itsNewtonian limit; we treat these in chapters 6 and 7.For now we concentrate our attention on lineariz<strong>ed</strong> theory. We followthe notation and conventions of Misner et al (1973) and Schutz (1985). In15


16 Elements of gravitational <strong>waves</strong>particular we choose units in which c = G = 1; Greek indices run from 0to 3; Latin indices run from 1 to 3; repeat<strong>ed</strong> indices are summ<strong>ed</strong>; commasin subscripts or superscripts denote partial derivatives; and semicolons denotecovariant derivatives. The metric has positive signature. These above twotextbooks and others referr<strong>ed</strong> to at the end of these chapters give more detailson the theory that we outline here. For an even simpler introduction, bas<strong>ed</strong> on ascalar analogy to general relativity, see [1].2.1 Mathematics of lineariz<strong>ed</strong> theoryConsider a perturb<strong>ed</strong> flat spacetime. Its metric tensor can be written asg αβ = η αβ + h αβ , |h αβ |≪1, α,β = 0,...,3 (2.1)where η αβ is the Minkowski metric (−1, 1, 1, 1) and h αβ is a very smallperturbation of the flat spacetime metric. Lineariz<strong>ed</strong> theory is an approximationto general relativity that is correct to first order in the size of this perturbation.Since the size of tensor components depends on coordinates, one must be carefulwith such a definition. What we require for lineariz<strong>ed</strong> theory to be valid is thatthere should exist a coordinate system in which equation (2.1) holds in a suitablylarge region of spacetime. Even though η αβ is not the true metric tensor, we arefree to define raising and lowering indices of the perturbation with η αβ ,asifitwere a tensor on flat spacetime. We writeh αβ := η αγ η βδ h γδ .This leads to the following equation for the inverse metric, correct to first order(all we want in lineariz<strong>ed</strong> theory):g αβ = η αβ − h αβ . (2.2)The mathematics is simpler if we define the trace-revers<strong>ed</strong> metricperturbation:¯h αβ := h αβ − 1 2 η αβh, (2.3)where h := η αβ h αβ . There is considerable coordinate fre<strong>ed</strong>om in the componentsh αβ , since we can wiggle and stretch the coordinate system with a comparableamplitude and change the components. This coordinate fre<strong>ed</strong>om is call<strong>ed</strong> gaugefre<strong>ed</strong>om, by analogy with electromagnetism. We use this fre<strong>ed</strong>om to enforce theLorentz (or Hilbert) gauge:¯h αβ ,β = 0. (2.4)In this gauge the Einstein field equations (neglecting the quadratic and higherterms in h αβ ) are just a set of decoupl<strong>ed</strong> linear wave equations:( )− ∂2∂t 2 +∇2 ¯h αβ =−16πT αβ . (2.5)


Using the TT gauge to understand gravitational <strong>waves</strong> 17To understand wave propagation we look for the easiest solution of the vacuumgravitational field equations:( )£ ¯h αβ ≡ − ∂2∂t 2 +∇2 ¯h αβ = 0. (2.6)Plane <strong>waves</strong> have the form:¯h αβ = e αβ exp(ik γ x γ ) (2.7)where the amplitude , polarization tensor e αβ and wavevector k γ are allconstants. (As usual one has to take the real part of this expression.)The Einstein equations imply that the wavevector is ‘light-like’, k γ k γ = 0,and the gauge condition implies that the amplitude and the wavevector areorthogonal: e αβ k β = 0.Lineariz<strong>ed</strong> theory describes a classical gravitational field whose quantumdescription would be a massless spin 2 field that propagates at the spe<strong>ed</strong> oflight. We expect from this that such a field will have only two independentdegrees of fre<strong>ed</strong>om (helicities in quantum language, polarizations in classicalterms). To show this classically we remember that h αβ is symmetric, so it hasten independent components, and that the Lorentz gauge applies four independentconditions to these, r<strong>ed</strong>ucing the fre<strong>ed</strong>om to six. However, the Lorentz gauge doesnot fully fix the coordinates. In fact if we perform another infinitesimal coordinatetransformation (x µ → x µ + ξ µ with ξ µ ,ν = O(h)) and impose £ξ µ = 0, weremain in Lorentz gauge. We can use this fre<strong>ed</strong>om to demand:e 0α = 0 ⇒ e ij k j = 0 (transverse wave), (2.8)e i i = 0 (traceless wave). (2.9)These conditions can only be appli<strong>ed</strong> outside a sphere surrounding the source.Together they put the metric into the transverse-traceless (TT) gauge. We willexplicitly construct this gauge in chapter 5.2.2 Using the TT gauge to understand gravitational <strong>waves</strong>The TT gauge leaves only two independent polarizations out of the original ten,and it ensures that h αβ = h αβ . In order to understand the polarization degrees offre<strong>ed</strong>om, let us take the wave to move in the z-direction, so that k z = ω, k 0 = ω,k x = 0, k y = 0; the TT gauge conditions in equations (2.8) and (2.9) lead toe 0α = e zα = 0 and e xx =−e yy . This leaves only two independent componentsof the polarization tensor, say e xx and e xy (which we denote by the symbols⊕, ⊗).A wave for which e xy = 0 (pure ⊕ polarization) produces a metric of theform:ds 2 =−dt 2 + (1 + h + ) dx 2 + (1 − h + ) dy 2 + dz 2 , (2.10)


18 Elements of gravitational <strong>waves</strong>’+’0.2h/2-0.2t’x’Figure 2.1. Illustration of two linear polarizations and the associat<strong>ed</strong> wave amplitude.where h + = e xx exp[−iω(t − z)]. Such a metric produces opposite effects onproper distance at the two transverse axes, contracting one while expanding theother.If e xx = 0 we have pure ⊗ polarization h × which can be obtain<strong>ed</strong> from theprevious case by a simple 45 ◦ rotation, as in figure 2.1. Since the wave equationand TT conditions are linear, a general wave will be a linear combination of thesetwo polarization tensors. A circular polarization basis would be:e R = 1 √2(e + + ie × ), e L = 1 √2(e + − ie × ), (2.11)where e + , e × are the two linear polarization tensors and e R and e Lare polarizations that rotate in the right-hand<strong>ed</strong> and left-hand<strong>ed</strong> directions,respectively. It is important to understand that, for circular polarization,the polarization pattern rotates around the central position, but test particlesthemselves rotate only in small circles relative to the central position.Now we compute the effects of a wave in the TT gauge on a particle at rest inthe flat background metric η αβ before the passage of the gravitational wave. Thegeodesic equationd 2 x µdτ 2 + dx α dx βƔµ αβdτ dτ = 0implies in this case:d 2 x idτ 2 =−Ɣi 00 =− 1 2 (2h i0,0 − h 00,i ) = 0, (2.12)so that the particle does not move. The TT gauge, to first order in h αβ , representsa coordinate system that is comoving with freely-falling particles. Becauseh 0α = 0, TT time is proper time on the clock of freely-falling particles at rest.Tidal forces show the action of the wave independently of the coordinates.Let us consider the equation of geodesic deviation, which governs the separationof two neighbouring freely-falling test particles A and B. If the particles are


Interaction of gravitational <strong>waves</strong> with detectors 19initially at rest, then as the wave passes it produces an oscillating curvature tensor,and the separation ξ of the two particles is:d 2 ξ idt 2 = Ri 0 j0ξ j . (2.13)To calculate the component R i 0 j0 of the Riemann tensor in equation (2.13), wecan use the metric in the TT gauge, because the Riemann tensor is gauge-invariantat linear order (see exercise (d) at the end of this chapter). Therefore, we canreplace R i 0 j0 by R i 0 j0 = 1 2 hTTi j,00 and write:d 2 ξ idt 2 = 1 2 hTTi j,00ξ j . (2.14)This equation, with an initial condition ξ j(0)= constant, describes the oscillationsof Bs location as measur<strong>ed</strong> in the proper reference frame of A. The validity ofequation (2.14) is the same as that of the geodesic deviation equation: geodesicshave to be close to one another, in a neighbourhood where the change in curvatureis small. In this approximation a gravitational wave is like an extra force, call<strong>ed</strong>a tidal force, perturbing the proper distance between two test particles. If thereare other forces on the particles, so that they are not free, then as long as thegravitational field is weak, one can just add the tidal forces to the other forces andwork as if the particle were in special relativity.2.3 Interaction of gravitational <strong>waves</strong> with detectorsWe have shown above that the TT gauge is a particular coordinate system inwhich the polarization tensor of a plane gravitational wave assumes a very simpleform. This gauge is comoving for freely-falling particles and so it is not thelocally Minkowskian coordinate system that would be us<strong>ed</strong> by an experimenterto analyse an experiment. In general relativity one must always be aware of howone’s coordinate system is defin<strong>ed</strong>.We shall analyse two typical situations:• the detector is small compar<strong>ed</strong> to the wavelength of the gravitational <strong>waves</strong>it is measuring; and• the detector is comparable to or larger than that wavelength.In the first case we can use the geodesic deviation equation above to representthe wave as a simple extra force on the equipment. Bars detectors can always beanalys<strong>ed</strong> in this way. Laser interferometers on the Earth can be treat<strong>ed</strong> this waytoo. In these cases a gravitational wave simply produces a force to be measur<strong>ed</strong>.There is no more to say from the relativity point of view. The rest of the detectionstory is the physics of the detectors. Sadly, this is not as simple as gravitationalwave physics!


20 Elements of gravitational <strong>waves</strong>In the second case, the geodesic deviation equation is not useful becausewe have to abandon the ‘local mathematics’ of geodesic deviation and return tothe ‘global mathematics’ of the TT gauge and metric components h TT αβ . Spacebas<strong>ed</strong>interferometers like LISA, accurate ranging to solar-system spacecraft andpulsar timing are all in this class. Together with ground interferometers, these arebeam detectors: they use light (or radio <strong>waves</strong>) to register the <strong>waves</strong>.To study these detectors, it is easiest to remain in the TT gauge and tocalculate the effect of the <strong>waves</strong> on the (coordinate) spe<strong>ed</strong> of light. Let usconsider, for example, the ⊕ metric from equation (2.10) and examine a nullgeodesic moving in the x-direction. The spe<strong>ed</strong> along this curve is:( dxdt) 2=11 + h +. (2.15)This is only a coordinate spe<strong>ed</strong>, not a contradiction to special relativity.To analyse the way in which detectors work, suppose one arm of aninterferometer lies along the x-direction and the wave, for simplicity, is movingin the z-direction with a ⊕ polarization of any waveform h + (t) along this axis (itis a plane wave, so its waveform does not depend on x). Then a photon emitt<strong>ed</strong> attime t from the origin reaches the other end, at a fix<strong>ed</strong> coordinate position x = L,at the coordinate time∫ L √t far = t + 1 + h+ (t(x)) dx, (2.16)0where the argument t(x) denotes the fact that one must know the time to reachposition x in order to calculate the wave field. This implicit equation can be solv<strong>ed</strong>in lineariz<strong>ed</strong> theory by using the fact that h + is small, so we can use the first-ordersolution of equation (2.15) to calculate h + (t) to sufficient accuracy.To do this we expand the square root in powers of h + , and consider as azero-order solution a photon travelling at the spe<strong>ed</strong> of light in the x-direction of aflat spacetime. We can set t(x) = t + x. The result is:t out = t + L + 1 2∫ L0h + (t + x) dx. (2.17)In an interferometer, the light is reflect<strong>ed</strong> back, so the return trip takes[ ∫ L∫ L]t return = t + L + 1 2h + (t + x) dx + h + (t + x + L) dx . (2.18)0What one monitors is changes in the time taken by a return trip as a function oftime at the origin. If there were no gravitational <strong>waves</strong> t return would be constantbecause L is fix<strong>ed</strong>, so changes indicate a gravitational wave.The rate of variation of the return time as a function of the start time t isdt return= 1 + 1 dt 2 [h +(t + 2L) − h + (t)]. (2.19)0


Analysis of beam detectors 21This depends only on the wave amplitude when the beam leaves and when itreturns.Let us consider now a more realistic geometry than the previous one, andin particular suppose that the wave travels at an angle θ to the z-axis in the x–zplane. If we r<strong>ed</strong>o this calculation, allowing the phase of the wave to depend onx in an appropriate way, and taking into account the fact that h TT+ xx is r<strong>ed</strong>uc<strong>ed</strong> ifthe wave is not moving in a direction perpendicular to x, we find (see exercise (a)at the end of this chapter for the details of the calculation)dt returndt= 1 {(1 − sin θ)hxx + (t + 2L) − (1 + sin θ)hxx +2 (t)+ 2 sin θh+ xx [t + L(1 − sin θ)]}. (2.20)This three-term relation is the starting point for analysing the response of all beamdetectors. This is directly what happens in radar ranging or in transpondingto spacecraft, where a beam in only one direction is us<strong>ed</strong>. In long-baselineinterferometry, one must analyse the second beam as well. We shall discuss thesecases in turn.2.4 Analysis of beam detectors2.4.1 Ranging to spacecraftBoth NASA and ESA perform experiments in which they monitor the return timeof communication signals with interplanetary spacecraft for the characteristiceffect of gravitational <strong>waves</strong>. For missions to Jupiter and Saturn, the return timesare of the order 2–4 × 10 3 s. Any gravitational wave event shorter than thiswill leave an imprint on the delay time three times: once when the wave passesthe Earth-bas<strong>ed</strong> transmitter, once when it passes the spacecraft, and once whenit passes the Earth-bas<strong>ed</strong> receiver. Searches use a form of pattern matching tolook for this characteristic imprint. There are two dominant sources of noise:propagation-time irregularities caus<strong>ed</strong> by fluctuations in the solar wind plasma,and timing noise in the clocks us<strong>ed</strong> to measure the signals. The plasma delaysdepend on the radio-wave frequency, so by using two transmission frequenciesone can model and subtract the plasma noise. Then if one uses the most stableatomic clocks, it is possible to achieve sensitivities for h of the order 10 −13 .Inthe future, using higher radio frequencies, such experiments may reach 10 −15 .No positive detections have yet been made, but the chances are not zero. Forexample, if a small black hole fell into a massive black hole in the centre ofthe Galaxy, it would produce a signal with a frequency of about 10 mHz and anamplitude significantly bigger than 10 −15 . Rare as this might be, it would be adramatic event to observe.


22 Elements of gravitational <strong>waves</strong>2.4.2 Pulsar timingMany pulsars, in particular old millisecond pulsars, are extraordinarily regularclocks, whose random timing irregularities are too small for even the bestatomic clocks to measure. Other pulsars have weak but observable irregularities.Measurements of or even upper limits on any of these timing irregularities forsingle pulsars can be us<strong>ed</strong> to set upper limits on any background gravitationalwave field with periods comparable to or shorter than the observing time. Here thethree-term formula is replac<strong>ed</strong> by a simpler two-term expression (see exercise (b)at the end of this chapter), because we only have a one-way transmission fromthe pulsar to Earth. Moreover, the transit time of a signal to Earth from the pulsarmay be thousands of years, so we cannot look for correlations between the twoterms in a given signal. Instead, the delay time is a combination of the effectsof uncorrelat<strong>ed</strong> <strong>waves</strong> at the pulsar when the signal was emitt<strong>ed</strong> and at the Earthwhen it is receiv<strong>ed</strong>.If one simultaneously observes two or more pulsars, the Earth-bas<strong>ed</strong> part ofthe delay is correlat<strong>ed</strong> between them, and this offers a means of actually detectinglong-period gravitational <strong>waves</strong>. Observations require a timescale of several yearsin order to achieve the long-period stability of pulse arrival times, so this methodis suit<strong>ed</strong> to looking for strong gravitational <strong>waves</strong> with periods of several years.2.4.3 InterferometryAn interferometer essentially measures changes in the difference in the returntimes along two different arms. It does this by looking for changes in theinterference pattern form<strong>ed</strong> when the returning light beams are superimpos<strong>ed</strong>on one another. The response of each arm will follow the three-term formulain equation (2.20), but with a different value of θ for each arm, depending in acomplicat<strong>ed</strong> way on the orientation of the arms relative to the direction of traveland the polarization of the wave. Ground-bas<strong>ed</strong> interferometers are small enoughto use the small-L formulae we deriv<strong>ed</strong> earlier. However, LISA, the space-bas<strong>ed</strong>interferometer that is describ<strong>ed</strong> by Bender in this book, is larger than a wavelengthof gravitational <strong>waves</strong> for frequencies above 10 mHz, so a detail<strong>ed</strong> analysis of itssensitivity requires the full three-term formula.2.5 Exercises for chapter 2Suggest<strong>ed</strong> solutions for these exercises are at the end of chapter 7.(a) 1. Derive the full three-term return equation, reproduc<strong>ed</strong> here:dt returndt= 1 2{(1 − sin θ)hxx + (t + 2L) − (1 + sin θ)hxx + (t)+ 2 sin θh+ xx [t + L(1 − sin θ)]}. (2.21)


Exercises for chapter 2 232. Show that, in the limit where L is small compar<strong>ed</strong> to the wavelength ofthe gravitational wave, the derivative of the return time is the derivativeof the excess proper distance δL = Lh+ xx(t)cos2 θ for small L. Makesure you know how to interpret the factor of cos 2 θ.3. Examine the limit of the three-term formula when the gravitational waveis travelling along the x-axis too (θ =± π 2): what happens to light goingparallel to a gravitational wave?(b) Derive the two-term formula governing the delays induc<strong>ed</strong> by gravitational<strong>waves</strong> on a signal transmitt<strong>ed</strong> only one-way, for example from a pulsar toEarth.(c) A frequently ask<strong>ed</strong> question is: if gravitational <strong>waves</strong> alter the spe<strong>ed</strong> of light,as we seem to have us<strong>ed</strong> here, and if they move the ends of an interferometercloser and further apart, might these effects not cancel, so that there wouldbe no measurable effects on light? Answer this question. You may want toexamine the calculation above: did we make use of the changing distancebetween the ends, and why or why not?(d) Show that the Riemann tensor is gauge-invariant in lineariz<strong>ed</strong> theory.


Chapter 3<strong>Gravitational</strong>-wave detectors<strong>Gravitational</strong> radiation is a central pr<strong>ed</strong>iction of general relativity and its detectionis a key test of the integrity of the theoretical structure of Einstein’s work.However, in the long run, its importance as a tool for observational astronomy islikely to be even more important. We have excellent observational evidence fromthe Hulse–Taylor binary pulsar system (describ<strong>ed</strong> in chapter 4) that the pr<strong>ed</strong>ictionsof general relativity concerning gravitational radiation are quantitatively correct.However, we have incomplete information from astronomy today about the likelysources of detectable radiation.The gravitational wave spectrum is completely unexplor<strong>ed</strong>, and whenever anew electromagnetic waveband has been open<strong>ed</strong> to astronomy, astronomers hav<strong>ed</strong>iscover<strong>ed</strong> completely unexpect<strong>ed</strong> phenomena. This seems to me just as likelyto happen again with gravitational <strong>waves</strong>, especially because gravitational <strong>waves</strong>carry some kinds of information that electromagnetic radiation cannot convey.<strong>Gravitational</strong> <strong>waves</strong> are generat<strong>ed</strong> by bulk motions of masses, and they encodethe mass distributions and spe<strong>ed</strong>s. They are coherent and their low frequenciesreflect the dynamical timescales of their sources.In contrast, electromagnetic <strong>waves</strong> come from individual electrons executingcomplex and partly random motions inside their sources. They are incoherent, andindividual photons must be interpret<strong>ed</strong> as samples of the large statistical ensembleof photons being emitt<strong>ed</strong>. Their frequencies are determin<strong>ed</strong> by microphysics onlength scales much smaller than the structure of the astronomical system emittingthem. From electromagnetic observations we can make inferences about thisstructure only through careful modelling of the source. <strong>Gravitational</strong> <strong>waves</strong>, bycontrast, carry information whose connection to the source structure and motionis fairly direct.A good example is that of massive black holes in galactic nuclei. Fromobservations that span the electromagnetic spectrum from radio <strong>waves</strong> to x-rays, astrophysicists have inferr<strong>ed</strong> that black holes of masses up to 10 9 M ⊙are responsible for quasar emissions and control the jets that power the giantradio emission regions. The evidence for the black hole is very strong but24


<strong>Gravitational</strong>-wave detectors 25indirect: no other known object can contain so much mass in such a smallvolume. <strong>Gravitational</strong> wave observations will tell us about the dynamics of theholes themselves, providing unique signatures from which they can be identifi<strong>ed</strong>,measuring their masses and spins directly from their vibrational frequencies.The interplay of electromagnetic and gravitational observations will enrich manybranches of astronomy.The history of gravitational-wave detection start<strong>ed</strong> in the 1960s with J Weberat the University of Maryland. He built the first bar detector: it was a massivecylinder of aluminium (∼2 × 10 3 kg) operating at room temperature (300 K)with a resonant frequency of about 1600 Hz. This early prototype had a modestsensitivity, around 10 −13 or 10 −14 .Despite this poor sensitivity, in the late 1960s Weber announc<strong>ed</strong> the detectionof a population of coincident events between two similar bars at a rate farhigher than expect<strong>ed</strong> from instrumental noise. This news stimulat<strong>ed</strong> a numberof other groups (at Glasgow, Munich, Paris, Rome, Bell Laboratories, Stanford,Rochester, LSU, MIT, Beijing, Tokyo) to build and develop bar detectors tocheck Weber’s results. Unfortunately for Weber and for the idea that gravitational<strong>waves</strong> were easy to detect, none of these other detectors found anything, even attimes when Weber continu<strong>ed</strong> to find coincidences. Weber’s observations remainunexplain<strong>ed</strong> even today. However, the failure to confirm Weber was in a realsense a confirmation of general relativity, because theoretical calculations hadnever pr<strong>ed</strong>ict<strong>ed</strong> that reasonable signals would be strong enough to be seen byWeber’s bars.Weber’s announcements have had a mix<strong>ed</strong> effect on gravitational-waveresearch. On the one hand, they have creat<strong>ed</strong> a cloud under which the fieldhas labour<strong>ed</strong> hard to re-establish its respectability in the eyes of many physicists.Even today the legacy of this is an extreme cautiousness among the major projects,a conservatism that will ensure that the next claim of a detection will be ironclad.On the other hand, the stimulus that Weber gave to other groups to build detectorshas directly l<strong>ed</strong> to the present advanc<strong>ed</strong> state of detector development.From 1980 to 1994 groups develop<strong>ed</strong> detectors in two different directions:• Cryogenic bar detectors, develop<strong>ed</strong> primarily at Rome/Frascati, Stanford,LSU and Perth (Australia). The best of these detectors reach below 10 −19 .They are the only detectors operating continuously today and they haveperform<strong>ed</strong> a number of joint coincidence searches, leading to upper limitsbut no detections.• Interferometers, develop<strong>ed</strong> at MIT, Garching (where the Munich groupmov<strong>ed</strong>), Glasgow, Caltech and Tokyo. The typical sensitivity of theseprototypes was 10 −18 . The first long coincidence observation withinterferometers was the Glasgow/Garching 100 hr experiment in 1989 [2].In fact, interferometers had apparently been consider<strong>ed</strong> by Weber, but at thattime the technology was not good enough for this kind of detector. Only 10–15 years later, technology had progress<strong>ed</strong>. Lasers, mirror coating and polishing


26 <strong>Gravitational</strong>-wave detectorstechniques and materials science had advanc<strong>ed</strong> far enough to allow the firstpractical interferometers, and it was clear that further progress would continueunabat<strong>ed</strong>. Soon afterwards several major collaborations were form<strong>ed</strong> to buildlarge-scale interferometric detectors:• LIGO: Caltech and MIT (NSF) LIGO;• VIRGO: France (CNRS) and Italy (INFN)• GEO600: Germany (Max Planck) and UK (PPARC).Later, other collaborations were form<strong>ed</strong> in Australia (AIGO) and Japan (TAMAand JGWO). At present there is still considerable effort in building successors toWeber’s original resonant-mass detector: ultra-cryogenic bars are in operation inFrascati and Padua, and they are expect<strong>ed</strong> to reach below 10 −20 . Further, thereare proposals for a new generation of spherical or icosah<strong>ed</strong>ral solid-mass detectorsfrom the USA (LSU), Brazil, the Netherlands and Italy. Arrays of smaller barshave been propos<strong>ed</strong> for observing the highest frequencies, where neutron starnormal modes lie.However, the real goal for the near future is to break through the 10 −21 level,which is where theory pr<strong>ed</strong>icts that it is not unreasonable to expect gravitational<strong>waves</strong> of the order of once per year (see the discussion in chapter 4 later). Thefirst detectors to reach this level will be the large-scale interferometers that arenow under construction. They have very long arms: LIGO, Hanford (WA) andLivingstone (LA), 4 km; VIRGO: Pisa, 3 km; GEO600: Hannover, 600 m;TAMA300: Tokyo, 300 m.The most spectacular detector in the near future is the space-bas<strong>ed</strong> detectorLISA, which has been adopt<strong>ed</strong> by ESA (European Space Agency) as aCornerstone mission for the twenty-first century. The project is now gaining aconsiderable amount of momentum in the USA, and a collaboration between ESAand NASA seems likely. This mission could be launch<strong>ed</strong> around 2010.3.1 <strong>Gravitational</strong>-wave observablesWe have describ<strong>ed</strong> earlier how different gravitational-wave observables are fromelectromagnetic observables. Here are the things that we want to measure whenwe detect gravitational <strong>waves</strong>:• h + (t), h × (t), phase(t): the amplitude and polarization of the wave, andthe phase of polarization, as functions of time. These contain most of theinformation about gravitational <strong>waves</strong>.• θ, φ: the direction on the sky of the source (except for observations of astochastic background).From this it is clear that gravitational-wave detection is not the same aselectromagnetic-radiation detection. In electromagnetic astronomy one almostalways rectifies the electromagnetic wave, while we can follow the oscillations of


<strong>Gravitational</strong>-wave observables 27the gravitational wave. Essentially in electromagnetism one detects the power inthe radiation, while for gravitational radiation, as we have said before, one detectsthe wave coherently.Let us consider now what we can infer from a detection. If the gravitationalwave has a short duration, of the order of the sampling time of the signal stream,then each detector will usually give just a single number, which is the amplitudeof the wave project<strong>ed</strong> on the detector (a projection of the two polarizations h +and h × ). If the wave lasts more than one sampling time, then this information isa function of time.If the signal lasts for a sufficiently long time, then both the amplitude andthe phase of the wave can be affect<strong>ed</strong> by the motion of the detector, which movesand turns with the motion of the Earth. This produces an amplitude and phasemodulation which is not intrinsic to the signal. If the signal’s intrinsic form isunderstood, then this modulation can be us<strong>ed</strong> to determine the location of thesource. We distinguish three distinct kinds of signals, from the point of view ofobservations.Bursts have a duration so short that modulation due to detector motion is notobservable. During the detection, the detector is effectively stationary. In thiscase we ne<strong>ed</strong> at least three, and preferably four, interferometers to triangulate thepositions of bursts on the sky and to find the two polarizations h + and h × . (Se<strong>ed</strong>iscussions in Schutz 1989.) A network of detectors is essential to extract all theinformation in this case.Continuous <strong>waves</strong> by definition last long enough for the motion of th<strong>ed</strong>etector to induce amplitude and phase modulation. In this case, assuming asimple model for the intrinsic signal, we can use the information imprint<strong>ed</strong> onthe signal (the amplitude modulation and phase modulation) to infer the positionand polarization amplitude of the source on the sky. A single detector, effectively,performs aperture synthesis, finding the position of the source and the amplitudeof the wave entirely by itself. However, in order to be sure that the signal is not anartefact, it will be important that the signal is seen by a second or third detector.Stochastic backgrounds can be detect<strong>ed</strong> just like noise in a single detector.If the detector noise is well understood, this excess noise may be detect<strong>ed</strong> asa stochastic background. This is closely analogous to the way the originalmicrowave background detection was discover<strong>ed</strong>.A more reliable method for detecting stochastic radiation is the crosscorrelationbetween two detectors, which experience the same cosmological noisebut have a different intrinsic noise. Coherent cross-correlation between twodetectors eliminates much detector noise and works best when detectors are closerthan a wavelength.In general, detection of gravitational <strong>waves</strong> requires joint observing by anetwork of detectors, both to increase the confidence of the detection and toprovide accurate information on other physical observables (direction, amplitudeand so on). Networks can be assembl<strong>ed</strong> from interferometers, bars, or both.


28 <strong>Gravitational</strong>-wave detectors3.2 The physics of interferometersInterferometric gravitational-wave detectors are the most sensitive instruments,and among the most complex, that have ever been construct<strong>ed</strong>. They areremarkable for the range of physics that is important for their construction.Interferometer groups work at the forefront of the development in lasers, mirrorpolishing and coating, quantum measurement, materials science, mechanicalisolation, optical system design and thermal science. In this section we shallonly be able to take a fairly superficial look at one of the most fascinatinginstrumentation stories of our age. A good introduction to interferometer designis Saulson (1994).Interferometers use laser light to compare the lengths of two perpendiculararms. The simplest design, originat<strong>ed</strong> by Michelson for his famous experimenton the velocity of light, uses light that passes up and down each arm once, asin the first panel in figure 3.1. Imagine such an instrument with identical armsdefin<strong>ed</strong> by mirrors that hang from supports, so they are free to move horizontallyin response to a gravitational wave. If there is no wave, the arms have the samelength, and the light from one arm returns exactly in phase with that from theother. When the wave arrives, the two arms typically respond differently. Thearms are no longer the same length, and so the light that arrives back at the centrefrom one arm will no longer be in phase with that arriving back from the otherarm. This will produce a shift in the interference fringes between the two beams.This is the principle of detection.Real detectors are design<strong>ed</strong> to store the light in each arm for longer thanjust one reflection (see figure 3.1(b)). It is optimum to store the light for halfof the period of the gravitational wave, so that on each reflection the light gainsan add<strong>ed</strong> phase shift. Michelson-type delay-line interferometers store the lightby arranging multiple reflections. Fabry–Perot interferometers store the light incavities in each arm, allowing only a small fraction to escape for the interferencemeasurement (figure 3.1(e)).An advantage of interferometers as detectors is that the gravitational-waveinduc<strong>ed</strong>phase shift of the light can be made larger simply by making the armlength larger, since gravitational <strong>waves</strong> act by tidal forces. A detector with an armlength l = 4 km responds to a gravitational wave with an amplitude of 10 −21 withδl gw ∼ 1 2 hl ∼ 2 × 10−18 m (3.1)where δl gw is the change in the length of one arm. If the orientation of theinterferometer is optimum, then the other arm will change by the same amountin the opposite direction, so that the interference fringe will shift by twice thislength.If the light path is fold<strong>ed</strong> or resonat<strong>ed</strong>, as in figure 3.1(b) and (d), then theeffective number of bounces can be trad<strong>ed</strong> off against overall length to achievea given desir<strong>ed</strong> total path length, or storage time. Shorter interferometers withmany bounces have a disadvantage, however: even though they can achieve the


The physics of interferometers 29 Figure 3.1. Five steps to a gravitational-wave interferometer. (a) The simple Michelson.Notice that there are two return beams: one goes toward the photodetector and the othertoward the laser. (b) Delay line: a Michelson with multiple bounces in each arm to enhancethe signal. (c) Power recycling. The extra mirror recycles the light that goes towards thelaser, which would otherwise be wast<strong>ed</strong>. (d) Signal recycling. The mirror in front of thephotodetector recycles only the signal sidebands, provid<strong>ed</strong> that in the absence of a signalno light goes to the photodetector. (e) Fabry–Perot interferometer. The delay lines areconvert<strong>ed</strong> to cavities with partially silver<strong>ed</strong> interior mirrors.same response as a longer interferometer, the extra bounces introduce noise fromthe mirrors, as discuss<strong>ed</strong> below. There is, therefore, a big advantage to long-arminterferometers.


30 <strong>Gravitational</strong>-wave detectorsThere are three main sources of noise in interferometers: thermal, shot andvibrational. To understand the way they are controll<strong>ed</strong>, it is important to thinkin frequency space. Observations with ground-bas<strong>ed</strong> detectors will be made in arange from perhaps 10 Hz up to 10 kHz, and initial detectors will have a muchsmaller observing bandwidth within this. Disturbances by noise that occur atfrequencies outside the observation band can simply be filter<strong>ed</strong> out. The goal ofnoise control is to r<strong>ed</strong>uce disturbances in the observation band.• Thermal noise. Interferometers work at room temperature, and vibrations ofthe mirrors and of the suspending pendulum can mask gravitational <strong>waves</strong>.To control this noise, scientists take advantage of the fact that thermal noisehas its maximum amplitude at the frequency of the vibrational mode, andif the resonance of the mode is narrow (a high quality factor Q) then theamplitude at other frequencies is small. Therefore, pendulum suspensionsare design<strong>ed</strong> with the pendulum frequency at about 1 Hz, well belowthe observing window, and mirror masses are design<strong>ed</strong> to have principalvibration modes above 1 kHz, well above the optimum observing frequencyfor initial interferometers. These systems are construct<strong>ed</strong> with high values ofQ (10 6 or more) to r<strong>ed</strong>uce the noise in the observing band. Even so, thermalnoise is typically a dominant noise below 100 or 200 Hz.• Shot noise. This is the principal limitation to sensitivity at higherfrequencies, above 200–300 Hz. It arises from the quantization of photons.When photons form interference fringes, they arrive at random timesand make random fluctuations in the light intensity that can look like agravitational wave signal; the more photons one uses, the smoother will bethe interference fringe. We can easily calculate this intrinsic noise. If N isthe number of photons emitt<strong>ed</strong> by the laser during our measurement, thenas a random process the fluctuation number δN is proportional to the squareroot of N. If we are using light with a wavelength λ (for example infrar<strong>ed</strong>light with λ ∼ 1 µm) one can expect to measure lengths to an accuracy ofλδl shot ∼2π √ N .To measure a gravitational wave at a frequency f , one has to make at least 2 fmeasurements per second, so one can accumulate photons for a time 1/2 f .If P is the light power, one hasN =Phc 1λ ·2 f.It is easy to work out from this that, for δl shot to be equal to δl gw inequation (3.1), one ne<strong>ed</strong>s light power of about 600 kW. No continuous lasercould provide this much light to an interferometer.The key to reaching such power levels inside the arms of a detector isa technique call<strong>ed</strong> power recycling (see Saulson 1994) first propos<strong>ed</strong> by


The physics of interferometers 31Drever and independently by Schilling. Normally, interferometers workon a ‘dark fringe’, that is they are arrang<strong>ed</strong> so that the light reachingthe photodetector is zero if there is no gravitational wave. Then, asshown in figure 3.1(a), the whole of the input light must emerge from theinterferometer travelling towards the laser. If one places another mirror,correctly position<strong>ed</strong>, between the laser and the beam splitter (figure 3.1(c)),it will reflect this wast<strong>ed</strong> light back into the interferometer in such a way thatit adds coherently in phase with light emerging from the laser. In this way,light can be recycl<strong>ed</strong> and the requir<strong>ed</strong> power levels in the arms achiev<strong>ed</strong>.Of course, there will be a maximum recycling gain, which is set by mirrorlosses. Light power builds up until the laser merely re-supplies the losses atthe mirrors, due to scattering and absorption. The maximum power gain isP =11 − R 2where 1 − R 2 is the total loss summ<strong>ed</strong> over all the optical surfaces. For thevery high-quality mirrors us<strong>ed</strong> in these projects, 1−R 2 ∼ 10 −5 . This r<strong>ed</strong>ucesthe power requirement for the laser by the same factor, down to about 6 W.This is attainable with modern laser technology.• Ground vibration and mechanical vibrations are another source of noisethat must be screen<strong>ed</strong> out. Typical seismic vibration spectra fall sharplywith frequency, so this is a problem primarily below 100 Hz. Pendulumsuspensions are excellent mechanical filters above the pendulum frequency:it is a familiar elementary-physics demonstration that one can wiggle thesuspension point of a pendulum vigorously at a high frequency and thependulum itself remains undisturb<strong>ed</strong>. Suspension designs typically involvemultiple pendula, each with a frequency around 1 Hz. These provide veryfat roll-off of the noise above 1 Hz. Interferometer spectra normally showa steep low-frequency noise ‘wall’: this is the expect<strong>ed</strong> vibrational noiseamplitude.In addition, there are noise sources that are not dominant in the presentinterferometers but will become important as sensitivity increases.• Quantum effects: uncertainty principle noise. Shot noise is a quantum noise,but in addition there are other effects similar to those that bar detectorsface, as describ<strong>ed</strong> below: zero-point vibrations of suspensions and mirrorsurfaces, and back-action of light pressure fluctuations on the mirrors. Theseare small compar<strong>ed</strong> to present operating limits of detectors, but they maybecome important in five years or so. Practical schemes to r<strong>ed</strong>uce this noisehave already been demonstrat<strong>ed</strong> in principle, but they ne<strong>ed</strong> to be improv<strong>ed</strong>considerably. This is the subject of considerable theoretical work at themoment.• Gravity gradient noise. <strong>Gravitational</strong>-wave detectors respond to any changesin the gradients (tidal forces) of the local gravitational field, not just


32 <strong>Gravitational</strong>-wave detectorsthose carri<strong>ed</strong> by <strong>waves</strong>. The environment always contains changes in theNewtonian fields of nearby objects. Besides obvious ones, like people,there are changes caus<strong>ed</strong> by density <strong>waves</strong> in ground vibrations, atmosphericpressure changes, and many other disturbances. Below about 1 Hz,these gravity gradient changes will be stronger than <strong>waves</strong> expect<strong>ed</strong> fromastronomical objects, and they make it impossible to do observing at lowfrequencies from Earth. This is the reason that scientists have propos<strong>ed</strong> theLISA mission, discuss<strong>ed</strong> later. Above 1 Hz, this noise does not affect thesensitivity of present detectors, but in ten years this could become a limitingfactor.Besides these noise sources, which are pr<strong>ed</strong>ictable and therefore can becontroll<strong>ed</strong> by detector design, it is possible that there will be unexpect<strong>ed</strong> orunpr<strong>ed</strong>ict<strong>ed</strong> noise sources. Interferometers will be instrument<strong>ed</strong> with many kindsof environmental monitors, but there may occasionally be noise that is impossibleto identify. For this reason, short bursts of gravitational radiation must beidentifi<strong>ed</strong> at two or more separat<strong>ed</strong> facilities. Even if detector noise is not atall understood, it is relatively easy to estimate from the observ<strong>ed</strong> noise profile ofthe individual detectors what the chances are of a coincident noise event betweentwo detectors.3.2.1 New interferometers and their capabilitiesInterferometers work over a broad bandwidth and they do not have any naturalresonance in their observing band. They are ideal for detecting bursts, since onecan perform pattern-matching over the whole bandwidth and detect such signalsoptimally. They are also ideal for searching for unknown continuous signals, suchas surveying the sky for neutron stars. And in observations of stochastic signalsby cross-correlating two detectors, they can give information about the spectrumof the signal.If an interferometer wants to study a signal with a known frequency, suchas known pulsars, then there is another optical technique available to enhance itssensitivity in a narrow bandwidth, at the expense of sensitivity outside that band.This is call<strong>ed</strong> signal recycling [3]. In this technique, a further mirror is plac<strong>ed</strong> infront of the photodetector, where the signal emerges from the interferometer (seefigure 3.1(d)). If the mirror is chosen correctly, it will build up the signal, but onlyin a certain bandwidth. This modifies the shot noise in the detector, but not othernoise sources. Therefore, it can improve sensitivity only at the higher frequencieswhere shot noise is the limiting factor.Four major interferometer projects are now under construction, and theycould begin acquiring good data in the period between 2000–2003. They will alloperate initially with a sensitivity approaching 10 −21 over a bandwidth between50–1000 Hz. Early detections are by no means certain, but recent work hasmade prospects look better for an early detection than when these detectors werefund<strong>ed</strong>.


The physics of interferometers 33Figure 3.2. TAMA300 sensitivity as a function of frequency. The vertical axis is the1σ noise level, measur<strong>ed</strong> in strain per root Hz. To get a limit on the gravitational waveamplitude h, one must multiply the height of the curve by the square root of the bandwidthof the signal. This takes into account the fact that the noise power at different frequenciesis independent, so the power is proportional to bandwidth. The noise amplitude is thereforeproportional to the square root of the bandwidth.TAMA300 [4] (Japan) is locat<strong>ed</strong> in Tokyo, and its arm length is 300 m.It began taking data without power recycling in 1999, but its sensitivity is notyet near 10 −21 . Following improvements, especially power recycling, it shouldget to within a factor of ten of this goal. However, it is not plann<strong>ed</strong> as anobserving instrument: it is a prototype for a kilometre-scale interferometer inJapan, currently call<strong>ed</strong> JGWO. By 2005 this may be operating, possibly withcryogenically cool<strong>ed</strong> mirrors.GEO600 [5] (Germany and Britain) is locat<strong>ed</strong> near Hannover (Germany).Its arm length is 600 m and the target date for first good data is now the end of<strong>2001</strong>. Unlike TAMA, GEO600 is design<strong>ed</strong> as a leading-<strong>ed</strong>ge-technologydetector,where high-performance suspensions and optical tricks like signal recycling canbe develop<strong>ed</strong> and appli<strong>ed</strong>. Although it has a short baseline, it will have a similarsensitivity to the larger LIGO and VIRGO detectors at first. At a later stage, LIGOand VIRGO will incorporate the advanc<strong>ed</strong> methods develop<strong>ed</strong> in GEO, and at thatpoint they will advance in sensitivity, leaving GEO behind.As we can see from figure 3.3 the sensitivity of GEO600 depends on itsbandwidth, which in its turn depends on the signal recycling factor. GEO600 canchange its observing bandwidth in response to observing goals. By choosing lowor high reflectivity for the signal recycling mirror, scientists can make GEO600wide-band or narrow-band, respectively. The centre frequency of the observingband (in the right-hand panel of figure 3.3 it is ∼600 Hz) can be tun<strong>ed</strong> to anydesir<strong>ed</strong> frequency by shifting the position of the signal recycling mirror, thus


34 <strong>Gravitational</strong>-wave detectorsNoise amplitude spectral density10 −2110 −2210 −23SeismicNoise10 −20 Frequency (Hz)TotalThermal NoiseUncertainty PrincipleShot Noise(broad band)10 2 10 3Noise amplitude spectral density10 −2110 −2210 −23SeismicNoise10 −20 Frequency (Hz)TotalShot Noise(narrow bandat 600 Hz)Thermal NoiseUncertainty Principle10 2 10 3Figure 3.3. GEO600 noise curves. As for the TAMA curve, these are calibrat<strong>ed</strong> in strainper root Hz. The figure on the left-hand side shows GEO’s wideband configuration; thaton the right-hand side shows a possible narrowband operating mode.changing the resonance frequency of the signal recycling cavity. This featurecould be useful when interferometers work with bars or when performing widebandsurveys.LIGO [6] (USA) is building two detectors of arm length 4 km. One islocat<strong>ed</strong> in Hanford WA and the other in Livingstone LA. The target date forobserving is mid-2002. The two detectors are plac<strong>ed</strong> so that their antenna patternsoverlap as much as possible and yet they are far enough apart that there will be ameasurable time delay in most coincident bursts of gravitational radiation. Thisdelay will give some directional information. The Hanford detector also containsa half-length interferometer to assist in coincidence searches. The two LIGOdetectors are the best plac<strong>ed</strong> for doing cross-correlation for a random backgroundof gravitational <strong>waves</strong>. LIGO’s expect<strong>ed</strong> initial noise curve is shown in figure 3.4.These detectors have been construct<strong>ed</strong> to have a long lifetime. With such longarms they can benefit from upgrades in laser power and mirror quality. LIGO hasdefin<strong>ed</strong> an upgrade goal call<strong>ed</strong> LIGO II, which it hopes to reach by 2007, whichwill observe at 10 −22 or better over a bandwidth from 10 Hz up to 1 kHz.VIRGO [7] (Italy and France) is building a 3 km detector near Pisa. Itstarget date for good data is 2003. Its expect<strong>ed</strong> initial noise curve is shown infigure 3.4. Like LIGO, it can eventually be push<strong>ed</strong> to much higher sensitivitieswith more powerful lasers and other optical enhancements. VIRGO specializesin sophisticat<strong>ed</strong> suspensions, and the control of vibrational noise. Its goal is toobserve at the lowest possible frequencies from the ground, at least partly to beable to examine as many pulsars and other neutron stars as possible.3.3 The physics of resonant mass detectorsThe principle of operation of bar detectors is to use the gravitational tidal forceof the wave to stretch a massive cylinder along its axis, and then to measure theelastic vibrations of the cylinder.


The physics of resonant mass detectors 3510 −1410 −16h (Hz −1/2 )10 −1810 −2010 −2210 −2410 1 10 2 10 3 10 4frequency (Hz)Figure 3.4. Noise curves of the initial LIGO (left) and VIRGO (right) detectors. TheVIRGO curve is in strain per root Hz, as the GEO curves earlier. The LIGO curve iscalibrat<strong>ed</strong> in metres per root Hz, so to convert to a limit on h one multiplies by the squareroot of the bandwidth and divides by the length of the detector arm, 4000 m.Let us suppose we have a typical bar with length L ∼ 1 m. (In the future,spheres may go up to 3 m.) Depending on the length of the bar and its material,the resonant frequency will be f ∼ 500 Hz to 1.5 kHz and mass M ∼ 1000 kg.A short burst gravitational wave h will make the bar vibrate with an amplitudeδl gw ∼ hl ∼ 10 −21 m.Unlike the interferometers, whose response is simply given by this equation,the bars respond in a complicat<strong>ed</strong> way depending on all their internal forces.However, if the duration of the wave is short, the amplitude will be of the sameorder as that given here. If the wave has long duration and is near the bars resonantfrequency, then the signal can build up to much larger amplitudes. Normally, bardetector searches have been target<strong>ed</strong> at short-duration signals.The main sources of noise that compete with this very small amplitude are:• Thermal noise. This is the most serious source of noise. Interferometers canlive with room-temperature thermal noise because their larger size makestheir response to a gravitational wave larger, and because they observe atfrequencies far from the resonant frequency, where the noise amplitude islargest. However, bars observe at the resonant frequency and have a veryshort length, so they must r<strong>ed</strong>uce thermal noise by going to low temperatures.The best ultra-cryogenic bars today operate at about T = 100 mK, wherethe rms amplitude of vibration is found by setting the kinetic energy of thenormal mode, M(δ ˙l) 2 /2, equal to kT/2, the equipartition thermal energy ofa single degree of fre<strong>ed</strong>om. This gives then〈δl 2 〉 1 2th=( kT4π 2 Mf 2 )12∼ 6 × 10 −18 m.


36 <strong>Gravitational</strong>-wave detectorsThis is far larger than the gravitational wave amplitude. In order to detectgravitational <strong>waves</strong> against this noise, bars are construct<strong>ed</strong> to have a veryhigh Q, of order 10 6 or better.The reason that bars ne<strong>ed</strong> a high Q is different from the reason thatinterferometers also strive for high-Q systems. To see how Q helps bars,we recall that Q is defin<strong>ed</strong> as Q = f · τ where f is the resonant frequencyof the mode and τ is the decay time of the oscillations. If Q is large, thenthe decay time is long. If the decay time is long, then the amplitude ofoscillation changes very slowly in thermal equilibrium. Essentially, the bar’smode of vibration changes its amplitude by a random walk with very smallsteps, taking time Q/ f ∼ 1000 s to change by the full amount. On the otherhand, a gravitational wave burst will cause an amplitude change in time ofthe order 1 ms, during which the thermal noise will have random walk<strong>ed</strong> toan expect<strong>ed</strong> amplitude change that is Q 1 2 = ( 1000 1ms s ) 1 2 times smaller. In thiscase(〈δl 2 〉 1 )1kT 22th:1ms=∼ 6 × 10 −214π 2 Mf 2 m.QThus, thermal noise only affects a measurement to the extent that it changesthe amplitude of vibration during the time of the gravitational-wave burst.This change is similar to that produc<strong>ed</strong> by a gravitational wave of amplitude6 × 10 −21 . It follows that, if thermal noise were the only noise source,bars would be operating at around 10 −20 today. Bar groups expect in fact toreach this level during the next few years, as they r<strong>ed</strong>uce the other competingsources of noise. Notice that the effect of thermal noise has nothing todo with the frequency of the disturbance, so it is not the reason that barsobserve near their resonant frequency. In fact, both thermal impulses andgravitational-wave forces are mechanical forces on the bar, and the ratio oftheir induc<strong>ed</strong> vibrations is the same at all frequencies for a given appli<strong>ed</strong>impulsive force.• Sensor noise. Because the oscillations of the bar are very small, bars requirea transducer to convert the mechanical energy of vibration into electricalenergy, and an amplifier that increases the electrical signal to a level whereit can be record<strong>ed</strong>. If the amplifier were perfect, then the detector wouldin fact be broadband: it would amplify the smaller off-resonant responsesjust as well as the on-resonance ones. Conversely, real bars are narrow-bandbecause of sensor noise, not because of their mechanical resonance.Unfortunately sensing is not perfect: amplifiers introduce noise and thismakes small amplitudes harder to measure. The amplitudes of vibration arelargest in the resonance band near the resonant frequency f 0 , so amplifiernoise limits the detector sensitivity to frequencies near f 0 . Now, the signal(a typical gravitational-wave burst) has a duration time τ w ∼ 1 ms, so theamplifier’s bandwidth should be at least 1τ w in order for it to be able torecord a signal every τ w . In other words, bars require amplifiers with very


The physics of resonant mass detectors 37small noise in a large bandwidth (∼1000 Hz) near f 0 (note that this band ismuch larger than f Q). Today typical bandwidths of realizable amplifiersare 1 Hz, but in the very near future it is hop<strong>ed</strong> to extend these to 10 Hz, andeventually to 100 Hz.• Quantum limit. According to the Heisenberg uncertainty principle, the zeropointvibrations of a bar with a frequency of 1 kHz have rms amplitude〈δl 2 〉 1 2quant =( )1Ð 2∼ 4 × 10 −21 m.2π MfThis is bigger than the expect<strong>ed</strong> signal, and comparable to the thermal limitover 1 ms. It represents the accuracy with which one can measure theamplitude of vibration of the bar. So as soon as current detectors improvetheir thermal limits, they will run into the quantum limit, which must beovercome before a signal at 10 −21 can be seen with such a detector. Oneway to overcome this limit is by increasing the size of the detector and evenby making it spherical. This increases its mass dramatically, pushing thequantum limit down below 10 −21 .Another way around the quantum limit is to avoid measuring δl, but insteadto measure other observables. After all, the goal is to infer the gravitationalwaveamplitude, not to measure the state of vibration of the bar. It is possibleto define a pair of conjugate observables that have the property that one ofthem can be measur<strong>ed</strong> arbitrarily accurately repeat<strong>ed</strong>ly, so that the resultinginaccuracy of knowing the conjugate variable’s value does not disturb thefirst variable’s value. Then, if the first variable responds to the gravitationalwave, the gravitational wave may be measur<strong>ed</strong> accurately, even though thefull state of the bar is poorly known. This method is call<strong>ed</strong> ‘back reactionevasion’. The theory was develop<strong>ed</strong> in a classic paper by Caves et al [8].However, no viable schemes to do this have been demonstrat<strong>ed</strong> for bardetectors so far.3.3.1 New bar detectors and their capabilitiesResonant-mass detectors are limit<strong>ed</strong> by properties of materials and, as we havejust explain<strong>ed</strong>, they have their best sensitivity in a narrow band around theirresonant frequency. However, they can usefully explore higher frequencies (above500 Hz), where the interferometer noise curves are rising (see earlier figures).From the beginning, bars were design<strong>ed</strong> to detect bursts. If the burstradiation carries significant energy in the bar’s bandwidth, then the bar can dowell. Standard assumptions about gravitational collapse suggest a signal with abroad spectrum to 1 kHz or more, so that most of the sensitive bars today wouldbe suit<strong>ed</strong> to observe such a signal. Binary coalescence has a spectrum that peaksat low frequencies, so bars are not partiularly well suit<strong>ed</strong> for such signals. Onthe other hand, neutron-star and stellar-mass black-hole normal modes range in


38 <strong>Gravitational</strong>-wave detectorsfrequency from about 1 kHz up to 10 kHz, so suitably design<strong>ed</strong> bars could, inprinciple, go after these interesting signals.A bar gets all of its sensitivity in a relatively narrow bandwidth, so if a barand an interferometer can both barely detect a burst of amplitude 10 −20 , thenthe bar has much greater sensitivity than the interferometer in its narrow band,and much worse at other frequencies. This has l<strong>ed</strong> recently to interest in bars asdetectors of continuous signals. If the signal frequency is in the observing band ofthe bar, it can do very well compar<strong>ed</strong> to interferometers. Signals from millisecondpulsars and possible signals from x-ray binaries are suitable if they have the rightfrequency. However, most known pulsars will radiate at frequencies rather lowcompar<strong>ed</strong> to the operating frequencies of present-day bars.The excellent sensitivity of bars in their narrow bandwidth also suits themto detecting stochastic signals. Cross-correlations of two bars or of bars withinterferometers can be better than searches with first-generation interferometers[9]. One gets no spectral information, of course, and in the long run expect<strong>ed</strong>improvements in interferometers will overtake bars in this regard.Today’s best bar detectors are orders of magnitude more sensitive than theoriginal Weber bar. Two ultra-cryogenic bars have been built and are operatingat thermodynamic temperatures below 100 mK: NAUTILUS [10] at Frascati, nearRome, and [11] in Legnaro. With a mass of several tons, these may be the coldestmassive objects ever seen anywhere in the universe. These are expect<strong>ed</strong> soon toreach a sensitivity of 10 −20 near 1 kHz. Already they are performing coincidenceexperiments with bars at around 4 K at Perth, Australia, and at LSU.Proposals exist in the Netherlands, Brazil, Italy, and the USA for sphericalor icosah<strong>ed</strong>ral detectors (see links from [10]). These detectors have moremass, so they could reach 10 −21 near 1 kHz. Because of their shape, theyhave omnidirectional antenna patterns; if they are instrument<strong>ed</strong> so that all fiveindependent fundamental quadrupolar modes of vibration can be monitor<strong>ed</strong>, theycan do all-sky observing and determine directions as well as verify detectionsusing coincidences between modes of the same antenna.3.4 A detector in spaceAs we have not<strong>ed</strong> earlier, gravitational <strong>waves</strong> from astronomical objects atfrequencies below 1 Hz are obscur<strong>ed</strong> by Earth-bas<strong>ed</strong> gravity-gradient noise.Detectors must go into space to observe in this very interesting frequency range.The LISA [12] mission is likely to be the first such mission to fly. LISAwill be a triangular array of spacecraft, with arm lengths of 5 × 10 6 km, orbitingthe Sun in the Earth’s orbit, about 20 ◦ behind the Earth. The spacecraft will bein a plane inclin<strong>ed</strong> to the ecliptic by 60 ◦ . The three arms can be combin<strong>ed</strong> indifferent ways to form two independent interferometers. During the mission theconfiguration of spacecraft rotates in its plane, and the plane rotates as well, sothat LISA’s antenna pattern sweeps the sky.


A detector in space 39LISA has been nam<strong>ed</strong> a cornerstone mission of the European Space Agency(ESA), and NASA has recently form<strong>ed</strong> its own team to study the same mission,with a view toward a collaboration with ESA. LISA will be sensitive in a rangefrom 0.3 mHz to about 0.1 Hz, and it will be able to detect known binarystar systems in the Galaxy and binary coalescences of supermassive black holesanywhere they occur in the universe. A joint ESA–NASA project looks verylikely, aiming at a launch around 2010. A technology demonstration missionmight be launch<strong>ed</strong> in 2005 or 2006.LISA’s technology is fascinating. We can only allude to the most interestingparts of the mission here. A full description can be found in the pre-Phase A studydocument [13]. The most innovative aspect of the mission is drag-free control. Inorder to guarantee that the interferometry is not disturb<strong>ed</strong> by external forces, suchas fluctuations in solar radiation pressure, the mirror that is the reference pointfor the interferometry is on a free mass inside the spacecraft. The spacecraft actsas an active shield, sensing the position of the free mass, firing jets to counteractexternal forces on itself and ensure that it does not disturb the free mass. Thejets themselves are remarkable, in that they must be very weak compar<strong>ed</strong> to mostspacecraft’s control jets, and they must be capable of very precise control. Theywill work by expelling streams of ions, accelerat<strong>ed</strong> and controll<strong>ed</strong> by a highvoltageelectric field. Fuel for these jets is not a problem: 1 g will be enough fora mission lifetime of ten years!LISA interferometry is not done with reflection from mirrors. When a laserbeam reaches one spacecraft from the other, it is too weak to reflect: the sendingspacecraft would only get the occasional photon! Instead, the incoming lightis sens<strong>ed</strong>, and an on-board laser is slav<strong>ed</strong> to it, returning an amplifi<strong>ed</strong> beamwith the same phase and frequency as the incoming one. No space mission hasyet implement<strong>ed</strong> this kind of laser-transponding. The LISA team had to ensurethat there was enough information in all the signals to compensate for inevitablefrequency fluctuations among all six on-board lasers.A further serious problem that the LISA team had to solve was howto compensate for the relative motions of the spacecraft. The laser signalsconverging on a single spacecraft from the other two corners will be Dopplershift<strong>ed</strong> so that their fringes change at MHz frequencies. This has to be sens<strong>ed</strong> onboard and remov<strong>ed</strong> from the signal that is sent back to Earth, which can only besampl<strong>ed</strong> a few tens of times per second.When LISA flies it will, on a technical as well as a scientific level, be aworthy counterpart to its Earth-bas<strong>ed</strong> interferometer cousins!3.4.1 LISA’s capabilitiesIn the low-frequency LISA window, most sources will be relatively long liv<strong>ed</strong>,at least a few months. During an observation, LISA will rotate and change itsvelocity by a significant amount. This will induce Doppler shifts into the signals,and modulate their amplitudes, so that LISA should be able to infer the position,


40 <strong>Gravitational</strong>-wave detectorsLISA and Binaries in the Galaxygravitational wave amplitude10 -2010 -2110 -2210 -2310 -2310 -24Closest resolv<strong>ed</strong> binariesResolv<strong>ed</strong> binaries at GC4U1820-30confusion noise threshold (est.)LISA instrumental noise threshold(1 yr observationat S/N = 5)known close binariesknown IWDB's10 -4 10 -3 10 -2 10 -1 10 0frequency (Hz)LISA and Massive Black Holes at z=1Effective amplitudes for filter<strong>ed</strong> signals from coalescing binary sytemseffective gravitational wave amplitude10 -1910 -2010 -2110 -2210 -232 x 10 7 M o2 x 10 6 M o2 x 10 5 M o2 x 10 4 M oconfusion noise threshold (est.)10 M o + 10 6 M o BH10 -4 10 -3 10 -2 10 -1 10 0frequency (Hz)LISA instrumental noise threshold(1 yr observation at S/N = 5)Figure 3.5. LISA sensitivity to binary systems in the Galaxy (top) and to massive blackhole coalescences (bottom). The top figure is calibrat<strong>ed</strong> in the intrinsic amplitude of thesignal, and the noise curve shows the detection threshold (5σ ) for a one-year observation.It also shows the confusion limit due to unresolv<strong>ed</strong> binary systems. The bottom panelshows the effective amplitude of signals from coalescences of massive black holes. Sincesome such events last less than one year, what is shown is the expect<strong>ed</strong> signal-to-noise ratioof the observation.polarization and amplitude of sources entirely from its own observations. Belowabout 1 mHz, this information weakens, because the wavelength of the radiationbecomes comparable to or greater than the radius of LISA’s orbit. The amplitudemodulation is the only directional information in this frequency range.


<strong>Gravitational</strong> and electromagnetic <strong>waves</strong> compar<strong>ed</strong> and contrast<strong>ed</strong> 413.5 <strong>Gravitational</strong> and electromagnetic <strong>waves</strong> compar<strong>ed</strong> andcontrast<strong>ed</strong>To conclude this lecture it is useful to discuss the most important differences andsimilarities between gravitational <strong>waves</strong> and electromagnetic ones. We do this inthe form of a table.Table 3.1.ElectromagnetismTwo signs of charges—large bodiesusually neutral—<strong>waves</strong> usually emitt<strong>ed</strong>by single particles, often incoherently—<strong>waves</strong> carry ‘local’ information.A genuine physical force, acting differentlyon different bodies. Detect<strong>ed</strong> bymeasuring accelerations.Maxwell’s equations are linear. Physicalfield is F µν (E and B). Gauge fieldis vector potential A.Source is charge-current density J µ .Charge creates electric field, currentmagnetic field.Moderately strong force on the atomicscale: e2 /4πɛ 0Gm 2 = 10 39 .pWave generation for A µ : ∂ β ∂ β A µ =4πɛ 0 J µ in a convenient gauge (Lorentzgauge).Propagate at the spe<strong>ed</strong> of light, amplitudefalls as 1/r.Conservation of charge ⇒ radiation bylow-velocity charges is dominat<strong>ed</strong> bydipole component.General relativityOne sign of mass—gravityaccumulates—<strong>waves</strong> emitt<strong>ed</strong> morestrongly by larger body—<strong>waves</strong> carry‘global’ information.Equivalence principle: gravity affectsall bodies in the same way. Represent<strong>ed</strong>as a spacetime curvature rather than aforce. Detect<strong>ed</strong> only by tidal forces—differential accelerations.Einstein’s equations are nonlinear.Physical field is Riemann curvature tensorR µναβ . Gauge fields are metric g µνand connection Ɣµν α . Gauge transformationsare coordinate transformations.Source is stress-energy tensor T µν .Mass creates a Newtonian-like field,momentum as gravito-magnetic effects.Stress creates field too.Weaker than ‘weak’ interaction.Wave generation for h µν = g µν − η µν :∂ β ∂ β (h µν − 1 2 η µνh α α) = 8πT µν in aconvenient gauge.Propagate at the spe<strong>ed</strong> of light, amplitudefalls as 1/r.Conservation of mass and momentum⇒ radiation by low-velocity masses isdominat<strong>ed</strong> by quadrupole component.


42 <strong>Gravitational</strong>-wave detectorsTable 3.1. (Continu<strong>ed</strong>)ElectromagnetismSimple detector: oscillating charge.Action is along a line, transverse to th<strong>ed</strong>irections of propagation. Spin s = 1and two states of linear polarization thatare inclin<strong>ed</strong> to each other at an angle of90 ◦ .Strength of force ⇒ <strong>waves</strong> scatter andrefract easily.Local energy and flux well defin<strong>ed</strong>:Poynting vector etc.Multipole expansion in slow-motionlimit is straightforward, radiation reactionwell defin<strong>ed</strong>.Exact solutions, containing <strong>waves</strong>, areavailable and can guide the constructionof approximation methods for morecomplicat<strong>ed</strong> situations.General relativitySimple detector: distort<strong>ed</strong> ring ofmasses. Action is elliptic in aplane transverse to the direction ofpropagation. Spin s = 2 and twostates of linear polarization that areinclin<strong>ed</strong> to each other at an angle of 45 ◦ .Equivalence principle ⇒ action dependsonly on h µν , which is dimensionless.Weakness of gravity ⇒ <strong>waves</strong> propagatealmost undisturb<strong>ed</strong> and transfer energyvery weakly. Dimensionless amplitudeh is small.Equivalence principle ⇒ local energydensity cannot be defin<strong>ed</strong> exactly. Onlyglobal energy balance is exact.Multipole expansion different if fieldsare weak or strong. For quasi-Newtonian case fields are weak, and theresulting post-Newtonian expansion isdelicate. Radiation reaction is still notfully understood.Fully realistic exact solutions for dynamicalsituations of physical interestare not available. Extensive reliance onapproximation methods.


Chapter 4Astrophysics of gravitational-wave sourcesThere are a large number of possible gravitational-wave sources in the observablewaveband, which spans eight orders of magnitude in frequency: from 10 −4 Hz(lower bound of current space-bas<strong>ed</strong> detector designs) to 10 4 Hz (frequency limitof likely ground-bas<strong>ed</strong> detectors). Some of these sources are highly relativisticand not too massive, especially above 10 Hz: a black hole of mass 1000M ⊙has a characteristic frequency of 10 Hz, and larger holes have lower frequenciesin inverse proportion to the mass. Neutron stars have even higher characteristicfrequencies. Other systems are well describ<strong>ed</strong> by Newtonian dynamics, such asbinary orbits.For nearly-Newtonian sources the post-Newtonian approximation (seechapter 6) provides a good framework for calculating gravitational <strong>waves</strong>. Morerelativistic systems, and unusual sources like the early universe, require moresophisticat<strong>ed</strong> approaches (see chapter 7).4.1 Sources detectable from ground and from space4.1.1 Supernovae and gravitational collapseThe longest expect<strong>ed</strong> and still probably the least understood source, gravitationalcollapse is one of the most violent events known to astronomy. Yet, because wehave little direct information about the deep interior, we cannot make reliablepr<strong>ed</strong>ictions about the gravitational radiation from it.Supernovae are trigger<strong>ed</strong> by the gravitational collapse of the interiordegenerate core of an evolv<strong>ed</strong> star. According to current theory the result shouldbe a neutron star or black hole. The collapse releases an enormous amountof energy, about 0.15M ⊙ c 2 , most of which is carri<strong>ed</strong> away by neutrinos; anuncertain fraction is convert<strong>ed</strong> into gravitational <strong>waves</strong>. One mechanism forproducing this radiation could be dynamical instabilities in the rapidly rotatingcore before it becomes a neutron star. Another likely source of radiation is the r-mode instability (see chapter 7). This could release ∼0.1M ⊙ c 2 in radiation everytime a neutron star is form<strong>ed</strong>.43


44 Astrophysics of gravitational-wave sourcesHowever, both kinds of mechanisms are difficult to model. The problem withgravitational collapse is that perfectly spherical motions do not emit gravitational<strong>waves</strong>, and it is still not possible to estimate in a reliable way the amountof asymmetry in gravitational collapse. Even modern computers are not ableto perform realistic simulations of gravitational collapse in three dimensions,including all the important nuclear reactions and neutrino- and photon-transport.Similarly, it is hard to model the r-mode instability because its evolution dependson nonlinear hydrodynamics and on poorly known physics, such as the coolingand viscosity of neutron stars.An alternative approach is to use general energy considerations. If, forexample, we assume that 1% of the available energy is convert<strong>ed</strong> into gravitationalradiation, then, from formulae we will derive in the next chapter, the amplitude hwould be large enough to be detect<strong>ed</strong> by the first ground-bas<strong>ed</strong> interferometers(LIGO/GEO600/VIRGO) at the distance of Virgo Cluster (18 Mpc) if theemission centres at 300 Hz. Moreover, bar and spherical-mass detectors withan effective sensitivity of 10 −21 and the right resonant frequency could see thesesignals as well.The uncertainties in our pr<strong>ed</strong>ictions have a positive aspect: it is clear that ifwe can detect radiation from supernovae, we will learn much that we do not knowabout the end stages of stellar evolution and about neutron-star physics.4.1.2 Binary starsBinary systems have given us our best proof of the reliability of general relativityfor gravitational <strong>waves</strong>. The most famous example of such systems is thebinary pulsar PSR1916+16, discover<strong>ed</strong> by Hulse and Taylor in 1974; they wereaward<strong>ed</strong> the Nobel Prize for this discovery in 1993. From the observations ofthe modulation of the pulse period as the stars move in their orbits, one knowsmany important parameters of this system (orbital period, eccentricity, masses ofthe two stars, etc), and the data also show directly the decrease of the orbitalperiod due to the emission of gravitational radiation. The observ<strong>ed</strong> value is2.4 × 10 −12 s/s. Post-Newtonian theory allows one to pr<strong>ed</strong>ict this from the othermeasur<strong>ed</strong> parameters of the system, without any free parameters (see chapter 7);the pr<strong>ed</strong>iction is 2.38 × 10 −12 , in agreement within the measurement errors.Unfortunately the radiation from the Hulse–Taylor system will be too weakand of too low a frequency to be detectable by LISA.4.1.3 Chirping binary systemsIf a binary gives off enough energy for its orbit to shrink by an observable amountduring an observation, it is said to chirp: as the orbit shrinks, the frequency andamplitude go up. LISA will see a few chirping binaries. If a binary systemis compact enough to radiate above 10 −3 Hz, it will always chirp within oneyear, provid<strong>ed</strong> its components have a mass above about 1M ⊙ . If they are above


Sources detectable from ground and from space 45Table 4.1. The range for detecting a 2 × 1.4M ⊙ NS binary coalescence. The threshold fordetection is taken to be 5σ . The binary and detector orientations are assum<strong>ed</strong> optimum.The average S/N ratio for randomly orient<strong>ed</strong> systems is r<strong>ed</strong>uc<strong>ed</strong> from the optimum by1/ √ 5.Detector: TAMA300 GEO600 LIGO I VIRGO LIGO IIRange (S/N = 5) 3 Mpc 14 Mpc 30 Mpc 36 Mpc 500 Mpcabout 10 3 M ⊙ , the binary will go all the way to coalescence within the one-yearobservation.Chirping binary systems are more easily detectable than gravitationalcollapse events because one can model with great accuracy the gravitationalwaveform during the inspiral phase. There will be radiation, possibly withconsiderable energy, during the poorly understood plunge phase (when the objectsreach the last stable orbit and fall rapidly towards one another) and during themerger event, but the detectability of such systems rests on tracking their orbitalemissions.The major uncertainty about this kind of source is the event rate. Currentpulsar observations suggest that there will be ∼1 coalescence per year of a Hulse-Taylor binary out to about 200 Mpc. This is a lower limit on the event rate, since itcomes from systems we actually observe. It is possible that there are other kindsof binaries that we have no direct knowl<strong>ed</strong>ge of, which will boost the event rate.Theoretical modelling of binary populations gives a wide spectrum ofmutually inconsistent pr<strong>ed</strong>ictions. Some authors [14] suggest that there may bea large population that escapes pulsar surveys but brings the nearest neutron starcoalescence in one year as far as 30 Mpc, only slightly farther than the Virgocluster; but other models [15] put the rate near to the observational limit.The most exciting motivation for detecting coalescing binaries is that theycould be associat<strong>ed</strong> with gamma-ray bursts. The event rates are consistent,and neutron stars are able to provide the requir<strong>ed</strong> energy. If gamma-burstsare associat<strong>ed</strong> with neutron-star coalescence, then observations of coalescenceradiation should be follow<strong>ed</strong> within a second or so by a strong gamma-ray burst.LISA will see a few chirping binaries in the Galaxy, but the sensitivity ofthe first generation of ground-bas<strong>ed</strong> detectors is likely to be too poor to see manysuch events (see table 4.1).A certain fraction of such systems could contain black holes instead ofneutron stars. In fact black holes should be overrepresent<strong>ed</strong> in binary systems(relative to their birth rate) because their formation is much less likely to disrupta binary system (there is much less mass lost) than the formation of a neutronstar would be. Pulsar observations have not yet turn<strong>ed</strong> up a black-hole/neutronstarsystem, and of course one does not expect to see binary black holes


46 Astrophysics of gravitational-wave sourcesTable 4.2. The range for detecting a 10M ⊙ black-hole binary. Conventions as in table 4.1.Detector: GEO600 LIGO I VIRGO LIGO IIRange (S/N = 5) 75 Mpc 160 Mpc 190 Mpc 2.6 Gpcelectromagnetically. So we can only make theoretical estimates, and there arebig uncertainties.Some evolution calculations [14] suggest that the coalescence rate of BH–BH systems may be of the same order as the NS–NS rate. Other models [15]suggest it could even be zero, because stellar-wind mass loss (significant in verymassive stars) could drive the stars far apart before the second BH forms, leadingto coalescence times longer than the age of the universe. A recent proposalidentifies globular clusters as ‘factories’ for binary black holes, forming binariesby three-body collisions and then expelling them [16]. Gamma-ray bursts mayalso come from black-hole/neutron-star coalescences. If the more optimistic eventrates are correct, then black-hole coalescences may be among the first sourcesdetect<strong>ed</strong> by ground-bas<strong>ed</strong> detectors (table 4.2).4.1.4 Pulsars and other spinning neutron starsThere are a number of ways in which a spinning neutron star may give off acontinuous stream of gravitational <strong>waves</strong>. They will be weak, so they will requirelong continuous observation times, up to many months. Here are some possibleemission mechanisms for neutron stars.The r-modes. Neutron stars are born hot and probably rapidly rotating.Before they cool (during their first year) they have a family of unstable normalmodes, the r-modes. These modes are excit<strong>ed</strong> to instability by the emissionof gravitational radiation, as pr<strong>ed</strong>ict<strong>ed</strong> originally by Andersson [17]. They areparticularly interesting theoretically because the radiation is gravitomagnetic,generat<strong>ed</strong> by mass currents rather than mass asymmetries. We will study thetheory of this radiation in chapter 6. In chapter 7 we will discuss how the emissionof this radiation excites the instability (the CFS instability mechanism).Being unstable, young neutron stars will presumably radiate away enoughangular momentum to r<strong>ed</strong>uce their spin and become stable. This could lower thespin of a neutron star to ∼100 Hz within one year after its formation [18]. Theenergy emitt<strong>ed</strong> in this way should be a good fraction of the star’s binding energy,so in principle this radiation could be detect<strong>ed</strong> from the Virgo Cluster by LIGOII, provid<strong>ed</strong> match<strong>ed</strong> filtering can be us<strong>ed</strong> effectively.We discuss a possible stochastic background of gravitational <strong>waves</strong> from ther-modes below.Accreting neutron stars (figure 4.1) are the central objects of most of the


Sources detectable from ground and from space 47Figure 4.1. Accreting neutron star in a low-mass x-ray binary system.binary x-ray sources in the Galaxy. Astronomers divide them into two distinctgroups: the low-mass and high-mass binaries, according to the mass of thecompanion star. In these systems mass is pull<strong>ed</strong> from the low- or high-mass giantby the tidal forces exert<strong>ed</strong> by its neutron star companion. In low-mass x-raybinaries (LMXBs) the accretion lasts long enough to spin the neutron star up tothe rotation rates of millisecond pulsars. Astronomers have therefore suppos<strong>ed</strong>for some time that the neutron stars in LMXBs would have a range of spins, fromnear zero (young systems) to near 500 or 600 Hz (at the end of the accretionphase). Until the launch of the Rossi X-ray Timing Explorer (RXTE), there wasno observational evidence for the neutron star spins. However, in the last twoyears there has been an accumulation of evidence that most, if not all, of thesestars have angular velocities in a narrow range around 300 Hz [19]. It is not knownyet what mechanism regulates this spin, but a strong candidate is the emission ofgravitational radiation.A novel proposal by Bildsten [20] suggests that the temperature gradientacross a neutron star that is accreting preferentially at its magnetic poles shouldlead to a composition and hence a density gradient in the deep crust. Spinning at300 Hz, such a star could radiate as much as it accretes. It would then be a steadysource for as long as accretion lasts, which could be millions of years.In this model the gravitational-wave energy flux is proportional to theobserv<strong>ed</strong> x-ray energy flux. The strongest source in this model is Sco X-1,which could be detect<strong>ed</strong> by GEO600 in a two-year-long narrow-band mode ifthe appropriate match<strong>ed</strong> filtering can be done. LIGO II would have no difficultyin detecting this source.Older stars may also be lumpy. For known pulsars, this is constrain<strong>ed</strong> by therate of spindown: the energy radiat<strong>ed</strong> in gravitational <strong>waves</strong> cannot exce<strong>ed</strong> the


48 Astrophysics of gravitational-wave sourcestotal energy loss. In most cases, this limit is rather weak, and stars would have tosustain strains in their crust of order 10 −3 or more. It is unlikely that crusts couldsustain this kind of strain, so the observational limits are probably significantoverestimates for most pulsars. However, millisecond pulsars have much slowerspindown rates, and it would be easier to account for the strain in their crusts,for example as a remnant Bildsten asymmetry. Such stars could, in principle, beradiating more energy in gravitational <strong>waves</strong> than electromagnetically.Observations of individual neutron stars would be rich with informationabout astrophysics and fundamental nuclear physics. So little is known aboutthe physics of these complex objects that the incentive to observe their radiationis great.However, making such observations presents challenges for data analysis,since the motion of the Earth puts a strong phase modulation on the signal, whichmeans that even if its rest-frame frequency is constant it cannot be found bysimple Fourier analysis. More sophisticat<strong>ed</strong> pattern-matching (match<strong>ed</strong>-filtering)techniques are ne<strong>ed</strong><strong>ed</strong>, which track and match the signal’s phase to within onecycle over the entire period of measurement. This is not difficult if the source’slocation and frequency are known, but the problem of doing a wide-area search forunknown objects is very challenging [21]. Moreover, if the physics of the sourceis poorly known, such as for LMXBs or r-mode spindown, the job of building anaccurate family of templates is a difficult one. These questions are the subject ofmuch research today, but they will ne<strong>ed</strong> much more in the future.4.1.5 Random backgroundsThe big bang was the most violent event of all, and it may have creat<strong>ed</strong> asignificant amount of gravitational radiation. Other events in the early universemay also have creat<strong>ed</strong> radiation, and there may be backgrounds from morerecent epochs. We have seen earlier, for example, that compact binary systemsin the Galaxy will merge into a confusion-limit<strong>ed</strong> noise background in LISAobservations below about 1 mHz.Let us consider the r-modes as another important example. This process mayhave occurr<strong>ed</strong> in a good fraction of all neutron stars form<strong>ed</strong> since the beginningof star formation. The sum of all of their r-mode radiation will be a stochasticbackground, with a spectrum that extends from a lower cut-off of about 200 Hz inthe rest frame of the emitter to an upper limit that depends on the initial angularvelocity of stars. If significant star formation start<strong>ed</strong> at, say, a r<strong>ed</strong>shift of five,then this background should extend down to about 25 Hz. If 10 −3 of the mass ofthe Galaxy is in neutron stars, and each of them radiates 10% of its mass in thisradiation, then the gravitational-wave background should have a density equal to10 −4 of the mean cosmological density of visible stars. Express<strong>ed</strong> as a fraction gw of the closure density of the universe, per logarithmic frequency interval, thisconverts to r-modesgw (25–1000 Hz) ≈ 10 −8 –10 −7 .


Sources detectable from ground and from space 49This background would be easily detectable by LIGO II.There may also be a cosmological background from either topologicaldefects (e.g. cosmic strings) or from inflation (which amplifies initial quantumgravitational fluctuations as it does the scalar ones that lead to galaxy formation).Limits from COBE observations suggest that standard inflation could not producea background stronger than inflationgw ∼ 10 −14 today. This is too weak for anyof the plann<strong>ed</strong> detectors to reach, but it remains an important long-range goalfor the field. However, there could also be a component of background radiationthat depends on what happen<strong>ed</strong> before inflation: string cosmological models, forexample, pr<strong>ed</strong>ict spectra growing with frequency [22].First-generation interferometers are not likely to detect these backgrounds:they may not be able to go below the upper limit set by the requirementthat gravitational <strong>waves</strong> should not disturb cosmological nucleosynthesis, whichis gw = 10 −5 . (This limit does not apply to backgrounds generat<strong>ed</strong> afternucleosynthesis, like the r-mode background.) Bar detectors may do as well orbetter than the first generation of interferometers for a broad-spectrum primordialbackground: as we have not<strong>ed</strong> earlier, their noise levels within their resonancebands are very low. However, their frequencies are not right for the r-modebackground.Second-generation interferometers may be able to reach to 10 −11 of closureor even lower, by cross-correlation of the output of the two detectors. However,they are unlikely to get to the inflation target of 10 −14 . LISA may be able to go aslowas10 −10 (if we have a confident understanding of the instrumental noise), butit is likely to detect only the confusion background of binaries, which is expect<strong>ed</strong>to be much stronger than a cosmological background in the LISA band.4.1.6 The unexpect<strong>ed</strong>At some level, we are bound to see things we did not expect. LISA, with itshigh signal-to-noise ratios for pr<strong>ed</strong>ict<strong>ed</strong> sources, is particularly well plac<strong>ed</strong> to dothis. Most of the universe is compos<strong>ed</strong> of dark matter whose existence we caninfer only from its gravitational effects. It would not be particularly surprising ifa component of this dark matter produc<strong>ed</strong> gravitational radiation in unexpect<strong>ed</strong>ways, such as from binaries of small exotic compact objects of stellar mass. Wewill have to wait to see!


Chapter 5Waves and energyHere we discuss wave-like perturbations h µν of a general background metric g µν .The mathematics is similar to that of lineariz<strong>ed</strong> theory: h µν is a tensor with respectto background coordinate transformations (as it was for Lorentz transformationsin lineariz<strong>ed</strong> theory) and it undergoes a gauge transformation when one makes aninfinitesimal coordinate transformation. As in lineariz<strong>ed</strong> theory, we will assumethat the amplitude of the <strong>waves</strong> is small. Moreover, the <strong>waves</strong> must have awavelength that is short compar<strong>ed</strong> to the radius of curvature of the backgroundmetric. These two assumptions allow us to visualize the <strong>waves</strong> as small ripplesrunning through a curv<strong>ed</strong> and slowly changing spacetime.5.1 Variational principle for general relativityWe start our analysis of the small perturbation h µν by introducing the standardHilbert variational principle for Einstein’s equations. The field equations ofgeneral relativity can be deriv<strong>ed</strong> from an action principle using the Ricci scalarcurvature as the Lagrangian density. The Ricci scalar (second contraction of theRiemann tensor) is an invariant quantity which contains in addition to g µν and itsfirst derivatives also the second derivatives of g µν , so our action can be writtensymbolically as:I[g µν ] = 1 ∫R(g µν , g µν,α , g µν,αβ ) √ −g d 4 x (5.1)16πwhere √ −g is the square root of the determinant of the metric tensor. As usual invariational principles, the metric tensor components are vari<strong>ed</strong> g µν → g µν +h µν ,and one demands that the resulting change in the action should vanish to firstorder in any small variation h µν of compact support:50δI = I[g µµ + h µν ] − I[g µν ]= 116π∫ δ(R√ −g)δg µνh µν d 4 x + O(2) (5.2)


Variational principles and the energy in gravitational <strong>waves</strong> 51=− 1 ∫G µν √h µν −g d 4 x + O(2) (5.3)16πwhere ‘O(2)’ denotes terms quadratic and higher in h µν . All the divergencesobtain<strong>ed</strong> in the interm<strong>ed</strong>iate steps of this calculation integrate to zero since h µνis of compact support. This variational principle therefore yields the vacuumEinstein equations: G µν = 0.Let us consider how this changes if we include matter. This will help usto see how we can treat gravitational <strong>waves</strong> as a new kind of ‘matter’ field onspacetime.Suppose we have a matter field, describ<strong>ed</strong> by a variable (which mayrepresent a vector, a tensor or a set of tensors). It will have a Lagrangian densityL m = L m (, ,α ,...,g µν ) that depends on the field and also on the metric.Normally derivatives of the metric tensor do not appear in L m , since by theequivalence principle, matter fields should behave locally as if they were in flatspacetime, where of course there are no metric derivatives. Variations of L m withrespect to will produce the field equation(s) for the matter system, but here weare more interest<strong>ed</strong> in variations with respect to g µν , which is how we will findthe matter field contribution to the gravitational field equations. The total actionhas the form:∫I = (R + 16π L m ) √ −g d 4 x, (5.4)whose variation is∫ √ ∫δ(R −g)δI =h µν d 4 x + 16π ∂(L √m −g)h µν d 4 x. (5.5)δg µν ∂g µνThis variation must yield full Einstein equations, so we must have the followingresult for the stress-energy tensor of matter:T µν√ −g = 2 ∂ L √m −g, (5.6)∂g µνleading toG µν = 8πT µν . (5.7)This way of deriving the stress-energy tensor of the matter field has deepconnections to the conservation laws of general relativity, to the way ofconstructing conserv<strong>ed</strong> quantities when the metric has symmetries and to the socall<strong>ed</strong>pseudotensorial definitions of gravitational wave energy (see Landau andLifshitz 1962) [23]. We shall use it in the latter sense.5.2 Variational principles and the energy in gravitational<strong>waves</strong>Before we introduce the mathematics of gravitational <strong>waves</strong>, it is important tounderstand which geometries we are going to examine. We have said that these


52 Waves and energygeometries consist of a slowly and smoothly changing background metric whichis alter<strong>ed</strong> by perturbations of small amplitude and high frequency. If L and λ arethe characteristic lengths over which the background and ‘ripple’ metrics changesignificantly, we assume that the ratio λ/L will be very much smaller than unityand that |h µν | is of the same order of smallness as λ/L. In this way the totalmetric remains slowly changing on a macroscopic scale, while the high-frequencywave, when averag<strong>ed</strong> over several wavelengths, will be the principal source of thecurvature of the background metric. This is the ‘short-wave’ approximation [24].Obviously this is a direct generalization of the treatment in chapter 2.5.2.1 Gauge transformation and invarianceConsider an infinitesimal coordinate transformation generat<strong>ed</strong> by a vector fieldξ α ,x α → x α + ξ α . (5.8)In the new coordinate system, neglecting quadratic and higher terms in h αβ ,itisnot hard to show that the general gauge transformation of the metric ish µν → h µν − ξ µ;ν − ξ ν;µ , (5.9)where a semicolon denotes the covariant derivative. We assume that th<strong>ed</strong>erivatives of the coordinate displacement field are of the same order as the metricperturbation: |ξ α,β |∼|h αβ |.Isaacson [24] show<strong>ed</strong> that the gauge transformation of the Ricci andRiemann curvature tensors has the property( ) λ 2¯R µν (1) − R(1) µν ≈ (5.10)L( )¯R (1)λ 2αµβν − R(1) αµβν ≈ Lwhere R µν (1) and R (1)αµβνare the first order of Ricci and Riemann tensors (inpowers of perturbation h µν ) and an overbar denotes their values after the gaugetransformation. In our high-frequency limit, therefore, these tensors are gaugeinvariantto linear order, just as in lineariz<strong>ed</strong> theory.5.2.2 <strong>Gravitational</strong>-wave actionLet us suppose that we are in vacuum so we have only the metric, no matter fields,but we work in the high-frequency approximation. The full metric is g µν (smoothbackground metric) +h µν (high-frequency perturbation). Our purpose is to showthat the wave field can be treat<strong>ed</strong> as a ‘matter’ field, with a Lagrangian and its


Variational principles and the energy in gravitational <strong>waves</strong> 53own stress-energy tensor. To do this we have to expand the action out to secondorder in the metric perturbation,∫I[g µµ + h µν ] = R(g µν + h µν , g µν,α + h µν,α ,...) √ −g[g µν + h µν ]d 4 x∫= R(g µν ,...) √ ∫ √ δ(R −g)−g d 4 x +h µν d 4 x+ 1 2δg µν∫ ( ∂ 2 (R √ −g)∂g µν ∂g αβh µν h αβ + 2 ∂2 (R √ −g)∂g µν ∂g αβ,γh µν h αβ,γ+ ∂2 (R √ −g)h µν,τ h αβ,γ + 2 ∂2 (R √ )−g)h µν h αβ,γ τ d 4 x∂g µν,τ ∂g αβ,γ ∂g µν ∂g αβ,γ τ+ O(3).The first term is the action for the background metric g µν . The second termvanishes (see equation (5.2)), since we assume that the background metric is asolution of the Einstein vacuum equation itself, at least to lowest order.If we compare the above equation with equation (5.4), we can see that thethird term, complicat<strong>ed</strong> as it seems, is an effective ‘matter’ Lagrangian for thegravitational field. Inde<strong>ed</strong>, if one varies it with respect to h µν holding g µνfix<strong>ed</strong> (as we would do for a physical matter field on the background), then thecomplicat<strong>ed</strong> coefficients are fix<strong>ed</strong> and one can straightforwardly show that onegets exactly the linear perturbation of the Einstein tensor itself. Its vanishing isthe equation for the gravitational-wave perturbation h µν . In this way we haveshown that, for a small amplitude perturbation, the gravitational wave can betreat<strong>ed</strong> as a ‘matter’field with its own Lagrangian and field equations.Given this Lagrangian, we should be able to calculate the effective stressenergytensor of the wave field by taking the variations of the effective Lagrangianwith respect to g µν , holding the ‘matter’field h µν fix<strong>ed</strong>:withL (GW)√ −g = 132πT (GW)αβ √ −g = 2 ∂ L(GW) [g µν , h µν ] √ −g∂g αβ(5.11)( ∂ 2 (R √ −g)∂g µν ∂g αβh µν h αβ + 2 ∂2 (R √ −g)∂g µν ∂g αβ,γh µν h αβ,γ+ ∂2 (R √ −g)∂g µν,τ ∂g αβ,γh µν,τ h αβ,γ + 2 ∂2 (R √ −g)∂g µν ∂g αβ,γ τh µν h αβ,γ τ).(5.12)This quantity is quadratic in the wave amplitude h µν . It could be simplifi<strong>ed</strong>further by integrations by parts, such as by taking a derivative off h αβ,γ τ . Thiswould change the coefficients of the other terms. We will not ne<strong>ed</strong> to worry aboutfinding the ‘best’ form for the expression (4.12), as we now show.


54 Waves and energyAs in lineariz<strong>ed</strong> theory, so also in the general case, the quantity h µν behaveslike a tensor with respect to background coordinate transformations, and so doesT µν (GW) . However, it is not gauge-invariant and so it is not physically observable.Since the integral of the action is independent of coordinate transformations thathave compact support, so too is the integral of the effective stress-energy tensor.In practical terms, this makes it possible to localize the energy of a wave to withina region of about one wavelength in size where the background curvature doesnot change significantly. In fact, if we restrict our gauge transformations to havea length scale of a wavelength, and if we average (integrate) the stress-energytensor of the <strong>waves</strong> over such a region, then any gauge changes will be smallsurface terms.By evaluating the effective stress-energy tensor on a smooth backgroundmetric in a Lorentz gauge, and performing the averaging (denot<strong>ed</strong> by symbol〈···〉), one arrives at the Isaacson tensor:T (GW)αβ= 132π 〈h µν;αh µν ;β〉. (5.13)This is a convenient and compact form for the gravitational stress-energytensor. It localizes energy in short-wavelength gravitational <strong>waves</strong> to regions ofthe order of a wavelength. It is interesting to remind ourselves that our onlyexperimental evidence of gravitational <strong>waves</strong> today is the observation of the effecton a binary orbit of the loss of energy to the gravitational <strong>waves</strong> emitt<strong>ed</strong> by thesystem. So this energy formula, or equivalent ones, is central to our understandingof gravitational <strong>waves</strong>.5.3 Practical applications of the Isaacson energyIf we are far from a source of gravitational <strong>waves</strong>, we can treat the <strong>waves</strong> bylineariz<strong>ed</strong> theory. Then if we adopt the TT gauge and specialize the stress-energytensor of the radiation to a flat background, we getT (GW)αβ= 132π 〈hTT ij,α hTTij ,β〉. (5.14)Since there are only two components, a wave travelling with frequency f(wavenumber k = 2π f ) and with a typical amplitude h in both polarizationscarries an energy F gw equal to (see exercise (f) at the end of this lecture)F gw = π 4 f 2 h 2 . (5.15)Putting in the factors of c and G and scaling to reasonable values gives[ ]F gw = 3mWm −2 h 2 [ ]f 21 × 10 −22 , (5.16)1 kHz


Practical applications of the Isaacson energy 55which is a very large energy flux even for this weak a wave. It is twice the energyflux of a full moon! Integrating over a sphere of radius r, assuming a total durationof the event τ, and solving for h, again with appropriate normalizations, gives[ ] 1 [ ]h = 10 −21 E gw 2 r −1 [ ]f −1 [ τ ] − 120.01M ⊙ c 2 . (5.17)20 Mpc 1 kHz 1msThis is the formula for the ‘burst energy’, normaliz<strong>ed</strong> to numbers appropriate to agravitational collapse occurring in the Virgo cluster. It explains why physicistsand astronomers regard the 10 −21 threshold as so important. However, thisformula could also be appli<strong>ed</strong> to a binary system radiating away its orbitalgravitational binding energy over a long period of time τ, for example.5.3.1 Curvature produc<strong>ed</strong> by <strong>waves</strong>We have assum<strong>ed</strong> that the background metric satisfies the vacuum Einsteinequations to linear order, but now it is possible to view the full action principleas a principle for the background with a wave field h µν on it, and to let the waveenergy affect the background curvature [24]. This means that the background willactually solve, in a self-consistent way, the equationG αβ [g µν ] = 8πTαβ GW [g µν + h µν ]. (5.18)This does not contradict the vanishing of the first variation of the action, which wene<strong>ed</strong><strong>ed</strong> to use above, because now we have an Einstein tensor that is of quadraticorder in h µν , contributing a term of cubic order to the first-variation of the action,which is of the same order as other terms we have neglect<strong>ed</strong>.5.3.2 Cosmological background of radiationThis self-consistent picture allows us to talk about, for example, a cosmologicalgravitational wave background that contributes to the curvature of the universe.Since the energy density is the same as the flux (when c = 1), we haveϱ gw = π 4 f 2 h 2 , (5.19)but now we must interpret h in a statistical way. This will be treat<strong>ed</strong> in thecontribution by Babusci et al, but basically it is done by replacing h 2 by astatistical mean square amplitude per unit frequency (Fourier transform power),so that the energy density per unit frequency is proportional to f 2 | ˜h| 2 . It is thenconventional to talk about the energy density per unit logarithm of frequency,which means multiplying by f . The result, after being careful about averagingover all directions of the <strong>waves</strong> and all independent polarization components, isdϱ gwdln f= 4π 2 f 3 | ¯h( f )| 2 . (5.20)


56 Waves and energyFinally, what is of the most interest is the energy density as a fraction of theclosure or critical cosmological density, given by the Hubble constant H 0 asϱ c = 3H0 2/8π. The resulting ratio is the symbol gw( f ) that we met in theprevious lecture: gw ( f ) = 32π 3f 3 | ¯h( f )| 2 . (5.21)5.3.3 Other approaches3H 2 0We finish this lecture by observing that there is no unique approach to definingenergy for gravitational radiation or inde<strong>ed</strong> for any solution of Einstein’sequations. Historically this has been one of the most difficult areas for physiciststo get to grips with. In the textbooks you will find discussions of pseudotensors, ofenergy measur<strong>ed</strong> at null infinity and at spacelike infinity, of Noether theorems andformulae for energy, and so on. None of these are worse than we have present<strong>ed</strong>here, and in fact all of them are now known to be consistent with one another, ifone does not ask them to do too much. In particular, if one wants only to localizethe energy of a gravitational wave to a region of the size of a wavelength, andif the <strong>waves</strong> have short wavelength compar<strong>ed</strong> to the background curvature scale,then pseudotensors will give the same energy as the one we have defin<strong>ed</strong> here.Similarly, if one takes the energy flux defin<strong>ed</strong> here and evaluates it at null infinity,one gets the so-call<strong>ed</strong> Bondi flux, which was deriv<strong>ed</strong> by H Bondi in one of thepioneering steps in the understanding of gravitational radiation. Many of theseissues are discuss<strong>ed</strong> in the Schutz–Sorkin paper referr<strong>ed</strong> to earlier [23].5.4 Exercises for chapter 5(e)In the notes above we give the general gauge transformationh µν → h µν − ξ µ;ν − ξ ν;µ .Use the formula for the derivation of Einstein’s equations from an actionprinciple,δI = 1 ∫ √ δ(R −g)h µν d 4 x16π δg µνwithδ(R √ −g)=−G µν√ −g,δg µνbut insert a pure gauge h µν . Argue that since this is merely a coordinatetransformation, the action should be invariant. Integrate the variation of theaction to prove the contact<strong>ed</strong> Bianchi identityG µν ,ν = 0.


Exercises for chapter 5 57(f)This shows that the divergence-free property of G µν is closely relat<strong>ed</strong> to thecoordinate invariance of Einstein’s theory.Suppose a plane wave, travelling in the z-direction in lineariz<strong>ed</strong> theory, hasboth polarization components h + and h × . Show that its energy flux in thez-direction, T (GW)0z ,is〈T (GW)0z 〉= k232π (A2 + + A2 × ),where the angle brackets denote an average over one period of the wave.


Chapter 6Mass- and current-quadrupole radiationIn this lecture we focus on the wave amplitude itself, and how it and thepolarization depend on the motions in the source. Consider an isolat<strong>ed</strong> sourcewith a stress-energy tensor T αβ . As in chapter 2, the Einstein equation is( )− ∂2∂t 2 +∇2 h αβ =−16πT αβ (6.1)(h αβ = h αβ − 1 2 ηαβ h and h αβ ,β = 0). Its general solution is the following retard<strong>ed</strong>integral for the field at a position x i and time t in terms of the source at a positiony i and the retard<strong>ed</strong> time t − R:h αβ (x i , t) = 4∫ 1R T αβ (t − R, y i ) d 3 y, (6.2)where we defineR 2 = (x i − y i )(x i − y i ). (6.3)6.1 Expansion for the far field of a slow-motion sourceLet us suppose that the origin of coordinates is in or near the source, and the fieldpoint x i is far away. Then we define r 2 = x i x i and we have r 2 ≫ y i y i . Wecan, therefore, expand the term R in the dominator in terms of y i . The lowestorder is r, and all higher-order terms are smaller than this by powers of r −1 .Therefore, they contribute terms to the field that fall off faster than r −1 , and theyare negligible in the far zone. Therefore, we can simply replace R by r in th<strong>ed</strong>ominator, and take it out of the integral.The R inside the time-argument of the source term is not so simple. If wesuppose that T αβ does not change very fast we can substitute t − R by t − r (theretard<strong>ed</strong> time to the origin of coordinates) and expand( ) 1t − R = t − r + n i y i + O , with n i = x ir r , ni n i = 1. (6.4)58


Expansion for the far field of a slow-motion source 59The two conditions r ≫ y i y i and the slow-motion source, can be express<strong>ed</strong>quantitatively as:r ≫ ¯λR ≪ ¯λwhere ¯λ is the r<strong>ed</strong>uc<strong>ed</strong> wavelength ¯λ = λ/2π and R is the size of source.The terms of order r −1 are negligible for the same reason as above, but thefirst term in this expansion must be taken into account. It depends on the directionto the field point, given by the unit vector n i . We use this by making a Taylorexpansion in time on the time-argument of the source. The combin<strong>ed</strong> effect ofthese approximations ish αβ = 4 r∫[T αβ (t ′ , y i ) + T αβ ,0(t ′ , y i )n j y j + 1 2 T αβ ,00(t ′ , y i )n j n k y j y k+ 1 6 T αβ ,000(t ′ , y i )n j n k n l y j y k y l +···]d 3 y. (6.5)We will ne<strong>ed</strong> all the terms of this Taylor expansion out to this order.The integrals in expression (5.5) contain moments of the components of thestress-energy. It is useful to give these names. Use M for moments of the densityT 00 , P for moments of the momentum T 0i and S for the moments of the stressT ij . Here is our notation:∫∫M(t ′ ) = T 00 (t ′ , y i ) d 3 y, M j (t ′ ) = T 00 (t ′ , y i )y j d 3 y,∫∫M jk (t ′ ) = T 00 (t ′ , y i )y j y k d 3 y, M jkl (t ′ ) = T 00 (t ′ , y i )y j y k y l d 3 y,∫∫P l (t ′ ) = T 0l (t ′ , y i ) d 3 y, P lj (t ′ ) = T 0l (t ′ , y i )y j d 3 y,∫P ljk (t ′ ) = T 0l (t ′ , y i )y j y k d 3 y,∫∫S lm (t ′ ) = T lm (t ′ , y i ) d 3 y, S lmj (t ′ ) = T lm (t ′ , y i )y j d 3 y.These are the moments we will ne<strong>ed</strong>.Among these moments there are some identities that follow from theconservation law in lineariz<strong>ed</strong> theory, T αβ ,β = 0, which we use to replace tim<strong>ed</strong>erivatives of components of T by divergences of other components and thenintegrate by parts. The identities we will ne<strong>ed</strong> areṀ = 0, Ṁ k = P k , Ṁ jk = P jk + P kj , Ṁ jkl = P jkl + P klj + P ljk ,(6.6)Ṗ j = 0, Ṗ jk = S jk , Ṗ jkl = S jkl + S jlk . (6.7)


60 Mass- and current-quadrupole radiationThese can be appli<strong>ed</strong> recursively to show, for example, two further very usefulrelationsd 2 M jkdt 2 = 2S jk d 3 M jkl,dt 3 = 6Ṡ ( jkl) (6.8)where the round brackets on indices indicate full symmetrization.Using these relations and notations it is not hard to show thath 00 (t, x i ) = 4 r M + 4 r P j n j + 4 r S jk (t ′ )n j n k + 4 r Ṡ jkl (t ′ )n j n k n l +···(6.9)h 0 j (t, x i ) = 4 r P j + 4 r S jk (t ′ )n k + 4 r Ṡ jkl (t ′ )n k n l +··· (6.10)h jk (t, x i ) = 4 r S jk (t ′ ) + 4 r Ṡ jkl (t ′ )n l +···. (6.11)In these three formulae there are different orders of time-derivatives, but in factthey are evaluat<strong>ed</strong> to the same final order in the slow-motion approximation. Onecan see that from the gauge condition h aβ ,β = 0, which relates time-derivativesof some components to space-derivatives of others.In these expressions, one must remember that the moments are evaluat<strong>ed</strong>at the retard<strong>ed</strong> time t ′ = t − r (except for those moments that are constant intime), and they are multipli<strong>ed</strong> by components of the unit vector to the field pointn j = x j /r.6.2 Application of the TT gauge to the mass quadrupole fieldIn the expression for the amplitude that we deriv<strong>ed</strong> so far, the final terms arethose that represent the current-quadrupole and mass-octupole radiation. Theterms before them represent the static parts of the field and the mass-quadrupoleradiation. In this section we treat just these terms, placing them into the TT gauge.This will be simpler than treating it all at once, and the proc<strong>ed</strong>ure for the nextterms will be a straightforward generalization.6.2.1 The TT gauge transformationsWe are already in Lorentz gauge, and this can be check<strong>ed</strong> by taking derivativesof the expressions for the field that we have deriv<strong>ed</strong> above. However, we aremanifestly not in the TT gauge. Making a gauge transformation consists ofchoosing a vector field ξ α and modifying the metric byThe corresponding expression for the potential h αβ ish αβ → h αβ − ξ α,β − ξ β,α . (6.12)h αβ → h αβ + ξ α,β + ξ α,β − η αβ ξ µ ,µ. (6.13)


Application of the TT gauge to the mass quadrupole field 61For the different components this implies changesδh 00 = ξ 0,0 + ξ j , j (6.14)δh 0 j = ξ 0, j + ξ j,0 (6.15)δh jk = ξ j,k + ξ k, j − δ jk ξ µ ,µ (6.16)where δ jk is the Kronecker delta (unit matrix). In practice, when takingderivatives, the algebra is vastly simplifi<strong>ed</strong> by the fact that we are keeping onlyr −1 terms in the potentials. This means that spatial derivatives do not act on1/r but only on t ′ = t − r. It follows that ∂t ′ /∂x j =−n j , and ∂h(t ′ )/∂x j =−ḣ(t ′ )n j .It is not hard to show that the following vector field puts the metric into theTT gauge to the order we are working:ξ 0 = 1 r Pk k + 1 r P jk n j n k + 1 r Sl lkn k + 1 r Sijk n i n j n k , (6.17)ξ i = 4 r Mi + 4 r Pij n j − 1 r Pk kn i − 1 r P jk n j n k n i + 4 r Sijk n j n k− 1 r Sl lkn k n i − 1 r S jlk n j n l n k n i . (6.18)6.2.2 Quadrupole field in the TT gaugeThe result of applying this gauge transformation to the original amplitudes ish TT00 = 4M r , (6.19)h TT0i = 0, (6.20)h TTij = 4 [⊥ ik ⊥ jl S lk + 1 ]r2 ⊥ij (S kl n k n l − S k k) . (6.21)Remember that here we are not including Ṡ jkl , because it is a third-order effect.The notation ⊥ ik represents the projection operator perpendicular to th<strong>ed</strong>irection n i to the field point.⊥ jk = δ jk − n j n k . (6.22)It can be verifi<strong>ed</strong> that this tensor is transverse to the direction n iprojection, in the sense that it projects to itselfand is a⊥ jk n k = 0, ⊥ jk ⊥ k l = ⊥ jl . (6.23)The spherical component of the field is not totally eliminat<strong>ed</strong> in this gaugetransformation: the time–time component of the metric must contain the constantNewtonian field of the source. (In fact we have succe<strong>ed</strong><strong>ed</strong> in eliminating the


62 Mass- and current-quadrupole radiationdipole, or momentum part of the field, which is also part of the non-wave solution.Our gauge transformation has incorporat<strong>ed</strong> a Lorentz transformation that has putus into the rest frame of the source.) The time-dependent part of the field is nowpurely spatial, transverse (because everything is multipli<strong>ed</strong> by ⊥), and traceless(as can be verifi<strong>ed</strong> by explicit calculation).The expression for the spatial part of the field actually does not depend onthe trace of S jk , as can be seen by constructing the trace-free part of the tensor,defin<strong>ed</strong> as:˜S jk = S jk − 1 3 δ jk S l l. (6.24)In fact, it is more conventional to use the mass moment here instead of the stress,so we also define˜M jk = M jk − 1 3 δ jk Ml l , ˜S jk = 1 d 2 ˜M jk2 dt 2 . (6.25)In terms of ˜M the far field is¯h TTij = 2 (⊥ ik ⊥ jl ¨˜M kl + 1 r2 ⊥ij ¨˜M )kl n l n k . (6.26)This is the usual formula for the mass-quadrupole field. In textbooks the notationis somewhat different than we have adopt<strong>ed</strong> here. In particular, our tensor ˜Mis what is call<strong>ed</strong> I– in Misner et al (1973) and Schutz (1985). It is the basis ofmost gravitational-wave source estimates. We have deriv<strong>ed</strong> it only in the contextof lineariz<strong>ed</strong> theory, but remarkably its form is identical if we go to the post-Newtonian approximation, where the gravitational <strong>waves</strong> are a perturbation ofthe Newtonian spacetime rather than of flat spacetime.Given this powerful formula, it is important to try to interpret it andunderstand it as fully as possible. One obvious conclusion is that the dominantsource of radiation, at least in the slow-motion limit, is the second tim<strong>ed</strong>erivativeof the second moment of the mass density T 00 (the mass-quadrupolemoment). This is a very important difference between gravitational <strong>waves</strong>and electromagnetism, in which the most important source is the electricdipole.In our case the mass-dipole term is not able to radiate because it isconstant, reflecting conservation of the linear momentum of the source. Inelectromagnetism, however, if the dipole term is absent for some reason (allcharges positive, for example) then the quadrupole term dominates and it looksvery similar to equation (6.26).6.2.3 Radiation patterns relat<strong>ed</strong> to the motion of sourcesThe projection operators in equation (6.26) show that the radiative field istransverse, as we expect. However, the form of equation (6.26) hides two equallyimportant messages:


Application of the TT gauge to the mass quadrupole field 63• the only motions that produce the radiation are the ones transverse to the lineof sight; and• the induc<strong>ed</strong> motions in a detector mirror the motions of the source project<strong>ed</strong>onto the plane of the sky.To see why these are true, we define the transverse traceless quadrupole tensorM TTij =⊥ k i ⊥ l j M kl − 1 2 ⊥ ij⊥ kl M kl . (6.27)(Notice that some of our definitions of tracelessness involve subtracting 1 3of thetrace, as in equation (6.24), and sometimes 1 2of the trace, as in equation (6.27).The appropriate factor is determin<strong>ed</strong> by the effective dimensionality (rank) ofthe tensor. Although we have three spatial dimensions, the projection tensor ⊥projects the mass-quadrupole tensor onto a two-dimensional plane, where thetrace involves only two components, not three.)Now, if in equation (6.26) we replace ˜M ij by its definition in terms of M ij ,and then collect terms appropriately, it is not hard to show that the equationsimplifies to its most natural form:¯h TTij = 2 r ¨M TTij . (6.28)This could of course have been deriv<strong>ed</strong> directly by applying the TT operation toequations (6.9)–(6.11). Since this equation involves only the TT part of M, ourfirst assertion above is prov<strong>ed</strong>. According to this equation, in order to calculatethe quadrupole radiation that a particular observer will receive, one ne<strong>ed</strong> onlycompute the mass-quadrupole tensor’s second time-derivative, project it onto theplane of the sky as seen by the observer looking toward the source, take away itstrace, and rescale it by a factor 2/r. In particular, the TT tensor that describes theaction of the wave (as in the polarization diagram in figure 2.1) is a copy of theTT tensor of the mass distribution. This proves our second assertion above.Looking again at figure 2.1 we imagine a detector consisting of two freemasses whose separation is being monitor<strong>ed</strong>. If the wave causes them to oscillaterelative to one another along the x-axis (the ⊕ polarization), this means that thesource motion contain<strong>ed</strong> a component that did the same thing. If the source is abinary, then the binary orbit project<strong>ed</strong> onto the sky must involve motion of thestars back and forth along either the x- or the y-axis.It is possible from this to understand many aspects of quadrupole radiationin a simple way. Consider a binary star system with a circular orbit. Seen by adistant observer in the orbital plane, the project<strong>ed</strong> source motion is linear, backand forth. The receiv<strong>ed</strong> polarization will be linear, the polarization ellipse align<strong>ed</strong>with the orbit. Seen by a distant observer along the axis of the orbit of thebinary, the project<strong>ed</strong> motion is circular, which is a superposition of two linearmotions separat<strong>ed</strong> in phase by 90 ◦ . The receiv<strong>ed</strong> radiation will also have circularpolarization. Because both linear polarizations are present, the amplitude of the


64 Mass- and current-quadrupole radiationwave emitt<strong>ed</strong> up the axis is twice that emitt<strong>ed</strong> in the plane. In this way we cancompletely determine the radiation pattern of a binary system.Notice that, when view<strong>ed</strong> at an arbitrary angle to the axis, the radiationwill be elliptically polariz<strong>ed</strong>, and the degree of ellipticity will directly measurethe inclination of the orbital plane to the line of sight. This is a very specialkind of information, which one cannot normally obtain from electromagneticobservations of binaries. It illustrates the complementarity of the two kinds ofobserving.6.3 Application of the TT gauge to the current-quadrupolefieldNow we turn to the problem of placing next-order terms of the wave field, thecurrent quadrupole and mass octupole, into the TT gauge. Our interest here is tounderstand current-quadrupole radiation in the same physical way as we have justdone for mass-quadrupole radiation. So we shall put the field into the TT gaugeand then see how to separate the current-quadrupole part from the mass-octupole,which we will discard from the present discussion.6.3.1 The field at third order in slow-motionThe next order terms in the non-TT metric bear a simple relationship to the massquadrupoleterms (see equations (6.9)–(6.11)). In each of the metric components,just replace S jk by Ṡ jkl n l to go from one order to the next.This means that we can just skip to the end of the application of the gaugetransformations in equations (6.17) and (6.18) and write the next order of the finalfield, only using S again, not M:¯h TTij = 4 [⊥ ik ⊥ jl Ṡ lkm n m + 1 ]r2 ⊥ij (Ṡ klm n k n l n m − Ṡ k kln l ) , (6.29)or more compactly¯h TTij = 4 r(⊥ ik ⊥ jl ˙˜S klm n m + 1 2 ⊥ij ˙˜S klm n l n k n m ). (6.30)The tilde on S represents a trace-free operation on the first two indices.˙˜S klm = Ṡ lkm − 1 3 δ kl Ṡi im.These are the indices that come from the indices of T jk , so the tensor is symmetricon these. By analogy with the quadrupole calculation, we can also define the TTpart of S ijk by doing the TT projection on the first two indices,S TTijm =⊥k i ⊥ l j S klm − 1 2 ⊥ ij⊥ kl S klm . (6.31)


Application of the TT gauge to the current-quadrupole field 65The TT projection of the equation for the metric ish TTij = 4 r ṠTTijk n k . (6.32)6.3.2 Separating the current quadrupole from the mass octupoleThe last equation is compact, but it does not have the ready interpretation thatwe have at quadrupole order. This is because the moment of the stress, S ijk ,does not have such a clear physical interpretation. We see from equations (6.6)–(6.8) that S ijk is a complicat<strong>ed</strong> mixture of moments of momentum and density.To gain more physical insight into radiation at this order, we ne<strong>ed</strong> to separatethese different contributions. It is straightforward algebra to see that the followingidentity follows from the earlier ones:Ṡ ijk = 1 ...6 M ijk + 2 ¨P [ jk]i 3+ 2 ¨P [ik] j 3, (6.33)where square brackets around indices mean antisymmetrization:A [ik] := 1 2 (Aik − A ki ).This is a complete separation of the mass terms (in M) from the momentum terms(in P) because the only identities relating the momentum moments to the massmoments involve the symmetric part of P ijk on its first two indices, and this isabsent from equation (6.33).The first term in equation (6.33) is the third moment of the density, and thisis the source of the mass-octupole field. It produces radiation through the thirdtime-derivative. Since we are in a slow-motion approximation, this is smaller thanthe mass-quadrupole radiation by typically a factor v/c. Unless there were somevery special symmetry conditions, one would not expect the mass octupole to beanything more than a small correction to the mass quadrupole. For this reason wewill not treat it here.The second and third terms in equation (6.33) involve the second momentof the momentum, and together they are the source of the current-quadrupolefield. It involves two time-derivatives, just as the mass quadrupole does, but theseare time-derivatives of the momentum moment, not the mass moment, so theseterms produce a field that is also v/c smaller than the typical mass-quadrupolefield. However, it requires less of an accident for the mass quadrupole to beabsent and the current quadrupole present. It just requires motions that leave th<strong>ed</strong>ensity unchang<strong>ed</strong> to lowest order. This happens in the r-modes. Therefore, thecurrent quadrupole deserves more attention, and we will work exclusively withthese terms from now on.The terms in equation (6.33) that we ne<strong>ed</strong> are the ones involving ¨P ijk .These are antisymmetriz<strong>ed</strong> on the first two indices, which involves effectivelya vector product between the momentum density (first index) and one of themoment indices. This is essentially the angular momentum density. To make


66 Mass- and current-quadrupole radiationthe angular momentum explicit and to simplify the expression, we introduce theangular momentum and the first moment of the angular momentum densityJ i := ɛ ijk P jk , (6.34)J il := ɛ ijk P l jk , (6.35)where ɛ ijk is the fully antisymmetric (Levi-Civita) symbol in three dimensions. Itfollows from this thatP [ jk]l = 1 2 ɛ jki J l i .These terms enter the TT projection of the field (6.32) with the last indexof S always contract<strong>ed</strong> with the direction n i to the observer from the source.According to equation (6.33), this contraction always occurs on one of theantisymmetriz<strong>ed</strong> indices, or if we use the form in the previous equation then wewill always have a contraction of n i with ɛ ijk . This is a simple object, which wecall⊥ɛ jk := n i ɛ ijk . (6.36)This is just the two-dimensional Levi-Civita object in the plane perpendicular ton i , which is the plane of the sky as seen by the observer. These quantities will beus<strong>ed</strong> in the current-quadrupole field, which contains projections on all the indices.Therefore, the only components of J jk that enter are those project<strong>ed</strong> onto the sky,and so it will simplify formulae to define the sky-project<strong>ed</strong> moment of the angularmomentum ⊥ J⊥ J ij := ⊥ i l ⊥ j m J lm . (6.37)Using this assembl<strong>ed</strong> notation, the current-quadrupole field ish TTij = 4 3r ( ⊥ɛ ik ⊥ ¨J k j + ⊥ ɛ jk ⊥ ¨J k i +⊥ ij ⊥ɛ km ⊥ ¨J km ). (6.38)This is similar in form and complexity to the mass-quadrupole fieldexpression. The interpretation of the contributions is direct. Only componentsof the angular momentum in the plane of the sky contribute to the field. Similarlyonly moments of this angular momentum transverse to the line of sight contribute.If one wants, say, the xx component of the field, then the ⊥ ɛ factor tells us it isdetermin<strong>ed</strong> by the y-component of momentum, i.e. the component perpendicularto the x-direction in the sky. In fact, it is much simpler just to write out theactual components, assuming that the wave travels toward the observer along thez-direction. Then we haveh TTxx = 4 3r ( ¨J xy + ¨J yx ), (6.39)h TTxy = 4 3r ( ¨J yy − ¨J xx ), (6.40)and the remaining components can be found from the usual symmetries of theTT-metric. I have dropp<strong>ed</strong> the prefix ⊥ on J because in this coordinate system thegiven components are already transverse.


Application of the TT gauge to the current-quadrupole field 67−++−Figure 6.1. A simple current-quadrupole radiator. The left-hand panel shows how the twowheels are connect<strong>ed</strong> with blade springs to a central axis. The wheels turn in opposit<strong>ed</strong>irections, each oscillating back and forth about its rest position. The right-hand panelshows the side view of the system, and the arrows indicate the motion of the near side ofthe wheels at the time of viewing. The + signs indicate where the momentum of the massof the wheel is toward the viewer and the − signs indicate where it is away from the viewer.The simplicity of these expressions is striking. There are two basic caseswhere one gets current-quadrupole radiation.• If there is an oscillating angular momentum distribution with a dipolemoment along the angular momentum axis, as project<strong>ed</strong> onto the sky, thenin an appropriate coordinate system ¨J xx will be nonzero and we will have ⊗radiation. To have a non-vanishing dipole moment, the angular momentumdensity could, for example, be symmetrical under reflection through theorigin along its axis, so that it points in opposite directions on opposite sides.• If there is an oscillating angular momentum distribution with a dipolemoment along an axis perpendicular to the angular momentum axis, asproject<strong>ed</strong> onto the sky, then in an appropriate coordiate system ¨J xy will benonzero and we will have ⊕ radiation.6.3.3 A model system radiating current-quadrupole radiationTo see that the first of these two leads to physically sensible results, let us considera simple model system that actually bears a close resemblance to the r-modesystem. Imagine, as in the left panel of figure 6.1, two wheels connect<strong>ed</strong> by anaxis, and the wheels are sprung on the axis in such a way that if a wheel is turn<strong>ed</strong>by some angle and then releas<strong>ed</strong>, it will oscillate back and forth about the axis.Set the two wheels into oscillation with opposite phases, so that when one wheelrotates clockwise, the other rotates anticlockwise, as seen along the axis.Then when view<strong>ed</strong> along the axis, the angular momentum has no componenttransverse to the line of sight, so there is no radiation along the axis. This


68 Mass- and current-quadrupole radiationis sensible, because when project<strong>ed</strong> onto the plane of the sky the two wheelsare performing exactly opposite motions, so the net effect is that there is zeroproject<strong>ed</strong> momentum density.When view<strong>ed</strong> from a direction perpendicular to the axis, with the axis alongthe x-direction, then the angular momentum is transverse, and it has opposit<strong>ed</strong>irection for the two wheels. There is therefore an x-moment of the x-componentof angular momentum, and the radiation field will have the ⊗ orientation.To see that this has a physically sensible interpretation, look back again at thepolarization diagram, figure 2.1, and look at the bottom row of figures illustratingthe ⊗ polarization. See what the particles on the x-axis are doing. They aremoving up and down in the y-direction. What motions in the source could beproducing this?At first one might guess that it is the up-and-down motion of the mass in thewheels as they oscillate, because in fact the near side of each wheel does exactlywhat the test particles at the observer are doing. However, this cannot be theexplanation, because the far side of each wheel is doing the opposite, and whenthey both project onto the sky they cancel. What in fact gives the effect is that atthe top of the wheel the momentum density is first positive (towards the observer)and then negative, while at the bottom of the wheel it is first negative and thenpositive. On the other wheel, the signs are revers<strong>ed</strong>.Current-quadrupole radiation is produc<strong>ed</strong>, at least in simple situations likethe one we illustrate here, by (the second time-derivative derivative of) thecomponent of source momentum along the line of sight. If this is positive in thesense that it is towards the observer, then the momentum density acts as a positivegravitational ‘charge’. If negative, then it is a negative ‘charge’. The wheelshave an array of positive and negative spots that oscillates with time, and thetest particles in the polarization diagram are drawn toward the positive ones andpush<strong>ed</strong> away from the negative ones. Interestingly, in electromagnetism, magneticdipole and magnetic quadrupole radiation are also generat<strong>ed</strong> by the component ofthe electric current along the line of sight.This is a rather simple physical interpretation of some rather more complexequations. It is possible to re-write equation (6.38) to show explicitly thecontribution of the line-of-sight momentum, but the expressions become evenmore complicat<strong>ed</strong>. Instead of dwelling on this, I will turn to the question ofcalculating the total energy radiat<strong>ed</strong> by the source.6.4 Energy radiat<strong>ed</strong> in gravitational <strong>waves</strong>We have calculat<strong>ed</strong> the energy flux in equation (5.14), and we now have the TTwave amplitudes. We ne<strong>ed</strong> only integrate the flux over a distant sphere to getthe total luminosity. We do this for the mass and current quadrupoles in separatesections.


6.4.1 Mass-quadrupole radiationEnergy radiat<strong>ed</strong> in gravitational <strong>waves</strong> 69The mass-quadrupole radiation field in equation (6.26) must be put into the energyflux formula, and the dependence on the direction n i can then be integrat<strong>ed</strong> overa sphere. It is not a difficult calculation, but it does require some angular integralsover multiple products of the vector n i , which depends on the angular directionon the sphere. By symmetry, integrals of odd numbers of factors of n i vanish. Foreven numbers of factors, the result is essentially determin<strong>ed</strong> by the requirementthat after integration the result must be fully symmetric under interchange of anytwo indices and it cannot have any special directions (so it must depend only onthe Kronecker delta δ i j ). The identities we ne<strong>ed</strong> are∫n i n j d = 4π 3 δij , (6.41)∫n i n j n k n l d = 4π15 (δij δ kl + δ ik δ jl + δ il δ jk ). (6.42)Using these, one gets the following simple formula for the total luminosity ofmass-quadrupole radiationL massgw = 1 ... ...5 〈 ˜M jk ˜M jk 〉 . (6.43)Here we still preserve the angle brackets of equation (5.14), because this formulaonly makes sense in general if we average in time over one cycle of the radiation.6.4.2 Current-quadrupole radiationThe energy radiat<strong>ed</strong> in the current quadrupole is nearly as simple to obtain as themass-quadrupole formula. The extra factor of n i in the radiation field makes theangular integrals longer, and requires two further identities:∫n i n j n k n l n p n q d = 4π 7 δ(ij δ kl δ pq) , (6.44)ɛ ijk ɛ i′ j ′ k ′ = δ ii′ δ jj′ δ kk′ + δ ij′ δ jk′ δ ki′ + δ ik′ δ ji′ δ kj′− δ ii′ δ jk′ δ kj′ − δ ij′ δ ji′ δ kk′ − δ ik′ δ jj′ δ ki′ , (6.45)where the round brackets indicate full symmetrization on all indices. Theexpression is simplest if we define˜J jk := ɛ jlm ˜P k lm + ɛ klm ˜P j lm ,where˜P kij := P kij − 1 3 δij P kl l.The result of the integration of the flux formula over a distant sphere is[18, 25], in our notation,L currentgw = 4 5 〈 ...˜J jk ...˜J jk 〉. (6.46)


70 Mass- and current-quadrupole radiation6.5 Radiation in the Newtonian limitThe calculation so far has been within the assumptions of lineariz<strong>ed</strong> theory. Realsources are likely to have significant self-gravity. This means, in particular,that there will be a significant component of the source energy in gravitationalpotential energy, and this must be taken into account.In fact a more realistic equation than equation (6.1) would be£ ¯h αβ =−16π(T αβ + t αβ, ) (6.47)where t αβ is the stress-energy pseudotensor of gravitational <strong>waves</strong>. This is hardto work with: equation (6.47) is an implicit equation because t αβ depends on ¯h αβ .Fortunately, the formulae that we have deriv<strong>ed</strong> are more robust than theyseem. It turns out that the leading order radiation field from a Newtonian sourcehas the same formula as in lineariz<strong>ed</strong> theory. By leading order we mean th<strong>ed</strong>ominant radiation. If there is mass-quadrupole radiation, then the mass-octupoleradiation from a Newtonian source will not be given by the formulae of thelineariz<strong>ed</strong> theory. On the other hand, current-quadrupole and mass-quadrupoleradiation can coexist, because they have different symmetries, so the work wehave done here is generally applicable.More details on how one calculates radiation to higher order in theNewtonian limit will be given in Blanchet’s contribution in this book. This isparticularly important for computing the radiation to be expect<strong>ed</strong> from coalescingbinary systems, whose orbits become highly relativistic just before coalescenceand which are, therefore, not well describ<strong>ed</strong> by lineariz<strong>ed</strong> theory.


Chapter 7Source calculationsNow that we have the formulae for the radiation from a system, we can use themfor some simple examples.7.1 Radiation from a binary systemThe most numerous sources of gravitational <strong>waves</strong> are binary stars systems. Injust half an orbital period, the non-spherical part of the mass distribution returnsto its original configuration, so the angular frequency of the emitt<strong>ed</strong> gravitational<strong>waves</strong> is twice the orbital angular frequency.We shall calculate here the mass-quadrupole moment for two stars of massesm 1 and m 2 , orbiting in the x–y plane in a circular orbit with angular velocity ,govern<strong>ed</strong> by Newtonian dynamics. We take their total separation to be R, whichmeans that the orbital radius of mass m 1 is m 2 R/(m 1 + m 2 ) while that of massm 2 is m 1 R/(m 1 + m 2 ). We place the origin of coordinates at the centre of massof the system. Then, for example, the xx-component of M ij is( )m2 R cos(t) 2 ( )m1 R cos(t) 2M xx = m 1 + m 2m 1 + m 2m 1 + m 2= µR 2 cos 2 (t), (7.1)where µ := m 1 m 2 /(m 1 + m 2 ) is the r<strong>ed</strong>uc<strong>ed</strong> mass. By using a trigonometricidentity and throwing away the part that does not depend on time (since we willuse only time-derivatives of this expression) we haveM xx = 1 2 µR2 cos(2t). (7.2)By the same methods, the other nonzero components areM yy =− 1 2 µR2 cos(2t),M xy = 1 2 µR2 sin(2t).This shows that the radiation will come out at twice the orbital frequency.71


72 Source calculationsIn this case the trace-free moment ˜M ij differs from M ij only by a constant,so we can use these values for M ij to calculate the field and luminosity.As an example of calculating the field, let us compute ¯h TTxx as seen by anobserver at a distance r from the system along the y-axis, i.e. lying in the planeof the orbit. We first ne<strong>ed</strong> the TT part of the mass-quadrupole moment, fromequation (6.27):M TTxx = M xx − 1 2 (M xx + M zz ).However, since M zz = 0, this is just M xx /2. Then from equation (6.28) we find¯h TTxx =−2 µ r (R)2 cos[2(t − r)]. (7.3)Similarly, the result for the luminosity isL gw = 325 µ2 R 4 6 . (7.4)The various factors in these two equations are not independent, because theangular velocity is determin<strong>ed</strong> by the masses and separations of the stars. Whenobserving such a system, we cannot usually measure R directly, but we can infer from the observ<strong>ed</strong> gravitational-wave frequency, and we may often be able tomake a guess at the masses (we will see below that we can actually measure animportant quantity about the masses). So we eliminate R using the Newtonianorbit equationR 3 = m 1 + m 2 2 . (7.5)If in addition we use the gravitational-wave frequency gw = 2, we get¯h TTxx =−2 1/3 Å5/3 gw2/3cos[ gw (t − r)], (7.6)r4L gw =5 × 2 1/3 (Å gw) 103 , (7.7)where we have introduc<strong>ed</strong> the symbol for the chirp mass of the binary system:Å := µ 3/5 (m 1 + m 2 ) 2/5 .Notice that both the field and the luminosity depend only on Å, not on theindividual masses in any other combination.The power represent<strong>ed</strong> by L gw must be suppli<strong>ed</strong> by the orbital energy,E =−m 1 m 2 /2R. By eliminating R as before we find the equationE =− 12 5/3 Å5/3 2/3gw .This is remarkable because it too involves only the chirp mass Å. By setting therate of change of E equal to the (negative of the) luminosity, we find an equationfor the rate of change of the gravitational-wave frequency˙ gw =12 × 21/3Å 5/3 11/3gw . (7.8)5


The r-modes 73As we mention<strong>ed</strong> in chapter 4, since the frequency increases, the signal is said to‘chirp’.These results show that the chirp mass is the only mass associat<strong>ed</strong> with thebinary that can be d<strong>ed</strong>uc<strong>ed</strong> from observations of its gravitational radiation, at leastif only the Newtonian orbit is important. Moreover, if one can measure the fieldamplitude (e.g. h TTxx ) plus gw and ˙ gw , one can d<strong>ed</strong>uce from these the valueof Å and the distance r to the system! A chirping binary with a circular orbit,observ<strong>ed</strong> in gravitational <strong>waves</strong>, is a standard candle: one can infer its distancepurely from the gravitational-wave observations. To do this one ne<strong>ed</strong>s the fullamplitude, not just its projection on a single detector, so one generally ne<strong>ed</strong>s anetwork of detectors or a long-duration observation with a single detector to getenough information.It is very unusual in astronomy to have standard candles, and they are highlypriz<strong>ed</strong>. For example, one can, in principle, use this information to measureHubble’s constant [26].7.1.1 CorrectionsIn the calculation above we made several simplifying assumptions. For example,how good is the assumption that the orbit is circular? The Hulse–Taylor binaryis in a highly eccentric orbit, and this turns out to enhance its gravitational-waveluminosity by more than a factor of ten, since the elliptical orbit brings the twostars much nearer to one another for a period of time than a circular orbit with thesame period would do. So there are big corrections for this system.However, systems emitting at frequencies observable from ground-bas<strong>ed</strong>interferometers are probably well approximat<strong>ed</strong> by circular orbits, because theyhave arriv<strong>ed</strong> at their very close separation by gravitational-wave-driven in-spiral.This process removes eccentricity from the orbit faster than it shrinks the orbitalradius, so by the time they are observ<strong>ed</strong> they have insignificant eccentricity.Another assumption is that the orbit is well describ<strong>ed</strong> by Newtoniantheory. This is not a good assumption in most cases. Post-Newtonian orbitcorrections will be very important in observations. This is not because thestars eventually approach each other closely. It is because they spend a longtime at wide separations where the small post-Newtonian corrections accumulatesystematically, eventually changing the phase of the orbit by an observableamount. So it is very important for observations that we match signals witha template containing high-order post-Newtonian corrections, as describ<strong>ed</strong> inBlanchet’s contribution. But even so, the information contain<strong>ed</strong> in the Newtonianpart of the radiation is still there, so all our conclusions above remain important.7.2 The r-modesWe consider rotating stars in Newtonian gravity and look at the effect that theemission of gravitational radiation has on their oscillations. One might expect


74 Source calculationsthat the loss of energy to gravitational <strong>waves</strong> would damp out any perturbations,and inde<strong>ed</strong> this is normally the case. However, it was a remarkable discovery ofChandrasekhar [27] that the opposite sometimes happens.A rotating star is idealiz<strong>ed</strong> as an axially symmetric perfect-fluid system.In the Newtonian theory the pulsations of a perturb<strong>ed</strong> fluid can be describ<strong>ed</strong>by normal modes which are the solutions of perturb<strong>ed</strong> Euler and gravitationalfield equations. If the star is stable, the eigenfrequencies σ of the normal modesare real; if the star is unstable, there is at least one pair of complex-conjugatefrequencies, one of which represents an exponentially growing mode and the othera decaying mode. (We take the convention that the time-dependence of a mode isexp(iσ t).)In general relativity, the situation is, in principle, the same, except that thereis a boundary condition on the perturbation equations that insists that gravitational<strong>waves</strong> far away be outgoing, i.e. that the star loses energy to gravitational <strong>waves</strong>.This condition forces all eigenfrequencies to be complex. The sign of theimaginary part of the frequency determines stability or instability.The loss of energy to gravitional radiation can destabilize a star that wouldotherwise (i.e. in Newtonian theory) be stable. This is because it opens a pathwayto lower-energy configurations that might not be accessible to the Newtonianstar. This normally happens because gravitational radiation also carries awayangular momentum, a quantity that is conserv<strong>ed</strong> in the Newtonian evolution ofa perturbation.The sign of the angular momentum lost by the star is a critical diagnosticfor the instability. A wave that moves in the positive angular direction around astar will radiate positive angular momentum to infinity. A wave that moves in theopposite direction, as seen by an observer at rest far away, will radiate negativeangular momentum. In a spherical star, both actions result in the damping ofthe perturbation because, for example, the positive-going wave has intrinsicallypositive angular momentum, so when it radiates its angular momentum decreasesand so its amplitude decreases. Similarly, the negative-going wave has negativeangular momentum, so when it radiates negative angular momentum its amplitud<strong>ed</strong>ecreases.The situation can be different in a rotating star, as first point<strong>ed</strong> out byFri<strong>ed</strong>man and Schutz [28]. The angular momentum carri<strong>ed</strong> by a wave dependson its pattern angular velocity relative to the star’s angular velocity, not relativeto an observer far away. If a wave pattern travels backwards relative to the star,it represents a small effective slowing down of the star and therefore carriesnegative angular momentum. This can lead to an anomalous situation: if a wavetravels backwards relative to the star, but forwards relative to an inertial observer(because its angular velocity relative to the star is smaller than the star’s angularvelocity), then it will have negative angular momentum but it will radiate positiveangular momentum. The result will be that its intrinsic angular momentum willget more negative, and its amplitude will grow.This is the mechanism of the Chandrasekar–Fri<strong>ed</strong>man–Schutz (CFS)


The r-modes 75instability. In an ideal star, it is always possible to find pressure-driven <strong>waves</strong>of short enough wavelength around the axis of symmetry (high enough angulareigenvalue m) that satisfy this condition. However, it turns out that even a smallamount of viscosity can damp out the instability in such <strong>waves</strong> , so it is not clearthat pressure-driven <strong>waves</strong> will ever be significantly unstable in realistic stars.However, in 1997 Andersson [17] point<strong>ed</strong> out that there was a class of modescall<strong>ed</strong> r-modes (Rossby modes) that no-one had previously investigat<strong>ed</strong>, andthat were formally unstable in all rotating stars. Rossby <strong>waves</strong> are well knownin oceanography, where they play an important role in energy transport aroundthe Earth’s oceans. They are hard to detect, having long wavelengths and verylow-density perturbations. They are mainly velocity perturbations of the oceans,whose restoring force is the Coriolis effect, and that is their character in neutronstars too. Because they have very small density perturbation, the gravitationalradiation they emit is dominat<strong>ed</strong> by the current-quadrupole radiation.For a slowly-rotating, nearly-spherical Newtonian star, the followingvelocity perturbation is characteristic of r-modes:δv a = ς(r)ɛ abc ∇ b r∇ c Y lm , (7.9)where ς(r) is some function of r determin<strong>ed</strong> by the mode equations. This velocityis a curl, so it is divergence-free; since it has no radial component, it does notchange the density. If the star is perfectly spherical, these perturbations are simplya small rotation of some of the fluid, and it continues to rotate. They have nooscillation, and have zero frequency.If we consider a star with a small rotational angular velocity , then thefrequency σ is no longer exactly zero and a Newtonian calculation to first orderin shows that there is a mode with pattern spe<strong>ed</strong> ω p =−σ/m equal toω p = [1 −2l (l + 1)]. (7.10)These modes are now oscillating currents that move (approximately) along theequipotential surfaces of the rotating star.For l 2, ω p is positive but slower than the spe<strong>ed</strong> of the star, so by the CFSmechanism these modes are unstable to the emission of gravitational radiation foran arbitrarily slowly rotating star.The velocity pattern given in equation (7.9) for (l = 2, m = 2) is closelyrelat<strong>ed</strong> to the wheel model we describ<strong>ed</strong> for current-quadrupole radiation infigure 6.1. Take two such wheels and orient their axes along the x- and y-axes,with the star rotating about the z-axis. Choose the sense of rotation so that thewheels at positive-x and positive-y are spinning in the opposite sense at any time,i.e. so that their adjacent <strong>ed</strong>ges are always moving in the same direction. Thenthis relationship will be reproduc<strong>ed</strong> for all other adjacent pairs of wheels: adjacent<strong>ed</strong>ges move together.When seen from above the equatorial plane, the line-of-sight momenta ofthe wheels reinforce each other, and we get the same kind of pattern that we saw


76 Source calculationswhen looking at one wheel from the side. However, in this case the pattern rotateswith the angular velocity 2/3 of equation (7.10). Since the pattern of line-ofsightmomenta repeats itself every half rotation period, the gravitational <strong>waves</strong>are circularly polariz<strong>ed</strong> with frequency 4/3. Seen along the x-axis, the wheelalong the x-axis contributes nothing, but the other wheel contributes fully, so theradiation amplitude in this direction is half that going out the rotation axis. Seenalong a line at 45 ◦ to the x-axis, the line-of-sight momenta of the wheels on thefront part of the star cancel those at the back, so there is no radiation. Thus,along the equator there is a characteristic series of maxima and zeros, leadingto a standard m = 2 radiation pattern. This pattern also rotates around the star,but the radiation in the equator remains linearly polariz<strong>ed</strong> because there is onlythe ⊗ component, not the ⊕. Again, the radiation frequency is twice the patternspe<strong>ed</strong> because the radiation goes through a complete cycle in half a wave rotationperiod.This discussion cannot go into the depth requir<strong>ed</strong> to understand the r-modesfully. There are many issues of principle: what happens beyond linear order in; what happens if the star is describ<strong>ed</strong> in relativity and not Newtonian gravity;what is the relation between r-modes and the so-call<strong>ed</strong> g-modes that can havesimilar frequencies; what happens when the amplitude grows large enough thatthe evolution is nonlinear; what is the effect of magnetic fields on the evolution ofthe instability? The literature on r-modes is developing rapidly. We have includ<strong>ed</strong>references where some of the most basic issues are discuss<strong>ed</strong> [17, 18, 29–31], butthe interest<strong>ed</strong> student should consult the current literature carefully.7.2.1 Linear growth of the r-modesWe have seen how the r-mode becomes unstable when coupl<strong>ed</strong> to gravitationalradiation, and now we turn to the practical question: is it important? This willdepend on the balance between the growth rate of the mode due to relativisticeffects and the damping due to viscosity.When coupl<strong>ed</strong> to gravitational radiation and viscosity, the mode has acomplex frequency. If we define I(σ ) := 1/τ, then τ is the characteristicdamping time. When radiation and viscosity are treat<strong>ed</strong> as small effects, theircontributions to the eigenfrequencies add, so we have that the total damping isgiven by1τ() = 1 + 1 ,τ GR τ v1τ v= 1 τ s+ 1 τ b, (7.11)where 1/τ GR ,1/τ v are the contributions due to gravitational radiation emissionand viscosity, and where the latter has been further divid<strong>ed</strong> between shearviscosity ( 1 τ s) and bulk viscosity ( 1 τ b).If we consider a ‘typical’ neutron star with a polytropic equation of statep = kρ 2 (for which k has been chosen so that a 1.5M ⊙ model has a radiusR = 12.47 km), and if we express the angular velocity in terms of the scale for


The r-modes 77Table 7.1. <strong>Gravitational</strong> radiation and viscous timescales, in seconds. Negative valuesindicate instability, i.e. a growing rather than damping mode.l m τ gw (s) p gw τ bv (s) p bv τ sv (s)2 2 −20.83 5.93 9.3 × 10 10 1.77 2.25 × 10 83 3 −316.1 7.98 1.89 × 10 10 1.83 3.53 × 10 7the approximate maximum spe<strong>ed</strong> √ πG ¯ρ and the temperature in terms of 10 9 K,then it can be shown that [30]1τ = 1( ) pgwτ 1ms gw +1Pτbv(1msP( ) 610 9 K) pbv+ 1T( ) T 2τ sv 10 9 , (7.12)Kwhere the scaling parameters ˜τ sv , ˜τ bv , ˜τ gw and the exponents p gw and p bv haveto be calculat<strong>ed</strong> numerically. Some representative values relevant to the r-modeswith 2 l 6 are in table 7.1 [30].The physics of the viscosity is interesting. It is clear from equation (7.12) thatgravitational radiation becomes a stronger and stronger destabilizing influenceas the angular velocity of a star increases, but the viscosity is much morecomplicat<strong>ed</strong>. There are two contributions: shear and bulk. Shear viscosity comesmainly from electrons scattering off protons and other electrons. This effect fallswith increasing temperature, just as does viscosity of everyday materials. So acold, slowly rotating star will not have the instability, where a hotter star might.However, at high temperatures, bulk viscosity becomes dominant. This effectarises in neutron stars from nuclear physics. Neutron-star matter always containssome protons and electrons. When it is compress<strong>ed</strong>, some of these react to formneutrons, emitting a neutrino. When it is expand<strong>ed</strong>, some of the neutrons betadecayto protons and electrons, again emitting a neutrino. The emitt<strong>ed</strong> neutrinois not trapp<strong>ed</strong> in the star; within a short time, of the order of a second or less, itescapes. This irreversible loss of energy each time the star is compress<strong>ed</strong> createsa bulk viscosity. Now, bulk viscosity acts only due to the density perturbation,which is small in r-modes. So the effect of bulk viscosity only dominates at veryhigh temperatures.The balance of the viscous and gravitational effects is illustrat<strong>ed</strong> in figure 7.1[30]. This is indicative, but not definitive: much more work is ne<strong>ed</strong><strong>ed</strong> on thephysics of viscosity and the structure of the modes at large values of (small P).7.2.2 Nonlinear evolution of the starOur description so far is only a linear approximation. To understand the fullevolution of the r-modes we have to treat the nonlinear hydrodynamical effects


78 Source calculations,,3 .3, ORJ 7.Figure 7.1. The balance of viscous and gravitational radiation effects in the r-modes isillustrat<strong>ed</strong> in a diagram of rotation spe<strong>ed</strong>, showing the ratio of the maximum period P kto the rotation period P versus the temperature of the star. The solid curve indicates theboundary between viscosity-dominat<strong>ed</strong> and radiation-dominat<strong>ed</strong> behaviour: stars abovethe line are unstable. The dash<strong>ed</strong> curves illustrate possible nonlinear evolution histories asa young neutron star cools.that become important as the modes grow. This could only be done with anumerical simulation, which some groups are now working on. However, it ispossible to make simple estimates analytically.We characterize the initial configuration with just two paramters: the uniformangular velocity , and the amplitude α of the r-modes perturbation. The star isassum<strong>ed</strong> to cool at the accept<strong>ed</strong> cooling rate for neutron stars, independently ofwhether it is affect<strong>ed</strong> by the r-mode instability or not. The star is assum<strong>ed</strong> tolose angular momentum to gravitational radiation at a rate given by the linearradiation field, with its large amplitude α. This loss is taken to drive the starthrough a sequence of equilibrium states of lower and lower angular momentum.Details of this approximation are in [31], here we report only the results. The


The r-modes 79evolution turns out to have three phases.(a) Initially the angular velocity of the hot rapidly rotating neutron star isnearly constant, evolving on the viscous timescale 1/τ v , while the amplitudeα grows exponentially on the gravitational radiation timescale 1/τ GR .(b) After a short time nonlinear effects become important and stop the growth ofthe amplitude α. Most of the initial angular momentum of the star is radiat<strong>ed</strong>away by gravitational radiation. The star spins down and evolves to a pointwhere the angular velocity and the temperature is sufficiently low that ther-mode is stable.(c) Finally gravitational radiation and viscosity damp out the r-mode and drivethe star into its final equilibrium configuration.This may take about a year, a timescale govern<strong>ed</strong> by the cooling time of thestar. During this year, the star would radiate away most of its angular momentumand rotational kinetic energy. This could be a substantial fraction of a solar massin energy.7.2.3 Detection of r-mode radiationThe large amount of energy radiat<strong>ed</strong> into the r-modes makes them attractive fordetection, but detection will not be trivial. The r-mode event occurs at the rateof supernovae: some fraction (hopefully large) of all supernovae leave behind arapidly spinning neutron star that spins down over a one-year period. This meanswe should have sufficient sensitivity to reach the Virgo Cluster (20 Mpc distance).Estimates [31] suggest that a neutron star in the Virgo Cluster could be detect<strong>ed</strong>by second generation of LIGO and VIRGO gravitational-wave detectors with anamplitude signal-to-noise of about eight, provid<strong>ed</strong> one can use match<strong>ed</strong> filtering(exact template matching).It will not be easy to use match<strong>ed</strong> filtering, since one must follow all cyclesof the signal as the star spins down, and we will not know this well because ofmany uncertainties: initial temperature, initial spin distribution, detail<strong>ed</strong> physicsof viscosity, and so on. However, it would be helpful to have a parametriz<strong>ed</strong>model to take account of the uncertainties, so that we could look for a significantfit to one or more of the parameters.In addition, it is likely that, if a significant proportion of all neutron starswent through the r-mode instability, then the universe has been fill<strong>ed</strong> by theirradiation. There should be a background with an energy density gw that is agood fraction of the closure density. Its lower frequency limit should be around200 Hz in the rest frame of the star. When we see radiation cosmologically, itslower frquency limit will indicate the epoch at which star formation began.It is clear that the discovery of this new source of gravitational <strong>waves</strong> willopen several prospects for astronomy. Observations could be us<strong>ed</strong> as supernova<strong>ed</strong>etectors, revealing supernovae hidden in clouds of dust, identifying them about ayear after they are form<strong>ed</strong>. The existence of the radiation raises several prospects


80 Source calculationsand questions about the physics of neutron stars, not least the interaction ofmagnetic fields with the instability.7.3 ConclusionThese lectures (chapters 2 to 7) have taken us through the basic theory ofgravitational radiation and its applications in astrophysics, so far as we canunderstand and pr<strong>ed</strong>ict them now. In a few years, perhaps as little as two, perhapsas many as eight, we will start to make observations of gravitational radiation fromastrophysical sources. If gravitational-wave astronomy follows other branches ofobservational astronomy, it will not be long before completely unexpect<strong>ed</strong> signalsare seen, or unexpect<strong>ed</strong> features in long-pr<strong>ed</strong>ict<strong>ed</strong> signals. To interpret these willrequire joining a physical understanding of the relationship between gravitationalradiation and its source to a wide knowl<strong>ed</strong>ge of astronomical phenomena. Weencourage the students who have attend<strong>ed</strong> these lectures, and others who maystudy them, to get themselves ready to contribute to this activity. It will be anexciting time!


ReferencesWe have divid<strong>ed</strong> the references into two sections. The first gives generally usefulreferences—books, conference summaries, etc—that interest<strong>ed</strong> students shouldgo to for background and a more complete discussion of the theory. The secondsection contains specific references to the research literature. General referencesare indicat<strong>ed</strong> in the text by the author name plus year, as Misner et al (1973). Thespecific references are indicat<strong>ed</strong> by numbers, for example [1,3].General referencesGeneral relativityThere are a number of good text books on GR. The following cover, at differentlevels of difficulty and completeness, lineariz<strong>ed</strong> theory, gauges and the definitionof energy:Ciufolini I and Wheeler J L 1995 Gravitation and Inertia (Princeton, NJ: PrincetonUniversity Press)Landau L and Lifshitz E M 1962 The Classical Theory of Fields (New York: Pergamon)Misner C W, Thorne K S and Wheeler J L 1973 Gravitation (San Francisco, CA: Freeman)Ruffini R and Ohanian H C 1997 Gravitazione e Spazio-Tempo (Bologna: Zanichelli)Schutz B F 1995 A First Course in General Relativity (Cambridge: Cambridge UniversityPress)Wald R M 1994 General Relativity (Chicago, IL: Chicago University Press)Weinberg S 1972 Gravitation and Cosmology (New York: Wiley)<strong>Gravitational</strong>-wave detectorsConference volumes on detector progress appear more than once per year thes<strong>ed</strong>ays. You can find progress reports on detectors on the web sites of the differentgroups, which you will find in the list of literature references below. The tworeferences below are more tutorial, aim<strong>ed</strong> at introducing the subject.Blair D G 1991 The Detection of <strong>Gravitational</strong> Waves (Cambridge: Cambridge UniversityPress)Saulson P R 1994 Fundamentals of Interferometric <strong>Gravitational</strong> Wave Detectors(Singapore: World Scientific)81


82 ReferencesSources of gravitational <strong>waves</strong>Again, there have been a number of conference publications on this subject. Thethird and fourth references are recent reviews of sources. The first two survey theproblem of data analysis.Schutz B F (<strong>ed</strong>) 1989 <strong>Gravitational</strong> Wave Data Analysis (Dordrecht: Kluwer)——1997 The detection of gravitational <strong>waves</strong> Relativistic Gravitation and <strong>Gravitational</strong>Radiation <strong>ed</strong> J A Marck and J P Lasota (Cambridge: Cambridge University Press)pp 447–75Thorne K S 1987 <strong>Gravitational</strong> radiation 300 Years of Gravitation <strong>ed</strong> S W Hawking andW Israel (Cambridge: Cambridge University Press) pp 330–458——1995 Proc. 1994 Summer Study on Particle and Nuclear Astrophysics and Cosmology<strong>ed</strong> E W Kolb and R Peccei (Singapore: World Scientific)Text references[1] Schutz B F 1984 <strong>Gravitational</strong> <strong>waves</strong> on the back of an envelope Am. J. Phys. 52412–19[2] Nicholson D et al 1996 Results of the first coincident observations by 2 laserinterferometricgravitational-wave detectors Phys. Lett. A 218 175–80[3] Meers B J 1988 Recycling in laser-interferometric gravitational wave detectors Phys.Rev. D 38 2317–26[4] TAMA300 project website, http://tamago.mtk.nao.ac.jp/[5] GEO600 project website, http://www.geo600.uni-hannover.de/[6] LIGO project website, http://ligo.caltech.<strong>ed</strong>u/[7] VIRGO project website, http://virgo4p.pg.infn.it/virgo/[8] Caves C M et al 1980 Rev. Mod. Phys. 52 341–92[9] Astone P, Lobo J A and Schutz B F 1994 Coincidence experiments betweeninterferometric and resonant bar detectors of gravitational <strong>waves</strong> Class. QuantumGrav. 11 2093–112[10] Rome gravitational wave group website, http://www.roma1.infn.it/rog/rogmain.html[11] Auriga detector website, http://axln01.lnl.infn.it/[12] LISA project websites, http://www.lisa.uni-hannover.de/ and http://lisa.jpl.nasa.gov/[13] Bender P et al LISA: Pre-Phase A Report MPQ 208 (Max-Planck-Institut fürQuantenoptik, Garching, Germany) (also see the 2nd <strong>ed</strong>n, July 1998)[14] Lipunov V M, Postnov K A and Prokhorov M E 1997 Formation and coalescence ofrelativistic binary stars: the effect of kick velocity Mon. Not. R. Astron. Soc. 288245–59[15] Yungelson L R and Swart S F P 1998 Formation and evolution of binary neutron starsAstron. Astrophys. 332 173–88[16] ZwartSFPandMcMillain S L W 2000 Black hole mergers in the universe Astrophys.J. 528 L17–20 (gr-qc/9910061)[17] Andersson N 1998 A new class of unstable modes of rotating relativistic starsAstrophys. J. 502 708–13 (gr-qc/9706075)[18] Lindblom L, Owen B J and Morsink S M 1998 <strong>Gravitational</strong> radiation instability inhot young neutron stars Phys. Rev. Lett. 80 4843–6 (gr-qc 9803053)


References 83[19] van der Klis M 1998 The Many Faces of Neutron Stars <strong>ed</strong> R Buccheri, J van Paradijsand M A Alpar (Dordrecht: Kluwer Academic)[20] Bildsten L 1998 <strong>Gravitational</strong> radiation and rotation of accreting neutron starsAstrophys. J. 501 L89–93 (astro-ph/9804325)See also the recent paper Ushomirsky G, Cutler C and Bildsten L 1999 Deformationsof accreting neutron star crusts and gravitational wave emission Preprint (astroph/0001136)[21] Brady P R, Creighton T, Cutler C and Schutz B F 1998 Searching for periodic sourceswith LIGO Phys. Rev. D 57 2101–16[22] Brustein R, Gasperini M, Giovannini M and Veneziano G 1996 <strong>Gravitational</strong>radiation from string cosmology Proc. Int. Europhysics Conf. (HEP 95) <strong>ed</strong>CVander Velde, F Verbeure and J Lemonne (Singapore: World Scientific)[23] Schutz B F and Sorkin R 1977 Variational aspects of relativistic field theories, withapplication to perfect fluids Ann. Phys., NY 107 1–43[24] Isaacson R Phys. Rev. 166 1263–71Isaacson R 1968 Phys. Rev. 166 1272–80[25] Thorne K S 1980 Multipole expansions of gravitational radiation Rev. Mod. Phys. 52299–340[26] Schutz B F 1986 Determining the Hubble constant from gravitational waveobservations Nature 323 310–11[27] Chandrasekar S 1970 Phys. Rev. Lett. 24 611[28] Fri<strong>ed</strong>man J L and Schutz B F 1978 Ap. J. 222 281[29] Fri<strong>ed</strong>man J L and Morsink S M 1998 Axial instability of rotating relativistic starsAstrophys. J. 502 714–20 (gr-qc/9706073)[30] Andersson N, Kokkotas K and Schutz B F 1999 <strong>Gravitational</strong> radiation limit on thespin of young neutron stars Astrophys. J. 510 846–8539 (astro-ph/9805225)[31] Owen B, Lindblom L, Cutler C, Schutz B F, Vecchio A and Andersson N 1998<strong>Gravitational</strong> <strong>waves</strong> from hot young rapidly rotating neutron stars Phys. Rev. D5808 4020 (gr-qc/9804044)


Solutions to exercisesChapter 2Exercise (a)1. Let us take the form of the wave to beh TT jk = e jk⊕ h +(t − ˆn · ˆx)where e jk⊕ is the polarization tensor for the ⊕ polarization, and where ˆn is theunit vector in the direction of travel of the wave. We will let h + be an arbitraryfunction of its phase argument.If the wave travels in the x–z plane at an angle θ to the z-direction, then theunit vector in our coordinates isˆn i = (sin θ,0, cos θ).We ne<strong>ed</strong> to calculate the polarization tensor’s components in x, y, z coordinates.We do this by rotating the ⊕ polarization tensor from its TT form in coordinatesparallel to the wavefront to its form in our coordinates. This requires a simplerotation around the y-axis. The transformation matrix is:( cos θ 0) sin θ j ′ k = 0 1 0 .− sin θ 0 cos θThe polarization tensor in our coordinates (prim<strong>ed</strong> indices) becomes:e j ′ k ′ = j ′ l k′ m e lm(cos 2 )θ 0 − sin θ cos θ= 0 −1 0 .− sin θ cos θ 0 sin 2 θNotice that the new polarization tensor is again traceless.The gravitational wave will be, at an arbitrary time t and position (x, z) inour (x, z)-plane,84h TT j ′ k ′ = e j ′ k ′ h + (t − x sin θ − z cos θ).


Solutions to exercises 85For this problem we ne<strong>ed</strong> the xx-component because the photon is propagatingalong this direction, and we will always stay at z = 0, so we haveh TTxx = cos 2 θh + (t − x sin θ).We see that for this geometry the wave amplitude is r<strong>ed</strong>uc<strong>ed</strong> by a factor of cos 2 θ.Generalizing the argument in the text, the relation between time and positionfor the photon on its trip outwards along the x-direction is t = t 0 + x, wheret 0 is the starting time. The analogous relation after the photon is reflect<strong>ed</strong> ist = t 0 + L + (L − x), since in this case x decreases in time from L to 0. If weput these into the equation for the lineariz<strong>ed</strong> corrections to the return time, we get{ ∫ Lt return = t 0 + 2L + 1 2 cos2 θ h + [t 0 + (1 − sin θ)x]dx0∫ L}× h + [t 0 + 2L − (1 + sin θ)x]dx .0This expression must be differentiat<strong>ed</strong> with respect to t 0 to find the variationof the return time as a function of the start time. The key point is how tohandle differentiation within the integrals. Consider, for example, the functionh + [t 0 + (1 − sin θ)x]. It is a function of a single argument,ξ := t 0 + (1 − sin θ)xso derivatives with respect to t 0 can be convert<strong>ed</strong> to derivatives with respect to xas followsdh += dh + dξ= dh +dt 0 dξ dt 0 dξ ;dh +dx = dh + dξdξ dx = (1 − sin θ)dh +dξ .It follows thatdh += dh +/(1 − sin θ).dt 0 dxOn the return trip the factor will be −(1 + sin θ). So when we differentiate wecan convert the derivatives with respect to t 0 inside the integrals into derivativeswith respect to x. Taking account of the factor cos 2 θ = (1 − sin θ)(1 + sin θ) infront of the integrals, the result isdt returndt 0= 1 + 1 2 (1 + sin θ) ∫ L+ 1 2 (1 − sin θ) ∫ L00dh +dx [t 0 + (1 − sin θ)x]dxdh +dx [t 0 + 2L − (1 + sin θ)x]dx.


86 Solutions to exercisesThe integrals can now be done, since they simply invert the differentiation by x.Evaluating the integrands at the end points of the integrals gives equation (2.20):dt return= 1 − 1dt2 (1 + sin θ)h +(t 0 ) + sin θh + [t 0 + (1 − sin θ)L]0+ 1 2 (1 − sin θ)h +(t 0 + 2L).2. If we Taylor-expand this equation in powers of L about L = 0, the leadingterm vanishes, and the first-order term is:dt return= L sin θ(1 − sin θ)ḣ + (t 0 ) + L(1 − sin θ)ḣ + (t 0 ),dt 0= L cos 2 θḣ + (t 0 ).This is just what was requir<strong>ed</strong>. The factor of cos 2 θ comes, as we saw above, fromthe projection of the TT field on the x-coordinate direction.3. All the terms cancel and there is no effect on the return time.Exercise (b)This is part of the calculation in the previous example. All we ne<strong>ed</strong> is the segmentwhere the light travels from the distant end to the centre:and so dt out /dt 0 is:t out = t 0 + 1 2 cos2 θ∫ L0h + [t 0 + L − (1 + sin θ)x]dxdt outdt 0= 1 + 1 2 (1 − sin θ)[h +(t 0 − sin θ L) − h + (t 0 + L)].Exercise (c)This question is frequently ask<strong>ed</strong>, but not by people who have done thecalculation. The answer is that the two effects occur in different gauges, notin the same one. So they cannot cancel. The apparent spe<strong>ed</strong> of light changes inthe TT gauge, but then the positions of the ends remain fix<strong>ed</strong>, so that the effect isall in the coordinate spe<strong>ed</strong>. In a local Lorentz frame ti<strong>ed</strong> to one mass, the ends domove back and forth, but then the spe<strong>ed</strong> of light is invariant.Exercise (d)To first order we haveR µ αβν = Ɣµ αν,β − Ɣµ αβ,ν ,Ɣ µ αβ,ν = 1 2 ηµσ (h σβ,αν + h σα,βν − h αβ,σ ν ), (i) (7.13)Ɣ µ αν,β = 1 2 ηµσ (h σν,αβ + h ασ,νβ − h αν,σβ ). (ii) (7.14)


Solutions to exercises 87The gauge transformation for a perturbation in lineariz<strong>ed</strong> theory isSubstituting (iii) into (i) and (ii), we obtainh ′ αβ = h αβ − ξ α,β − ξ β,α . (iii) (7.15)(i) = 1 2 ηµσ (h ′ σβ,αν + ξ σ,βαν + h ′ σα,βν + ξ σ,αβν − h ′ αβ,σ ν )(ii) = 1 2 ηµσ (h ′ σν,αβ + ξ σ,ναβ + h ′ σα,βν + ξ σ,αβν − h ′ αν,σβ ).If we find the difference between the two formulae above we getChapter 5Exercise (e)R µ αβν= (ii) − (i) = Ɣ′µαν,β − Ɣ′µ αβ,ν = R′µ αβν .The action principle is:∫ √ ∫δ(R −g)δI =h µν d 4 x =−δg µνG µν√ −gh µν d 4 x = 0. (i) (7.16)If we perform an infinitesimal coordinate transformation x µ → x µ + ξ µ withoutotherwise varying the metric, then the action I must not change:∫0 = δI = G µν (ξ µ;ν + ξ ν;µ ) √ −g d 4 x∫= 2 G µν ξ µ;ν d 4 x.This can be transform<strong>ed</strong> in the following way:∫∫δI = (G µν √ξ µ ) ;ν −g d 4 x − (G µν ;νξ µ ) √ −g d 4 x = 0.The first integral is a divergence and vanishes.arbitrariness of ξ µ , gives the Bianchi’s identities:G µν ;ν = 0.The second, because of theExercise (f)The two polarization components are h + xx yy=−h+ = A +e −ik(t−z) and h xy× =A × e −ik(t−z) . The energy flux is the negative of〈T (GW)0z〉= 132π 〈hij ,0h ij,z 〉=− k216π (A2 + + A2 × )〈sin2 k(t − z)〉=− k232π (A2 + + A2 × ).


PART 2GRAVITATIONAL-WAVE DETECTORSGuido Pizzella, Angela Di Virgilio, Peter Bender andFrancesco Fucito


Chapter 8Resonant detectors for gravitational <strong>waves</strong>and their bandwidthG PizzellaUniversity of Rome Tor Vergata and Laboratori Nazionali diFrascati, PO Box 13, I-00044 Frascati, ItalyThe sensitivity of the resonant detectors for gravitational <strong>waves</strong> (GW) isdiscuss<strong>ed</strong>. It is shown that with a very low-noise electronic amplifier it ispossible to obtain a frequency bandwidth up to 50 Hz. Five resonant detectorsare presently in operation, with a spectral amplitude sensitivity of the order of˜h = 3 × 10 −22 (1/ √ Hz). Initial results are present<strong>ed</strong>, including the first searchfor coincidences and a measurement of the GW stochastic background.8.1 Sensitivity and bandwidth of resonant detectorsThe detectors of GW now operating [1–5] use resonant transducers (and thereforethere are two resonance modes coupl<strong>ed</strong> to the gravitational field) in order to obtainhigh coupling and high-Q.However, for a discussion on the detectors’ sensitivity and frequencybandwidth it is sufficient to consider the simplest resonant antenna, a cylinderof high-Q material, strongly coupl<strong>ed</strong> to a non-resonant transducer follow<strong>ed</strong> by avery low-noise electronic amplifier. The equation for the end bar displacement ξis¨ξ + 2β 1 ˙ξ + ω0 2 ξ = f (8.1)mwhere f is the appli<strong>ed</strong> force, m the oscillator r<strong>ed</strong>uc<strong>ed</strong> mass (for a cylinderm = M/2) and β 1 = ω 0 /2Q is the inverse of the decay time of an oscillationdue to a delta excitation.We consider here only the noise which can be easily modell<strong>ed</strong>, the sum oftwo terms: the thermal (Brownian) noise and the electronic noise. The power91


92 Resonant detectors for gravitational <strong>waves</strong> and their bandwidthspectrum due to the thermal noise isS f = 2ω 0Q mkT e (8.2)where T e is the equivalent temperature which includes the effect of the backactionfrom the electronic amplifier.By referring the noise to the displacement of the bar ends, we obtain thepower spectrum of the displacement due to Brownian noise:S B ξ= S fm 2 1(ω 2 − ω 2 0 )2 + ω2 ω 2 0Q 2 . (8.3)From this we can calculate the mean square displacement¯ ξ 2 = kT emω 2 0(8.4)that can also be obtain<strong>ed</strong>, as is well known, from the equipartition of the energy.To this noise we must add the wide-band noise due to the electronic amplifier(the contribution to the narrow-band noise due to the amplifier has already beeninclud<strong>ed</strong> in T e ).For the sake of simplicity we consider an electromechanical transducer thatconverts the displacement of the detector in a voltage signalV = αξ (8.5)with the transducer constant α (typically of the order of 10 7 Vm −1 ). Thus, theelectronic wide-band power spectrum, S 0 , is express<strong>ed</strong> in units of V 2 Hz −1 andthe overall noise power spectrum referr<strong>ed</strong> to the bar end is given byS n ξ = 2kT eω 0mQ1+ S 0(ω 2 − ω0 2)2 + ω2 ω02 α 2 . (8.6)Q 2We now calculate the signal due to a gravitational wave with amplitude h andwith optimum polarization impinging perpendicularly to the bar axis. The bardisplacement corresponds [6] to the action of a forcef = 2 mLḧ. (8.7)π 2The bar end spectral displacement due to a flat spectrum of GW (as for a deltaexcitation)is similar to that due to the action of the Brownian force. Therefore, ifonly the Brownian noise were present, we would have a nearly infinite bandwidth,


Sensitivity and bandwidth of resonant detectors 93in terms of the signal-to-noise ratio (SNR). For a GW excitation with powerspectrum S h (ω), the spectrum of the corresponding bar end displacement isWe can then write the SNR asS ξ = 4L2 ω 4 S hπ 4 1(ω 2 − ω 2 0 )2 + ω2 ω 2 0Q 2 . (8.8)SNR = S ξS n ξ= 4L2 ω 4 S h1π 4 S f1 + Ɣ(Q 2 (1 − ω2m 2ω 2 0) 2 + ω2 )ω02(8.9)where the quantity Ɣ is defin<strong>ed</strong> by [7]Ɣ = S 0β 1α 2 ¯ξ 2 ∼ T n(8.10)β QT eT n is the noise temperature of the electronic amplifier and β indicates the fractionof energy which is transferr<strong>ed</strong> from the bar to the transducer. It can readily beseen that Ɣ ≪ 1.The GW spectrum that can be detect<strong>ed</strong> with SNR = 1 is:⎛ ⎛ ( ) ⎞⎞S h (ω) = π 2 kT e ω03 21⎝1MQv 2 ω 3 + Ɣ ⎝Q 2 1 − ω2ωω02 + ω2 ⎠⎠ ω02 (8.11)where v is the sound velocity in the bar material (v = 5400 m s −1 in aluminum).For ω = ω 0 we obtain the highest sensitivityS h (ω 0 ) = π 2 kT e 1MQv 2 (8.12)ω 0having consider<strong>ed</strong> Ɣ ≪ 1.Another useful quantity often us<strong>ed</strong> is the spectral amplitude˜h = √ S h . (8.13)We remark that the best spectral sensitivity, obtain<strong>ed</strong> at the resonancefrequency of the detector, only depends, according to equation (8.12), on thetemperature T , on mass M and on the quality factor Q of the detector, provid<strong>ed</strong>T = T e . Note that this condition is rather different from that requir<strong>ed</strong> for optimumpulse sensitivity (see later). The bandwidth of the detector is found by imposingthat S h (ω) is equal to twice the value of S h (ω 0 ). We obtain, in terms of thefrequency f = ω/2π, that△ f = f 0Q1√Ɣ. (8.14)


94 Resonant detectors for gravitational <strong>waves</strong> and their bandwidthFigure 8.1. Spectral amplitudes ˜h present and plann<strong>ed</strong> (see equations (8.13) and (8.11))using the parameters given in table 8.1 versus frequency (Hz). The presently operatingdetectors have bandwidth ∼1 Hz. The bandwidth will be much larger in the near futurewith improv<strong>ed</strong> transducers.Table 8.1. Bandwith and sensitivity for presently operating detectors and for futur<strong>ed</strong>etectors with improv<strong>ed</strong> transducers.T △ f(˜h min )(K) Ɣ Q (Hz) √ 1Hz0.1 10 −6 8.5 × 10 5 1.1 5.3 × 10 −220.1 10 −11 4.2 × 10 6 70 2.3 × 10 −22The present detector bandwidths are of the order of 0.5 Hz, but it is expect<strong>ed</strong> thatthe bandwidths will become of the order of 50 Hz, by improving the amplifiernoise temperature T n , the coupling parameter β and the quality factor Q.In figure 8.1 we show the spectral amplitude ˜h for the present aluminiumresonant detectors with mass M = 2270 kg operating at temperature T = 0.1 Kand the target ˜h plann<strong>ed</strong> to be reach<strong>ed</strong> with improv<strong>ed</strong> transducers. The parametersus<strong>ed</strong> for calculating the spectral amplitudes are given in table 8.1.


8.2 Sensitivity for various GW signalsSensitivity for various GW signals 95Let us consider a signal s(t) in the presence of noise n(t) [8]. The availableinformation is the sumx(t) = s(t) + n(t) (8.15)where x(t) is the measurement at the output of the low-noise amplifier and n(t) isa random process with known properties. Let us start by applying to x(t) a linearfilter which must be such to maximize the SNR at a given time t 0 (we emphasizethe fact that we search the signal at a given time t 0 ).Indicating the impulse response of the filter as w(t) (to be determin<strong>ed</strong>) andwith y s (t) = s(t) ∗ w(t) and y n (t) = n(t) ∗ w(t), respectively, the convolutionsof the signal and of the noise, we haveSNR = |y s(t 0 )| 2E[|y n (t 0 )| 2 ] . (8.16)The expectation of the noise after the filter, indicating with N(ω) the powerspectrum of the noise n(t), isE[|y n (t 0 )| 2 ] = 1 ∫ ∞N(ω)|W(ω)| 2 dω (8.17)2π −∞where W(ω) is the Fourier transform of the unknown w(t).At t = t 0 the output due to s(t) with Fourier transform S(ω) is given byy s (t 0 ) = 1 ∫ ∞S(ω)W(ω)e jωt 0dω. (8.18)2π −∞We now apply [8] the Schwartz’ inequality to the integral (8.18) and usingthe identityS(ω)W(ω) =S(ω) √ N(ω)W(ω) √ N(ω) (8.19)we obtainSNR ≤ 1 ∫ ∞|S(ω)| 2d f. (8.20)2π −∞ N(ω)It can be shown [8] that the equals sign holds if and only ifW(ω) = constant S(ω)∗N(ω) e− jωt 0. (8.21)Applying this optimum filter to the data we obtain the maximum SNRSNR = 1 ∫ ∞|S(ω)| 2dω (8.22)2π −∞ N(ω)


96 Resonant detectors for gravitational <strong>waves</strong> and their bandwidthwhere S(ω) and N(ω) as already specifi<strong>ed</strong> are, respectively, the Fourier transformof the signal and the power spectrum of the noise at the end of the electronic chainwhere the measurement x(t) is taken.Let us apply the above result to the case of measurements x(t) done at theend of a chain of two filters with transfer functions W a (representing the bar) andW e (representing the electronics).Let S uu be the white spectrum of the Brownian noise entering the bar andS ee the white spectrum of the electronics noise. The total noise power spectrumisN(ω) = S uu |W a | 2 |W e | 2 + S ee |W e | 2 . (8.23)The Fourier transform of the signal isS(ω) = S g (ω)W a W e . (8.24)where S g is the Fourier transform of the GW signal at the bar entrance.The optimum filter will have, applying equation (8.21), the transfer functionW(ω) = S∗ g e− jωt 0S uu1W a W e11 + Ɣ|W a | 2 (8.25)whereƔ = S ee. (8.26)S uuUsing equations (8.23) and (8.24), from equation (8.22) we obtainSNR = 12π S uu∫ ∞−∞|S g (ω)| 2 dω1 + Ɣ|W a | 2 . (8.27)We now apply the above result to a delta GW with Fourier transform S gindependent of ω. The remaining integral of equation (8.27) can be easily solv<strong>ed</strong>if we make use of a lock-in device which translates the frequency, bringing theresonant frequency ω 0 to zero. Then the total noise becomes [7]with the bar transfer function given byN(ω) ⇒ N(ω − ω 0 ) + N(ω + ω 0 ) (8.28)W a = β 1β 1 + iω . (8.29)We now estimate the signal and noise just after the first transfer function, beforethe electronic wide-band noise. For the signal due to a delta excitation we haveV (t) = 12π∫ ∞−∞β 1β 1 + iω S g dω = S g β 1 e −β 1t = V s e −β 1t(8.30)


Sensitivity for various GW signals 97where V s is the maximum signal. For the thermal noise we haveVnb 2 = 1 ∫ ∞β122π −∞ β1 2 + S ω2 uu dω = S uuβ 12(8.31)where Vnb 2 = α2 kT emω02(8.5)) .Introducing the signal energy E svariable y =β ω 1)SNR = 1 Sg 2β ∫1 ∞2 2π S uuis the mean square narrow-band noise (see equations (8.4) and−∞= 1 2 mω2 0 ( V sα )2 we calculate (using th<strong>ed</strong>y1 + Ɣ(1 + y 2 ) = V s2√ =Ɣ8V 2 nbE s4kT e√Ɣ. (8.32)The factor of 1 2has been introduc<strong>ed</strong> because we have suppos<strong>ed</strong> the signal to beall at a given phase, while we add to the noise in phase the noise in quadrature,thus r<strong>ed</strong>ucing the SNR by a factor of two. The effective noise temperature is [9]T eff = 4T e√Ɣ. (8.33)For a continuous source we can directly apply equation (8.22). With a totalmeasuring time t m the continuous source with amplitude h 0 and angular frequencyω 0 appears as a wavepacket with its Fourier transform at ω 0S(ω 0 ) =( )h0 t 2 m(8.34)2and with bandwidthδ f = 2(8.35)t mwhich is very small for long observation times. Indicating with N(ω 0 ) the powerspectrum of the measur<strong>ed</strong> noise at the resonance, we obtain the amplitude of thewave that can be observ<strong>ed</strong> with SNR = 1:h 0 =√2N(ω 0 )t m. (8.36)A similar result has been obtain<strong>ed</strong> in the past [10] using a different proc<strong>ed</strong>ure.In practical cases it is often not possible to calculate the Fourier spectrumN(ω 0 ) from experimental data over the entire period of measurement t m , eitherbecause the number of steps in the spectrum would be too large for a computeror because the physical conditions change as, for instance, a change in frequencydue to the Doppler effect. It is then necessary to divide the period t m in n subperiodsof length △t = t mn. In the search for a monochromatic wave we then haveto consider two cases: (a) The wave frequency is exactly known. In this case we


98 Resonant detectors for gravitational <strong>waves</strong> and their bandwidthcan combine n Fourier spectra in one unique spectrum taking into considerationalso the phase of the signal. The final spectrum has the same characteristics ofthe spectrum over the entire period t m and equation (8.36) still applies. (b) Theexact frequency is unknown. In this case when we combine the n spectra we loseinformation on the phase. The result is that the final combin<strong>ed</strong> spectrum over theentire period has a larger variance and the left part of equation (8.36) has to bechang<strong>ed</strong> to√2N(ω 0 )h = √tm △t . (8.37)We come now to the measurement of the GW stochastic background [11,12,16]. Using one detector, the measurement of the noise spectrum corresponding toequation (8.11) only provides an upper limit for the GW stochastic backgroundspectrum, since the noise is not so well known that we can subtract it. Theestimation of the GW stochastic background spectrum can be considerablyimprov<strong>ed</strong> by employing two (or more) antennae, whose output signals are crosscorrelat<strong>ed</strong>. Let us consider two antennae, that may in general be different, withtransfer functions T 1 and T 2 , and displacements ξ 1 and ξ 2 : the displacement crosscorrelationfunction∫R ξ1 ξ 2(τ) = ξ 1 (t + τ)ξ 2 (t) dt (8.38)only depends on the common excitation of the detectors, due to the GW stochasticbackground spectrum S gw acting on both of them, and is not affect<strong>ed</strong> by the noisesacting independently on the two detectors.The Fourier transform of equation (8.38) is the displacement cross spectrum.This spectrum, multipli<strong>ed</strong> by T 1 T 2 (4L 2 /π 4 ), is an estimate of the gravitationalbackground S gw . The estimate, obtain<strong>ed</strong> over a finite observation time t m , has astatistical error. It can be shown [11] that√Nh1 (ω)N h2 (ω)δS gw (ω) = S gw (ω) = √ (8.39)tm δ fwhere t m is the total measuring time and δ f is the frequency step in the powerspectrum. From equation (8.11) we have the obvious result that, for resonantdetectors, the error is smaller at the resonances. If the resonances of the twodetectors coincide the error is even smaller. In practice, it is better to have twodetectors with the same resonance and bandwidth. If one bandwidth is smaller, theminimum error occurs in a frequency region overlapping the smallest bandwidth.From the measur<strong>ed</strong> S gw one can calculate the value of the energy density ofthe stochastic GW referr<strong>ed</strong> to as the critical density (the energy density ne<strong>ed</strong><strong>ed</strong>for a clos<strong>ed</strong> universe). We have = 4π 2 f 23 H 2 S gw( f ) (8.40)where H is the Hubble constant.


Recent results obtain<strong>ed</strong> with the resonant detectors 99Table 8.2. Sensitivity of the resonant detectors in operation.Resonance ˜h = √ S h Frequency Minimum h Minimum hfrequency(at resonance)bandwidth for 1 ms for continuous Minimum(Hz)√ 1δ f (Hz) bursts <strong>waves</strong> Hz900–700 3–20 × 10 −22 0.5–1 4× 10 −19 2 × 10 −25 0.1Table 8.3. Target sensitivity for Auriga and Nautilus.˜h = √ S h Frequency Minimum h Minimum hat(resonance)bandwidth for 1 ms for continuous Minimum√ 1δ f (Hz) bursts <strong>waves</strong> Hz2 × 10 −22 50 3 × 10 −21 2 × 10 −26 10 −48.3 Recent results obtain<strong>ed</strong> with the resonant detectorsThe present five cryogenic bars in operation [1–5] (Allegro, Auriga, Explorer,Nautilus and Niobe) have roughly the same experimental sensitivity as given intable 8.2.Niobe, made with niobium, has a resonance frequency of 700 Hz, the otherones with aluminium, have resonance frequencies near 900 Hz. These minimumvalues for monochromatic <strong>waves</strong> and for the quantity have been estimat<strong>ed</strong> byconsidering one year of integration time (for we suppose to use the crosscorrelationof two identical antennae).The burst sensitivity for all bars can be increas<strong>ed</strong> by improving the transducerand associat<strong>ed</strong> electronics. It has been estimat<strong>ed</strong> that these improvements canincrease the frequency bandwidth up to 50 Hz.In addition to increasing the bandwidth, Auriga and Nautilus can improve(see table 8.3) their spectral sensitivity by making full use of their capabilityto go down in temperature to T = 0.10 K. At present the major difficulty isdue to excess noise, sometimes of unknown origin, and work is in progress foreliminating this noise.The search for signals due to GW bursts is done after the raw data havebeen filter<strong>ed</strong> with optimum filter algorithms [13, 14]. These algorithms may havevarious expressions but they all have in common an optimum integration timethat is roughly the inverse of the detector bandwidth and all give approximatelythe same value of T eff (all algorithms being optimal filters for short bursts).The data record<strong>ed</strong> by the various detectors are now being analys<strong>ed</strong>,searching, in particular, for coincidences above the background. The coincidence


100 Resonant detectors for gravitational <strong>waves</strong> and their bandwidthTable 8.4. Results of a coincidence search between data from Explorer and Nautilus. Seethe text for an explanation.Number of Number of Number of p poisson p expdays Explorer events Nautilus events 〈n〉 n c (%) (%)29.2 8527 5679 11.0 19 1.5 1.44technique is a powerful mean for r<strong>ed</strong>ucing the noise. As an example we showin table 8.4 some results obtain<strong>ed</strong> recently [15] by searching for coincidencesbetween Explorer and Nautilus during 1995 and 1996.In the first column we give the number of days when both antennae wereoperating. The small number of useful days shows that it is difficult to keep a GWantenna in operation continuously with good behaviour. In future it should bepossible to increase the useful time to 70% of the total time, considering that sometime is always lost for cryogenic maintenance. In the second and third columnwe show the number of candidate events. The candidate events are obtain<strong>ed</strong> byintroducing a proper threshold on the data filter<strong>ed</strong> with an optimum filter for shortburst detection. We notice the large number of candidate events that make itpractically impossible, using one detector alone, to search for a particular signaldue to a GW. A big improvement is obtain<strong>ed</strong> by the comparison of at least twodetectors.In the fourth column we give the expect<strong>ed</strong> number of accidental coincidencesmeasur<strong>ed</strong> by means of 10 000 shifts of the event times of one detector with respectto the other and using a coincidence window of w = ±0.29 s (one samplingtime for EXPLORER and NAUTILUS). This number of accidentals is smallenough to start considering the possibility of searching for a coincidence excess(though, according to astrophysical expectations, this excess should be muchsmaller than the observ<strong>ed</strong> accidentals). In this case the number of coincidences n c ,report<strong>ed</strong> in column five, turns out to be slightly larger than the expect<strong>ed</strong> numberof accidentals.Finally, in column six we report the probability calculat<strong>ed</strong> with the Poissonformula and in column seven the experimental probability, obtain<strong>ed</strong> by countinghow many times we had a number of accidental coincidences equal or largerthan n c and dividing this number by the number of trials, 10 000. Theagreement between theoretical values and experimental values is good, indicatinga poissonian statistical behaviour of the data.A new and interesting result has recently been obtain<strong>ed</strong> by cross-correlatingthe data record<strong>ed</strong> with EXPLORER and NAUTILUS [17]. The measur<strong>ed</strong> spectralamplitudes of Explorer and Nautilus, shown in figure 8.2, were correlat<strong>ed</strong>. Theresult of such cross-correlation at a resonance of 907.2 Hz (the same for the twodetectors) has been the determination of an upper limit for that measures the


Discussion and conclusions 101Figure 8.2. Spectral amplitude ˜h of EXPLORER and NAUTILUS versus frequency (Hz).closure of the universe. It has been found, using equation (8.40), that ≥ 60,where a factor of six has been also includ<strong>ed</strong> for taking into consideration the factthat EXPLORER and NAUTILUS are separat<strong>ed</strong> by about 600 km. The value of we have obtain<strong>ed</strong> is much larger than the expect<strong>ed</strong> value, but we remark thatit is the first measurement of this type made with two cryogenic resonant GWdetectors.Finally, we would like to mention that recently, making use of the detectorNAUTILUS, we have been able to observe the passage of cosmic rays 1 . Theobserv<strong>ed</strong> events in the bar have amplitude and frequency of occurrence inagreement with the pr<strong>ed</strong>iction (event energy of the order of 1 mK), showing thatNAUTILUS is properly working. In particular, the efficiency of the filter aim<strong>ed</strong>at detecting a small signal emb<strong>ed</strong>d<strong>ed</strong> into noise is well proven.8.4 Discussion and conclusionsIn previous literature the sensitivity for resonant detectors of gravitational <strong>waves</strong>has usually been express<strong>ed</strong> in terms of T eff , the minimum energy deliver<strong>ed</strong> bya GW burst that can be detect<strong>ed</strong> by the apparatus. However, what the resonantdetectors really measure is essentially the Fourier transform (over a certainfrequency band) of the GW adimensional amplitude.For studying the operation of a resonant antenna as a GW detector ofstochastic GW we had to deal with noise spectrums. This has made us reconsider1 This result has been present<strong>ed</strong> at the School by Dr Evan Mauceli.


102 Resonant detectors for gravitational <strong>waves</strong> and their bandwidththe sensitivity to bursts and other types of GW in a somewhat different manner,which improv<strong>ed</strong> our understanding of the role play<strong>ed</strong> by the electromechanicaltransducer and its associat<strong>ed</strong> electronics. The noise spectrum of the apparatus isexpress<strong>ed</strong> by equation (8.11), which also directly gives the sensitivity for the GWbackground.We notice that the optimum sensitivity (at the resonance) depends essentiallyon the ratio T e /MQ and is independent of the transducer. In practice, thetransducer and electronics determine only the bandwidth of the apparatus,express<strong>ed</strong> by equation (8.14).For the measurement of the GW stochastic background, that should notchange drastically in a frequency band of a few hertz, it might therefore besufficient to make use of a very simple transducer and even small bandwidthelectronics. This makes the resonant detectors, in particular, well suit<strong>ed</strong> formeasuring the GW stochastic background.The use of a sophisticat<strong>ed</strong> transducer, with larger bandwidth, improves thesensitivity to GW bursts. As far as the search for monochromatic <strong>waves</strong> a largerbandwidth is better, in the sense that it allows a larger frequency region in whichto search.References[1] Astone P et al 1993 Phys. Rev. D 47 362[2] Mauceli E et al 1996 Phys. Rev. D 54 1264[3] Blair D G et al 1995 Phys. Rev. Lett. 74 1908[4] Astone P et al 1997 Astropart. Phys. 7 231–43[5] Cerdonio M et al 1995 <strong>Gravitational</strong> wave experiments Proc. First Edoardo AmaldiConf. (Frascati, 1994) <strong>ed</strong> E Coccia, G Pizzella and F Ronga (Singapore: WorldScientific)[6] Pizzella G 1975 Riv. Nuovo Cimento 5 369[7] Pallottino G V and Pizzella G 1998 Data Analysis in Astronomy vol III, <strong>ed</strong> V Di Gesuet al (New York: Plenum) p 361[8] Papoulis A 1984 Signal Analysis (Singapore: McGraw-Hill) pp 325–6[9] Pizzella G 1979 Nuovo Cimento 2 C209[10] Pallottino G V and Pizzella G 1984 Nuovo Cimento 7 C155[11] Bendat J S and Piersol A G 1966 Measurement and Analysis of Random Data (NewYork: Wiley)[12] Brustein R, Gasperini M, Giovannini M and Veneziano G 1995 Phys. Lett. B 36145–51[13] Astone P, Buttiglione C, Frasca S, Pallottino G V and Pizzella G 1997 Nuovo Cimento20 9[14] Astone P, Pallottino G P and Pizzella G 1998 Gen. Rel. Grav. 30 105[15] Astone P et al 1999 Astropart. Phys. 10 83–92[16] Astone P, Pallottino G V and Pizzella G 1997 Class. Quantum Grav. 14 2019–30[17] Astone P et al 1999 Astron. Astrophys. 351 811–14


Chapter 9The Earth-bas<strong>ed</strong> large interferometer Virgoand the Low Frequency FacilityAngela Di VirgilioINFN Sezione di Pisa, via Vecchia Livornese 1265,56010 S. Piero a Grado (Pisa), ItalyE-mail: angela.divirgilio@pi.infn.itThe principle of the Earth-bas<strong>ed</strong> interferometer and the parameters of Virgo, theCNRS-INFN antenna under construction in Cascina near Pisa, are present<strong>ed</strong>. TheSuper-Attenuator (SA), the suspension ad hoc design<strong>ed</strong> for Virgo, and the LowFrequency Facility, the project devot<strong>ed</strong> to study the SA with a sensitivity close tothe Virgo one, are shortly describ<strong>ed</strong>.9.1 IntroductionThe detection of gravitational <strong>waves</strong> (GW), produc<strong>ed</strong> in astrophysical processes,is one of the most challenging fields of modern physics, aim<strong>ed</strong> at findinggravitational-wave astronomy. Lectures about GW production and expect<strong>ed</strong>sources can be found in this book, but let me very briefly remind you ofthe essentials. General relativity pr<strong>ed</strong>icts that accelerat<strong>ed</strong> matter producesgravitational <strong>waves</strong>, ripples of spacetime travelling with the spe<strong>ed</strong> of light. GWcan be defin<strong>ed</strong> as a perturbation of the metric, h αβ , and the observational effect ofits passage through the Earth is a change among the relative distance of massivebodies. If A and B are the space-state locations of the two testmasses and if weset AB = (x α ), the vector (x α ) obeys the equation of geodesic deviation in weakfield:d 2 x αdt 2 = 1 ∂ 2 h TTαβ2 ∂t 2 x β (9.1)103


104 The Earth-bas<strong>ed</strong> large interferometer Virgowhere TT denotes the transverse-traceless gauge. The maximum change in th<strong>ed</strong>istance AB is then:δL = h 2 L (9.2)where h is the dimensionless gravitational-wave amplitude and L is the distanceAB at rest. One of the most important gravitational-wave sources is theSuperNovae explosion. In order to detect several SuperNovae events per year, it isnecessary to have a sensitivity of h close to 10 −21 for a millisecond signal; in factthis is the expect<strong>ed</strong> amplitude for gravitational <strong>waves</strong> from SuperNovae comingfrom the closest cluster of galaxies: Virgo, distant 10–20 Mpc. The idea of usingan interferometer to detect gravitational <strong>waves</strong> has been propos<strong>ed</strong> independently,about 30 years ago, by Weber and two Russian physicists, Gerstenstein andPustovoit [1] and a lot of work has been done to develop the optical design ofthe large interferometers presently under construction [2, 3]. Let me mention(see, for example, [4]) the work done in Garching, Glasgow, Caltech and MIT,where interferometers of 30–40 m arms have been develop<strong>ed</strong> and are still runningwith displacement sensitivity equal to the shot noise limit. The power spectrumof the displacement sensitivity obtain<strong>ed</strong> so far is of the order of 10 −19 m/ √ Hz.Following the above equation it is straightforward to see that a few kilometrearmlength is necessary in order to reach the sensitivity of h = 10 −21 for themillisecond signal. Several antennae are under construction [5]: two antennae4 km arms in the USA (LIGO), one 0.6 km in Germany (GEO600), one inJapan 0.3 km (TAMA), in Italy 3 km (Virgo) and one is plann<strong>ed</strong> with 3 kmarms in Australia (ACIGA). Most of them (TAMA, LIGO and GEO) have beenconstruct<strong>ed</strong> and are making working together parts of the interferometer. TAMAis already operational, and work is concentrat<strong>ed</strong> in order to reach the plann<strong>ed</strong>sensitivity; moreover Japan already plans to build a longer one. The aboveinterferometers, all Earth bas<strong>ed</strong>, will cover the 10–10 000 Hz window, whereevents from SuperNovae explosions, coalescing binaries and pulsars events areexpect<strong>ed</strong>. Using the signals of the resonant bars, and the other kind of detectorsalready operational, in the near future gravitational-wave astronomy is expect<strong>ed</strong>to become a reality. Furthermore, a big space antenna is plann<strong>ed</strong> to be launch<strong>ed</strong>in this decade, it is call<strong>ed</strong> LISA [5], and will cover the frequency spectrum 10 −4 –10 −1 Hz. In the following the essential characteristics of the Virgo antenna [6,7],the Virgo SuperAttenuator suspension, the Low Frequency Facility [9] and the Rand D experiment of the Virgo project, are describ<strong>ed</strong>.9.1.1 Interferometer principles and Virgo parametersFigure 9.1 shows a simple Michelson interferometer: the passage of agravitational wave changes the phase shift between the two outgoing beamsinterfering on the beam splitter. For the sake of simplicity it is assum<strong>ed</strong> thatthe wave is optimally polariz<strong>ed</strong> and travels perpendicularly to the interferometer


Introduction 105Figure 9.1. The simple Michelson.plane, moreover the Michelson is assum<strong>ed</strong> ideal [10, 11] (contrast C = 1):δφ = π + α + gw . (9.3)Here α is the offset between the two interferometer arms and gw is the effect ofthe gravitational wave. The transmitt<strong>ed</strong> outgoing power P out has a simple relationwith the input power P in .P out = P in sin 2 α + gw2(9.4)for gw ≪ 1P out ≃ P in[sin 2 α 2 + 1 2 sin α gw]. (9.5)Equation (9.5) clearly states that P out is proportional to gw , the gravitational<strong>waves</strong>ignal, and the fundamental limit of the measurement comes from thefluctuation of the incoming power, i.e. the fluctuation of the number of photons.For a coherent state of light, the number of photons fluctuates with Poissonianstatistics, and it can be shown that the signal-to-noise ratio (SNR), when C = 1,is:√P in tSNR =¯hω cos a gw (9.6)where ω is the circular frequency of the laser light, t is the time of the measureand ¯h is the Planck constant. The SNR is maximum when α = 0 and the interferometeris tun<strong>ed</strong> to a dark fringe. The minimum phase shift detectable, per unittime, is evaluat<strong>ed</strong> assuming SNR = 1 mingw= √¯hωP in(9.7)


106 The Earth-bas<strong>ed</strong> large interferometer VirgoFigure 9.2. The interferometer with Fabry–Perot cavity and power recycling.and it is call<strong>ed</strong> the shot noise limit. The contrast C < 1 does not change thisresult. The large interferometer design has been optimiz<strong>ed</strong> in order to enhance thesensitivity: inside the arm there is a resonant Fabry–Perot cavity; this is equivalentto increasing the interferometer armlength, by a coefficient proportional to thefinesse of the Fabry–Perot cavity. Moreover, as shown in equation (9.5), the SNRincreases with the square root of the power and, since the output is at the darkfringe, the unus<strong>ed</strong> light, reflect<strong>ed</strong> back from the interferometer, is re-inject<strong>ed</strong> witha technique call<strong>ed</strong> recycling. Figure 9.2 shows the basic scheme of most of theinterferometers under construction [3].The other fundamental noise of the interferometer is the so call<strong>ed</strong> ‘thermalnoise’ [11, 12]. The Virgo test masses are in thermal equilibrium with theenvironment. An average energy of kT, where k is the Boltzmann constantand T is the temperature, is associat<strong>ed</strong> with each degree of fre<strong>ed</strong>om, givingorigin, through the ‘fluctuation dissipation theorem’, to position fluctuations ofthe macroscopic coordinates. Using the equation of motion and the fluctuationdissipationtheorem, the displacement noise is estimat<strong>ed</strong>. There are two maindissipation mechanisms: the structural and the viscous damping. It is assum<strong>ed</strong>that mechanisms of viscous damping are pretty much r<strong>ed</strong>uc<strong>ed</strong>, and only thestructural one survives. Internal friction in materials is describ<strong>ed</strong> by adding animaginary term to the elastic constant, and the equation of motion for a springbecomes:mẍ + K (1 + i)x = 0 (9.8)where x is the coordinate of interest, m is the mass associat<strong>ed</strong> with the oscillator,K is the elastic constant and takes care of the dissipation and is relat<strong>ed</strong> to theusual Q of the oscillator.As most of the interferometers under construction, Virgo [6, 7] will be arecycl<strong>ed</strong> interferometer, with resonant Fabry–Perot cavity in each arm. Thearm length is 3 km, the recycling factor 40 and the finesse of the cavity 40;the light source is a 25 W Nd:YAG laser (λ = 1.06 µm). The Virgo Projectis under construction in Italy at Cascina, near Pisa, by CNRS (France) andINFN (Italy). The participating laboratories are: ESPCI-Paris, Nice, LAL-


h(sqrt(1/Hz))10 −610 −810 −1010 −1210 −1410 −1610 −1810 −2010 −2210 −2410 −2610 −28Introduction 107Horizontal SeismVert. Seism * 10 −2Thermal NoiseQuantum LimitShot NoiseNewtonianSensitivity curve10 −1 10 0 10 1 10 2 10 3 10 4frequency (Hz)Figure 9.3. Virgo sensitivity curves.Orsay, LAPP-Annecy, INPN-Lyon for CNRS and Firenze, Frascati, Napoli,Perugia, Pisa, Roma for INFN. The aim is to reach, in a useful detectionbandwidth between a few hertz up to a few kilohertz, a sensitivity which allowsus to detect gravitational <strong>waves</strong> emitt<strong>ed</strong> by coalescing binary stars, gravitationalcollapses, spinning neutron stars or constituting the stochastic background of thegravitational radiation. The Virgo sensitivity curve is shown in figure 9.3, asa function of frequency; the vertical scale shows the linear spectral density ofthe minimum detectable dimensionless gravitational amplitude: at low frequencythere is the pendulum thermal noise, then the mirror substrate thermal noise andat high frequency the shot noise of the laser light, the interferometer sensitivitydecreases at high frequency for the high storage time of the 3 km Fabry–Perotcavity. Figure 9.3 shows the thermal and shot noise and the other important noiseof the Virgo antenna.The laser has to be frequency stabiliz<strong>ed</strong> at the level of 10 −1 Hz/ √ (Hz) inthe frequency band of interest [8], the mirror losses have to be below 1 ppmin order to obtain the requir<strong>ed</strong> recycling factor and cavity finesse; moreover,power stabilization is requir<strong>ed</strong>, and a mode cleaner 144 m long, with finesse1000, is necessary in order to have a good quality TM 00 mode inject<strong>ed</strong> into theinterferometer, and to contribute to the amplitude stability, since such a long modecleaner has a pole at 10 Hz. All the optical components will be suspend<strong>ed</strong> invacuum by anti-seismic suspensions call<strong>ed</strong> Super Attenuators (SA); the opticalpath will be completely under vacuum. The vacuum system will consist of: twoorthogonal vacuum tubes, 3 km long, 1.2 m in diameter, containing the arm opticalpaths; one vacuum tube, 144 m long, 0.3 m in diameter for the mode cleaner


108 The Earth-bas<strong>ed</strong> large interferometer VirgoFigure 9.4. Scheme of the Virgo interferometer.cavity; ten vertical vacuum towers, up to 11 m high, 2 m in diameter, containingthe different SAs. Figure 9.4 shows the Virgo scheme. The construction of Virgois plann<strong>ed</strong> in two main steps: the realization of the central interferometer by late2000, and the realization of the 3 km arm full interferometer by late 2002.9.2 The SA suspension and requirements on the controlRoughly speaking, above a few hertz, the seismic noise is a function of thefrequency ν and goes as ν −2 . In our site the measur<strong>ed</strong> spectrum is about 10 −6 –10 −7 /ν 2 m/ √ Hz, in all degrees of fre<strong>ed</strong>om; this implies a factor of at least10 10 of the noise r<strong>ed</strong>uction at 10 Hz. An harmonic oscillator, with resonantfrequency ν 0 acts as a low pass filter if N oscillators are connect<strong>ed</strong> in cascade,and for frequencies above the resonant frequencies, the displacement noise of theattachment point is transmitt<strong>ed</strong> to the test-mass with a transfer function going as∼1/ν 2N . Each SA is a multistage pendulum [13], acting as a multipendulum inall six degrees of fre<strong>ed</strong>om. The basic tool is a piece call<strong>ed</strong> a mechanical filter. Itis essentially a rigid steel cylinder (70 cm in diameter, 18 cm in height and with amass of about 130 kg), with a thin steel wire. In this way the horizontal oscillatorin two degrees of fre<strong>ed</strong>om and the torsion spring are made. The attachment pointsof the two wires are connect<strong>ed</strong> close to the centre of mass of the filter: in this waythe pitch spring is form<strong>ed</strong>. The vertical string is a mechanical one, a triangularcurl<strong>ed</strong> piece of a special steel (MARAGING). The SA blades are design<strong>ed</strong> tohold about 50 kg, at rest they are bent and when load<strong>ed</strong> act as a spring andbecome flat. Each filter of the SA is equipp<strong>ed</strong> with a number of blades from12–4, in function of the load. Each filter of the SA is compos<strong>ed</strong> of two main


18.5 cmThe SA suspension and requirements on the control 109to the previousstagecrossbarmagnetsmagnetscenteringwirescenteringwiresbellowbellow moving element bladecenteringwiresto the nextstagecentralcolumnbladehinge70 cmFigure 9.5. The filter.parts: the vessel, roughly speaking a cylinder with a hole in the centre and amovable part, compos<strong>ed</strong> of a piece call<strong>ed</strong> ‘cross’ attach<strong>ed</strong> to a tube call<strong>ed</strong> ‘centralcolumn’, rigidly connect<strong>ed</strong> together. The ‘central column’ is free to move alongthe vertical inside the vessel, for about a centimetre. The upper wire is attach<strong>ed</strong>to the vessel, while the lower is attach<strong>ed</strong> to the column. The two pieces areconstrain<strong>ed</strong> to move along the vertical by means of centring wires. The baseof each blade is rigidly attach<strong>ed</strong> to the outer diameter of the vessel, while the tipsare connect<strong>ed</strong> to the central column. When the filter is load<strong>ed</strong> the blades act as aspring, and are free to move for several millimetres, with a resonant frequencyabove 1 Hz. In order to decrease the vertical resonant frequency a magneticantispring acts between the vessel and the ‘cross’, in parallel with the blades.It is compos<strong>ed</strong> of an array of permanent magnets attach<strong>ed</strong> to the ‘cross’, plac<strong>ed</strong>in front of an equal array of magnets attach<strong>ed</strong> to the vessel; the two array aread hoc built in order to make repulsive force, proportional to the vertical offset,for a few millimetres. The effective elastic constant is the difference between theblade spring and the magnetic antispring, and each filter is tun<strong>ed</strong> to oscillate at∼400 mHz. In figure 9.3 with the other noise sources, the seismic motion of theVirgo test masses is shown [14], it is possible to see that it dominates below 2 Hz,and the resonance peaks of the SA are visible.Ground motion can be of the order of hundr<strong>ed</strong>s of micrometres, due to tidaldeformation of the Earth’s crust and thermal effects on the surface. In order to


110 The Earth-bas<strong>ed</strong> large interferometer VirgoFigure 9.6. The complete SA suspension, some details of the last stage control are shown.keep the interferometer running and compensate such a large drift, the SA isequipp<strong>ed</strong> with a special stage call<strong>ed</strong> an invert<strong>ed</strong> pendulum (IP). It is compos<strong>ed</strong>of three aluminium legs, about 6 m tall. When it is load<strong>ed</strong> with the 1.2 ton of theSA, it becomes a very soft spring, with about 30 mHz frequency. With magnetson top of the IP and coils attach<strong>ed</strong> to the ground, forces are produc<strong>ed</strong>, and it ispossible to displace to SA by a few centimetres. It is necessary not to introduceseismic noise through the forces to be appli<strong>ed</strong> to the test mirror in order to keepthe interferometer in its working position. This is an important task of the SA,and all the control forces are done by means of coils and magnets acting between


A few words about the Low Frequency Facility 111different stages of the SA. The last part of the SA is compos<strong>ed</strong> of steering filters,which support four coils, the marionetta, a cross shap<strong>ed</strong> piece of steel, whichsupports four magnets, position<strong>ed</strong> in the form of four coils, see figure 9.6. In thisway the mirror can translate along the beam direction and along the vertical, rotatearound the vertical axis and around the horizontal axis perpendicular to the beam.The test-mass (the interferometer mirror) and the reference mass are attach<strong>ed</strong> aspendulums to the marionetta [15]. Magnets and coils produce forces between thereference mass and the mirror.The control has two main parts: the damping and the mirror control. Th<strong>ed</strong>amping is a local control, which r<strong>ed</strong>uces the large motion of the SA, due to theresonance frequencies below 1 Hz: on top of the invert<strong>ed</strong> pendulum fe<strong>ed</strong>backsare produc<strong>ed</strong> using accelerometer signals and electromagnetic actuators [16]. Theother control uses the interferometer signal itself and it is aim<strong>ed</strong> to keep align<strong>ed</strong>the whole Virgo interferometer in its working point. The actuators, couples ofmagnet-coil, are divid<strong>ed</strong> per frequency band: below 100 mHz the fe<strong>ed</strong>back goesto the IP actuators, in the interm<strong>ed</strong>iate band to the marionetta, and above to thereference mass. This kind of control, call<strong>ed</strong> hierarchical is necessary for the hug<strong>ed</strong>isplacement range requir<strong>ed</strong> and the necessity not to induce electronic noise atthe mirror level.9.3 A few words about the Low Frequency FacilityMuch work is being done around interferometers at the moment, but the plann<strong>ed</strong>sensitivity is not enough to guarantee the detection of events. A large community[17] is working hard to improve the sensitivity of the antennae. LIGO plansto make an upgrade in five years from now. The effort is mainly concentrat<strong>ed</strong>on improving the suspension and keeping as low as possible the thermal noiseof pendulum and mirror substrate. With the aim of r<strong>ed</strong>ucing thermal noise,several methods have been propos<strong>ed</strong>, like cryogenic cooling, choice of low lossmaterial for the test mass, as zaffire or silicon, and novel cancellation techniques.Figure 9.3 shows that, below 600 Hz, Virgo is entirely dominat<strong>ed</strong> by the thermalnoise of the suspension. The improvement below 600 Hz implies changes inthe suspension. In order to demonstrate that a new choice, once appli<strong>ed</strong> to theantenna, makes a real improvement in the sensitivity of the whole antenna it isnecessary to make tests in conditions as close as possible to the real one. TheLow Frequency Facility (LFF) [9, 19], has been conceiv<strong>ed</strong> in order to measurethe noise of test masses attach<strong>ed</strong> to the SA, with a sensitivity at 10 Hz close tothe Virgo one: 10 −18 m/ √ Hz. The goal is to take a measurement of the thermalnoise of the Virgo suspension, while the system is under active control of the lowfrequencymotion. In this way, it will be possible to characterize the suspensionin an environment close to the real one.The use of a Fabry–Perot cavity as a displacement transducer relies on theexcellent frequency stabilization of the laser beam. A frequency stabilization, for


112 The Earth-bas<strong>ed</strong> large interferometer Virgoa Nd-YAG, of the order of 10 −1 –10 −2 Hz/ √ Hz at 10 Hz has been obtain<strong>ed</strong>, forexample by the Nice-Virgo team [8]. A copy of this circuit provides the frequencystabilization for the LFF. It is already operational and has obtain<strong>ed</strong> a clos<strong>ed</strong> loopbehaviour similar to the Virgo prestabilization circuit. The technique to r<strong>ed</strong>ucethe influence of power fluctuation is to extract signals from the Fabry–Perot withthe well known Pound–Drever–Hall, where the inject<strong>ed</strong> light is phase modulat<strong>ed</strong>at high frequency (usually around 10 MHz, where the laser is shot noise limit<strong>ed</strong>).The signal is extract<strong>ed</strong> in reflection, the noise is due to the shot noise, and theSNR for the cavity detuning δν is:SNR = P in8J 0 (m)J 1 (m)F(1 − A loss Fπ) 2L c δν2 √ . (9.9)ηP DC hν 0where P in represents the incident power on the cavity, A loss the total loss in thecavity during one round trip, P DC is the DC term of the power reflect<strong>ed</strong> fromthe cavity, √ 2ηP DC hν is the shot noise expression with η the quantum efficiencyof the photodiode, h is the Planck constant, ν 0 is the laser frequency, L is thecavity length, F is the finesse, c is the spe<strong>ed</strong> of light and J 0 and J 1 are Besselfunctions, functions of the modulation depth m. The LFF cavity parameters are:Finesse 3000, P 0 = 30 mW, cavity length L = 1 cm. With these parameters,the sensitivity at 10 Hz is limit<strong>ed</strong> by the frequency fluctuation, and a powerspectrum density of 10 −1 Hz/ √ Hz is necessary in order to reach the displacementsensitivity of 10 −18 m/ √ Hz at 10 Hz. A prototype of the SA, call<strong>ed</strong> R andD SA, has been assembl<strong>ed</strong> in Pisa, and has test<strong>ed</strong> the assembly proc<strong>ed</strong>ure anddevelop<strong>ed</strong> the control loop. LFF will suspend to the R and D SA two mirrors,in a way to create a very high finesse Fabry–Perot; of whose element, one is thestandard Virgo mirror, with a mass above 20 kg, while the other [18], usuallycall<strong>ed</strong> (AX), will be different. In the early phase of the experiment the auxiliarymirror will be a standard small mirror; the auxiliary mirror will be load<strong>ed</strong> withan extra mass, in order to r<strong>ed</strong>uce the displacement noise due to low-frequencyradiation pressure. The Fabry–Perot cavity will act as a displacement transducerand allow the measurement of the combin<strong>ed</strong> thermal noise of the two mirrors, andrelat<strong>ed</strong> suspension. Figure 9.7 shows the experimental apparatus: the R and D SAholding the two mirrors of the cavity and the table with the optical circuit, whichwill inject light into the cavity.9.4 ConclusionSeveral gravitational-wave interferometers will be operational in a few years,and a lot of effort is being concentrat<strong>ed</strong> in building and making them work.Expectations are great around such experiments, since they can give essentialinformation on general relativity, fundamental physics and astrophysics. TheVirgo antenna will be operational in 2–3 years from now. A large study is goingon around the SA suspension. A special prototype call<strong>ed</strong> R and D SA is under


Conclusion 113Figure 9.7. The R and D SA and the injection table of the LFF, inside vacuum tanks; theenlarg<strong>ed</strong> picture shows the two mirrors forming a Fabry–Perot cavity.


114 The Earth-bas<strong>ed</strong> large interferometer Virgotest in Pisa. The R and D SA is part of the LFF experiment, aim<strong>ed</strong> to measurethermal noise of the test mass suspend<strong>ed</strong> to the R and D SA, and to become a testarea for future improv<strong>ed</strong> antenna suspensions.References[1] Weber J 1960 Phys. Rev. 117 306Gerstenstein and Pustovoit 1963 Sov. Phys.–JETP 16 433[2] Blair D G (<strong>ed</strong>) 1991 The Detection of <strong>Gravitational</strong> Waves (Cambridge: CambridgeUniversity Press)[3] Giazotto A Phys. Rev. 182 365[4] Shoemaker D et al 1988 Phys. Rev. 38 423[5] See articles in: 1999 <strong>Gravitational</strong> Waves, 3rd Edoardo Amaldi Conf. (Caltec, July)(Singapore: World Scientific)[6] Brillet A et al 1992 VIRGO final conceptual designGamaitoni L, Kovalik J and Punturo M 1997 The VIRGO sensitivity curve VIRGOInternal Note NOT-PER-1390-84, 30-3[7] 1997 Virgo—Final Design Report May[8] Bondu F et al 1996 Opt. Lett. 14 582[9] Di Virgilio A et al 1997 Virgo Note VIR-NOT-PIS-8000-101Bernardini M et al 1998 Phys. Lett. A 243 187–94[10] Hello P 1998 Progress in Optics vol XXXVIII, <strong>ed</strong> E Wolf (Amsterdam: Elsevier)[11] See, for example, Saulson P R 1994 Fundamentals of Interferometric <strong>Gravitational</strong>Wave Detectors (Singapore: World Scientific)[12] Bondu F, Hello P and Vinet J-Y 1998 Phys. Lett. A 246 227–36Levin Yu 1998 Phys. Rev. D 57 659–63[13] Braccini S et al 1996 Rev. Sci. Instrum. 67 2899–902Beccaria M et al 1997 Nucl. Instrum. Methods A 394 397–408Braccini S et al 1997 Rev. Sci. Instrum. 68 3904–6De Salvo R et al 1999 Nucl. Instrum. Methods A 420 316–35Losurdo G et al 1999 Rev. Sci. Instrum. 70 2507–15[14] Vicerè A 2000 Introduction to the mechanical simulation of seismic isolationsystems Experimantal Physics of <strong>Gravitational</strong> Waves <strong>ed</strong> M Barcone, G Calamai,M Mazzoni, R Stanga and F Vetrano (Singapore: World Scientific)[15] Bernardini A et al 1999 Rev. Sci. Instrum. 70 3463–71[16] Losurdo G 2000 Active controls in interferometric detectors of gravitational<strong>waves</strong>: inertial damping of VIRGO super attenuators Experimantal Physics of<strong>Gravitational</strong> Waves <strong>ed</strong> M Barcone, G Calamai, M Mazzoni, R Stanga andF Vetrano (Singapore: World Scientific)Losurdo G 2000 Preprint gr-qc/00020006[17] Contribution to the Aspen Winter Conf. on <strong>Gravitational</strong> Waves, and their Detectioncan be found on: http://www.ligo.caltech.<strong>ed</strong>u/[18] Cella G et al 1999 Suspension for the Low Frequency Facility Phys. Lett. A accept<strong>ed</strong>[19] Benabid F et al 2000 The Low Frequency Facility, R&D experiment of the VirgoProject J. Opt. B accept<strong>ed</strong> (special issue)


Chapter 10LISA: A propos<strong>ed</strong> joint ESA–NASAgravitational-wave missionPeter L Bender, on behalf of the LISA Study and MissionDefinition TeamsJILA, National Institute of Standards and Technology andUniversity of Colorado, Boulder, Colorado 80309-0440, USA10.1 Description of the LISA mission10.1.1 IntroductionThe evidence for supermassive black holes at the centres of quasars and activegalactic nuclei (AGNs) has been strong but not conclusive for several decades.Recently, the evidence for massive black holes in one particular AGN, in our ownGalaxy, and in the Local Group Galaxy M32 has become extremely convincing.Thus, questions concerning gravitational <strong>waves</strong> generat<strong>ed</strong> by the interaction ofmassive black holes with smaller compact objects and with each other havebecome of strong interest. Signals from such sources also are likely to providethe strongest possible tests of general relativity.Plans are being develop<strong>ed</strong> in both Europe and the USA for flying ad<strong>ed</strong>icat<strong>ed</strong> gravitational wave mission call<strong>ed</strong> the Laser Interferometer SpaceAntenna (LISA). The antenna will measure gravitational <strong>waves</strong> in the frequencyrange from roughly 1 µHz to 1 Hz, and thus will strongly complement the resultsexpect<strong>ed</strong> from ground-bas<strong>ed</strong> detectors. The primary objectives are to obtainunique new information about massive black holes throughout the universe, andto map the metric around massive black holes (MBHs) with much higher accuracythan otherwise would be possible. Other important objectives include studies ofresolv<strong>ed</strong> signals from thousands of compact binary star systems in our Galaxy,and looking for a possible gravitational-wave cosmic background at millihertzfrequencies.115


116 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionThe possibility of making sensitive gravitational-wave measurements at lowfrequencies by laser interferometry between freely floating test masses in widelyseparat<strong>ed</strong> spacecraft appears to have been first suggest<strong>ed</strong> in print about 1972 [1,2].More extensive discussions start<strong>ed</strong> in 1974. An initial proposal similar to thatfor the LISA mission was present<strong>ed</strong> at the Second International Conference onPrecision Measurement and Fundamental Constants in 1981 [3], and at the ESAColloquium on Kilometric Optical Arrays in Space in 1984 [4]. Work on theconcept was support<strong>ed</strong> initially by the National Bureau of Standards, and later inthe USA by NASA. However, the concept became much better defin<strong>ed</strong> and widelyknown in the 1993–1994 period, when more extensive studies were carri<strong>ed</strong> out inEurope under ESA support. Since then, further studies have been carri<strong>ed</strong> out byboth ESA and NASA of a propos<strong>ed</strong> joint ESA–NASA mission, that could fly asearly as 2010 if all goes well.The currently propos<strong>ed</strong> mission is describ<strong>ed</strong> in section 1 of this chapter. Thisincludes the overall antenna and spacecraft design, the optics and interferometrysystem, the free mass sensors, the requir<strong>ed</strong> micronewton thrusters for thespacecraft, and the mission scenario. The emphasis will be on aspects of theantenna that are quite different from those for ground-bas<strong>ed</strong> detectors. Section 2describes the main scientific results that seem likely to be obtain<strong>ed</strong> by LISA.This includes unique new information on three major astrophysical questionsconcerning MBHs and a nearly ideal test of general relativity, as well as th<strong>ed</strong>etection of thousands of compact binaries in our Galaxy. In addition, somespeculations will be given on possible future prospects for gravitational-waveobservations in space after the LISA mission.Much more information on most of the above subjects can be found in ‘LaserInterferometer Space Antenna’, the Proce<strong>ed</strong>ings of the Second InternationalLISA Symposium [5], in the LISA Pre-Phase A Study [6], and in a special issueof Classical and Quantum Gravity [7], which is the Proce<strong>ed</strong>ings of the FirstInternational LISA Symposium. More recent information from the 1999–2000ESA Industrial Study of the LISA mission is being provid<strong>ed</strong> in the report of thatstudy and in several papers in the Proce<strong>ed</strong>ings of the Third International LISASymposium (Albert Einstein Institute, Potsdam, 11–14 July 2000).10.1.2 Overall antenna and spacecraft designThe basic geometry is shown schematically in figure 10.1. Three spacecraft forman equilateral triangle 5000 000 km on a side, and laser beams are sent both waysalong each side of the triangle. A Y-shap<strong>ed</strong> thermal shield inside each spacecraftcontains the sensitive parts of the scientific payload, consisting of two separateoptical assemblies mount<strong>ed</strong> in the two top arms of the Y, so that they are aim<strong>ed</strong>along the two adjacent sides of the triangle. The spacecraft instrumentation and acover over the sunward side of the spacecraft are not shown.Each spacecraft is in a one year period solar orbit with an eccentricity e ofabout 0.01 and an orbit inclination to the ecliptic of 3 0.5 × e [8]. By choosing


Description of the LISA mission 117Figure 10.1. Basic geometry of the LISA antenna. The sides of the triangle are5000 000 km long.the phasing and the orientations of the orbits properly, the three spacecraft willform the desir<strong>ed</strong> nearly equilateral triangle, which is tipp<strong>ed</strong> at 60 ◦ to the ecliptic(figure 10.2). The plane of the triangle will precess around the pole of the eclipticonce per year, and the triangle will rotate in that plane at the same rate. Thecentre of the triangle is chosen to be about 50 000 000 km (20 ◦ ) behind the Earth.The arm lengths for the triangle will stay constant to about 1% for a number ofyears. Locating the triangle 60 ◦ from the Earth would keep the arm lengths moreconstant, but at the expense of more propulsion requir<strong>ed</strong> to reach the desir<strong>ed</strong> orbitsand more telemetry capability to send the data back.The layout of one of the two optical assemblies in each spacecraft is shownin figure 10.3 (see [9]). Each optical assembly contains an optical bench,a transmit/receive telescope, and a low-power electronics package. Each ofthese three subassemblies is mount<strong>ed</strong> by low-thermal-conductivity struts from astiffen<strong>ed</strong> support cylinder that forms the outside of the optical assembly. Becausethe distances between the spacecraft will change by up to 1% during the year,the angles between the sides will also change by roughly one degree. Thus, it isnecessary to provide some adjustment for the angle between the axes of the two


118 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionFigure 10.2. Location of the LISA antenna, 50 000 000 km behind the Earth in orbit aroundthe Sun. The plane of the antenna is tipp<strong>ed</strong> at 60 ◦ to the ecliptic.Figure 10.3. Layout of one of the two optical assemblies in each spacecraft. The maincomponents are the optical bench at the centre and the transmit/receive telescope at oneend.optical assemblies in each spacecraft. This is accomplish<strong>ed</strong> by supporting thelower part of each optical assembly by a flexure from a common block near thebase of the Y-shap<strong>ed</strong> shield, and the upper part by two adjustable displacement


Description of the LISA mission 119Figure 10.4. Mounting of optical assemblies inside main thermal shield. Each opticalassembly must rotate slowly and smoothly during the year over about a one degree range.Figure 10.5. Optical bench for one optical assembly. Light from the laser is mainly sentout to the telescope, with a little going to the photodiode. Receiv<strong>ed</strong> light from the other endof the arm passes through the beamsplitter to the test mass, and then goes to the photodiode.Changes in the phase of the beat signal determine the changes in the arm length.mechanisms from an arm of the Y, as shown in figure 10.4 (see [10]). Themechanisms provide a very smooth one degree change in the angle between thearms with mainly a one year period.At the centre of each optical bench, shown in figure 10.5 (see [9]), is afreely floating test mass that is protect<strong>ed</strong> as much as possible from sources ofspurious accelerations. The basic measurements made by the LISA antenna


120 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionare the changes with time in the distances between test masses in the differentspacecraft. Capacitive sensors on the inside of a housing surrounding each testmass sense the position of the spacecraft with respect to the test mass. Thisposition information is f<strong>ed</strong> via a proportional control servosystem to micronewtonthrusters that keep the spacecraft essentially fix<strong>ed</strong> with respect to the test mass.Each test mass plus the housing around it and the associat<strong>ed</strong> electronics willbe referr<strong>ed</strong> to as a ‘free mass sensor’. The names ‘inertial sensor’, ‘drag-freesensor’, and ‘disturbance r<strong>ed</strong>uction sensor’ also have been us<strong>ed</strong> frequently, buteach choice has some drawbacks. ‘Inertial sensor’ causes confusion with thegyroscopic sensors us<strong>ed</strong> in inertial navigation systems. ‘Drag-free sensor’ or‘disturbance r<strong>ed</strong>uction sensor’ emphasizes the role of the sensor in permittingexternal forces on the spacecraft to be accurately compensat<strong>ed</strong> for, but do notindicate the importance of keeping the spurious accelerations of the test mass farbelow the residual accelerations of the spacecraft.At frequencies below roughly 3 mHz, the threshold sensitivity of the LISAantenna will be determin<strong>ed</strong> mainly by the spurious accelerations of the testmasses, despite the care that is taken to minimize such disturbances. However,at higher frequencies the main limitation will come from how well the relativechanges in the 5000 000 km distances between the test masses can be measur<strong>ed</strong>.The optical interferometry system for measuring these distances will be describ<strong>ed</strong>in the next section. The free mass sensors and the way they are us<strong>ed</strong> will b<strong>ed</strong>iscuss<strong>ed</strong> in detail in a later section.In describing the operation of the gravitational-wave antenna, it is useful todistinguish between the ‘main spacecraft’ and the ‘instrument package’. Here theinstrument package will mean the Y-shap<strong>ed</strong> thermal shield and everything insideit. In addition, it will include the lasers and a thermal radiation plate they arelocat<strong>ed</strong> on that is attach<strong>ed</strong> to the bottom of the Y. With this arrangement, theheat generat<strong>ed</strong> by the lasers can be radiat<strong>ed</strong> by the radiation plate out through alarge hole in the bottom of the spacecraft to space. Everything else, including thespacecraft structure and all the other equipment mount<strong>ed</strong> on it, will be referr<strong>ed</strong> toas the main spacecraft.Probably the most important factor in the design of the LISA mission, afterminimizing spurious forces on the test masses, is the ne<strong>ed</strong> to keep the temperatur<strong>ed</strong>istribution throughout each main spacecraft and instrument package as constantas possible. This is essential for two main reasons. One is that mass displacementsdue to local temperature changes within the desir<strong>ed</strong> measurement band, roughly10 −6 Hz to 1 Hz, will change the gravitational force on the test mass and look likea real signal [10, 11]. The other is that a variation in the temperature differencebetween the two sides of the housing in a free mass sensor will cause a differencein the thermal radiation pressure force on the two sides of the test mass, and thusgive a spurious acceleration.The thermal design of the LISA spacecraft can be understood on the basisof the following rough model. The main spacecraft will have some thermal timeconstant for responding to changes in the incident solar flux or in the electrical


Description of the LISA mission 121power dissipat<strong>ed</strong> in its equipment. The Y-shap<strong>ed</strong> thermal shield is mount<strong>ed</strong>from the main spacecraft by low-thermal-conductivity flexural strips or stress<strong>ed</strong>fibreglass bands, and forms a second passive thermal isolation stage. The thirdstage is the support cylinder around each optical assembly and the fourth stageis the optical bench that contains the free mass sensor. Since low-thermalconductivitysupports are us<strong>ed</strong> between each successive stage, and with lowinfrar<strong>ed</strong> emissivity coatings on as much of the surfaces as possible, quite longthermal time constants can be achiev<strong>ed</strong> for each stage.With the above type of multistage passive thermal isolation approach, quitelow levels of temperature fluctuations can be achiev<strong>ed</strong> at frequencies of 0.1 mHzand higher. But there are practical limits to how long the various thermal timeconstants can be made. To achieve high thermal stability at lower frequencies,some active thermal control will be ne<strong>ed</strong><strong>ed</strong>. Even a fairly crude servosystem forthe temperature of the Y-shap<strong>ed</strong> thermal shield would help considerably, and someactive control of the temperature of the main spacecraft probably will be desirablealso. Such active temperature control systems may be add<strong>ed</strong> to the baselinemission design later to meet a goal of improv<strong>ed</strong> antenna sensitivity at frequenciesbelow 0.1 mHz, but are not includ<strong>ed</strong> in the present mission requirements.A fundamental feature of the mission design is the high stability of the solarenergy dissipation in the spacecraft. In an ecliptic-bas<strong>ed</strong> reference frame thatrotates once per year about the pole of the ecliptic, the three spacecraft forma nearly equilateral triangle in a plane that is tipp<strong>ed</strong> at 60 ◦ to the ecliptic, asdiscuss<strong>ed</strong> earlier. Thus, the direction to the Sun makes a nearly constant 30 ◦angle with the normal to the sunward side of each spacecraft, where the solarcells providing power for the spacecraft are locat<strong>ed</strong>. Because of the stable solarillumination geometry, only fluctuations in the intrinsic solar intensity and inthe electrical power dissipation in the various parts of the spacecraft producesignificant temperature variations. Measures such as keeping the microwavetransmitters operating at the same power level whether data is being transmitt<strong>ed</strong>or not are taken to r<strong>ed</strong>uce thermal changes. Also, the power ne<strong>ed</strong><strong>ed</strong> in the opticalassemblies and particularly on the optical benches is kept as low as possible, andis highly regulat<strong>ed</strong>.10.1.3 Optics and interferometry systemThe present parameters assum<strong>ed</strong> for the LISA optical system are the use of 1 Wof output power from cw NdYAG lasers operating at 1.064 µm wavelength, andof 30 cm diameter telescopes to transmit and receive the laser beams betweendifferent spacecraft. Perhaps the most impressive thing about the mission is thatthese fairly modest parameter values lead to roughly 2 × 10 8 photoelectrons s −1being detect<strong>ed</strong> at the far end of a 5000 000 km arm, which would correspondto 2 × 10 −21 precision in measuring changes in the arm length in one second ifthere were no other sources of error. Since almost all expect<strong>ed</strong> types of LISAgravitational-wave sources can be observ<strong>ed</strong> for a year or more, the potential


122 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionmeasurement accuracy clearly is extremely high.Each optical assembly has two lasers for r<strong>ed</strong>undancy, and a switchingmechanism for switching between them. Each laser is capable of 2 W outputover a mean lifetime of several years, but is operat<strong>ed</strong> at1Wtoextend the lifetimeby a substantial factor. A schematic drawing of the optical bench for one opticalassembly was shown earlier in figure 10.5. Light from the active laser is takenby a single mode fibre onto the optical bench, and then form<strong>ed</strong> into a beam ofroughly 5 mm diameter. A few per cent of the beam is split off for other purposes,and then a beamsplitter sends almost all of the power out to the telescope, whereit is transmitt<strong>ed</strong> to the matching optical assembly on the other end of the arm. Thereturn beam from the laser in the distant optical assembly is sent by the telescopeto the test mass, where it is reflect<strong>ed</strong> from the face of the test mass, and then beatagainst a small amount of power from the local laser.The beat signals detect<strong>ed</strong> in the six optical assemblies are the main data us<strong>ed</strong>to make the gravitational-wave measurements. In the present baseline scheme,one laser in spacecraft SC-1 is consider<strong>ed</strong> the master laser and lock<strong>ed</strong> to a stablereference cavity on its optical bench. The other laser in SC-1 is phase lock<strong>ed</strong>to the first with a small offset frequency of perhaps a couple of kilohertz. Thetwo laser beams are transmitt<strong>ed</strong> along the adjacent arms of the triangle to SC-2and SC-3, where they are reflect<strong>ed</strong> from the test masses in the receiving opticalassemblies. The local lasers are then phase lock<strong>ed</strong> to the receiv<strong>ed</strong> beams, alsowith small frequency offsets, and their beams are sent back to SC-1. Finally, thesecond lasers in SC-2 and SC-3 are offset lock<strong>ed</strong> to the first lasers, their beams aresent over the third arm of the triangle, and the beats between them are record<strong>ed</strong>on both ends of the side, after reflection from the test masses.The frequencies of the beat signals can range from well under 1 MHz toroughly 15 MHz, depending on the exact orbits of the spacecraft, or rather theorbits of the freely floating test masses in them. In one possible measurementscheme, the signals are beat down to a convenient frequency range for precisephase measurements, such as perhaps 10 kHz, using outputs from frequencysynthesizers. The drive frequencies for the synthesizers come from ultra-stablecrystal oscillators (USOs) on each spacecraft. In a second measurement scheme,the signals are first sampl<strong>ed</strong> at a high rate such as 40 MHz, and the results areanalys<strong>ed</strong> in the software. In this case, the sampling times are controll<strong>ed</strong> by theUSOs.If only the signals from the two arms adjacent to SC-1 are consider<strong>ed</strong>,the antenna is like a single groundbas<strong>ed</strong> detector, with changes in the armlength difference being the quantity of interest. However, it does not appearpractical to keep the difference in arm length nearly constant, since this wouldrequire applying quite large forces to the test masses to overcome the changinggravitational forces on them due to the Sun and to the planets. If such large forceswere appli<strong>ed</strong>, it would be very difficult to avoid noise in those forces withinthe frequency band of interest from giving undesir<strong>ed</strong> changes that could not b<strong>ed</strong>istinguish<strong>ed</strong> from real signals. To avoid this problem, the test masses are left


Description of the LISA mission 123nearly free, except for small differential appli<strong>ed</strong> forces on the two test masses ina given spacecraft that are necessary to compensate for the difference in spuriousforces on them at dc and at frequencies below the useful measurement band.Because of the up to 1% difference in arm lengths expect<strong>ed</strong>, the phasenoise in the master laser has to be correct<strong>ed</strong> for too high accuracy. The originalsuggestion that this was possible was made by Faller in the abstract for a 1981conference [3]. The basic idea is to use the apparent changes in the sum ofthe lengths of the two arms to estimate the laser phase noise, and then to applythe corresponding correction to the measur<strong>ed</strong> difference in the arm lengths.Approximate algorithms for doing this were publish<strong>ed</strong> by Giampieri et al in1996 [12]. More recently, Armstrong, Tinto and Estabrook [13, 14] have givenwhat are believ<strong>ed</strong> to be rigorously correct algorithims, and it is plann<strong>ed</strong> to usethese during the LISA mission.So far, only the use of the signals from the two detectors on SC-1 has beendiscuss<strong>ed</strong>. In addition, the signal from one of the detectors on each of SC-2 andSC-3 is us<strong>ed</strong> to phase lock a laser to the receiv<strong>ed</strong> laser beam from SC-1. So thesignals measur<strong>ed</strong> on SC-2 and SC-3 from the beams sent between them can becombin<strong>ed</strong> to give the changes in length of the third arm. As discuss<strong>ed</strong> by Cutler[15], the length of the third arm minus the average of the lengths of the othertwo is an observable that gives the other polarization from the one determin<strong>ed</strong> bythe difference in lengths of the first two. Thus, having the measurements overthe third arm gives a valuable addition to the scientific information that can beobtain<strong>ed</strong> from a mission like LISA.In the recent papers of Armstrong, Tinto and Estabrook [13,14], the analysisis done in a way that assumes all the lasers are running independently, presumablystabiliz<strong>ed</strong> by locking to their own stable reference cavities, but not lock<strong>ed</strong> toeach other. This appears to put some additional requirements on the missionmeasurement system, but they point out that there is additional information inthe resulting signals at the shorter gravitational wavelengths that is not availablewith the current baseline measurement system. It appears that this may beequivalent to the information that would be obtain<strong>ed</strong> from running LISA as aSagnac interferometer, but this possibility has not yet been consider<strong>ed</strong> in detail.In addition to providing the second polarization if all is going well, havingthe capability of making measurements over the third arm has another veryimportant benefit for the LISA mission. With two fully functional opticalassemblies on each of the three spacecraft, if something in one of the six were notoperating properly, the antenna would still provide two-arm gravitational-wav<strong>ed</strong>ata with the full plann<strong>ed</strong> sensitivity. Even if a second optical assembly were outof operation, provid<strong>ed</strong> the two were not on the same spacecraft, the two-arm datacould still be obtain<strong>ed</strong>. Thus, the third arm capability provides an essential levelof r<strong>ed</strong>undancy for the mission.In addition to the shot noise in the photocurrents from the receiv<strong>ed</strong> laserbeams, other noise sources in measuring changes in the arm length differencesalso have to be consider<strong>ed</strong>. One such source is fluctuations in the attitude of


124 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missioneach spacecraft. If the transmitt<strong>ed</strong> beam gave a perfectly spherical wavefront inthe far field, attitude changes for the transmitting spacecraft would not give phasechanges in the receiv<strong>ed</strong> signal at the distant spacecraft. However, the combinationof diffraction due to the finite size transmitting aperture plus imperfections inthe optical system cause the wavefronts at the distant spacecraft to be somewhatdistort<strong>ed</strong>. Thus, the attitude of the spacecraft and any other sources of jitter in thepointing of the transmitt<strong>ed</strong> beam have to be controll<strong>ed</strong> quite closely.There are two methods under consideration for measuring jitter in thespacecraft attitude. In one, perhaps 10% of the receiv<strong>ed</strong> light from the distantspacecraft is pick<strong>ed</strong> off and focus<strong>ed</strong> on a CCD array or quadrant diode via a fairlylong effective focal length optical system. Changes in attitude then give changesin the position of the focal spot on the detector, which are us<strong>ed</strong> in a servo systemto control the spacecraft attitude, along with the similar signals from the secondoptical assembly on the spacecraft. In the second approach, all of the light goesto the main detector for measuring changes in the arm length, but the detectoris replac<strong>ed</strong> by a quadrant device. If the spacecraft tips slightly with respect tothe receiv<strong>ed</strong> wavefronts, the differences in phase of the four detect<strong>ed</strong> signals willchange, and these changes are us<strong>ed</strong> as the inputs to the attitude control system.Even if the attitude jitter is made small, there still is a ne<strong>ed</strong> to set themean beam pointing direction carefully. Diffraction plus a defocus of thetransmitt<strong>ed</strong> beam would result in the phase of the receiv<strong>ed</strong> wavefront varyingonly quadratically with the angular offset from the optical axis of the transmittingsystem. However, non-axisymmetric defects in the transmitt<strong>ed</strong> wavefronts canmake the change in the receiv<strong>ed</strong> phase vary linearly with the attitude change,even on what otherwise would have been the optical axis. To avoid this increas<strong>ed</strong>sensitivity to beam pointing, the pointing along each axis of each optical assemblyis modulat<strong>ed</strong> at a known frequency, and the resulting apparent changes in the armlength differences at the modulation frequencies and their second harmonics ar<strong>ed</strong>etect<strong>ed</strong>. This information permits the outputs at the modulation frequencies tobe minimiz<strong>ed</strong> by small offsets in the dc pointing directions, which corresponds tohaving just the quadratic variations in the receiv<strong>ed</strong> phase due to attitude jitter.As a measure of the remaining requirement on pointing jitter, a convenienttest case is to assume only astigmatic error in the transmitt<strong>ed</strong> wavefronts withan error of a tenth of a wavelength rms. For this example, if the distance errordue to pointing is allow<strong>ed</strong> to be equal to that from shot noise, the errors in th<strong>ed</strong>c pointing offset and in the pointing jitter can be as large as 10 milliarcsec and4 milliarcsec/rtHz, respectively. (Here and later, /rtHz stands for ‘per square rootHz’, and the given error or noise is the spectral amplitude of the error, which isthe square root of the power spectral density.) These error allocations are eachabout three times larger than those given in section 3.1.8 of [6], since the errorallocation in that case was assum<strong>ed</strong> to be only 10% of the error due to shot noise.Further information on the optical path error allocation budget for LISAis given in sections 4.2.1 and 4.2.2 of [6]. The total error allocation for themeasurement of the difference in the round-trip path lengths for two arms of


Description of the LISA mission 125the antenna is 40 nm/rtHz. With only shot noise consider<strong>ed</strong>, the correspondingvalue would be 22 nm/rtHz, and with beam pointing jitter includ<strong>ed</strong> also, thevalue becomes 30 nm/rtHz. Other error sources consider<strong>ed</strong> explicitly are asfollows: residual laser phase noise after the correction proc<strong>ed</strong>ure discuss<strong>ed</strong> earlieris appli<strong>ed</strong>; noise in the ultra-stable oscillators after a similar correction proc<strong>ed</strong>ure;noise in the laser phase measurements and phase locks; and scatter<strong>ed</strong> light effects.The 40 nm/rtHz total error allocation for the difference in round-trip paths for twoarms corresponds to 20 nm/rtHz for determining the difference in lengths for thetwo arms.10.1.4 Free mass sensorsHistorically, the first free mass sensor was develop<strong>ed</strong> jointly by the Johns HopkinsUniversity Appli<strong>ed</strong> Physics Laboratory and by Stanford University, and was flownon the TRIAD satellite in 1972 [16]. A spherical test mass was contain<strong>ed</strong> in aspherical cavity with three opposing pairs of capacitive electrodes on the insideof the housing for sensing the relative position of the test mass. Whenever theatmospheric drag caus<strong>ed</strong> the housing to move a few millimetres with respect tothe test mass, puls<strong>ed</strong> thrusters on the satellite fir<strong>ed</strong> to keep it centr<strong>ed</strong> on averageon the test mass. Care was taken to keep forces on the test mass other than thos<strong>ed</strong>ue to external gravitational fields as small as possible. Thus, the orbit of thesatellite was nearly drag-free, and was much more pr<strong>ed</strong>ictable than normal. Thetheory of such drag-free systems had been given earlier in 1964 by Lange [17].There is a close connection between free mass sensors and the highperformanceforce-rebalance accelerometers develop<strong>ed</strong> over the last three decadesby the Office National d’Etudes et de Recherches Aerospatiales (ONERA) in Parisfor flight on various missions. The basic approach in the accelerometers is tomeasure the position of a free or nearly free test mass inside a housing by meansof capacitive electrodes on the inside of the housing. Forces are then appli<strong>ed</strong> tothe test mass by means of voltages on the electrodes to keep the test mass centr<strong>ed</strong>in the housing. The requir<strong>ed</strong> voltages are measures of the accelerations along thethree perpendicular axes.The first such accelerometer design<strong>ed</strong> and built by ONERA was flown from1975 to 1979 on the French CASTOR-D5B satellite. It had a spherical test mass.However, later ONERA-design<strong>ed</strong> accelerometers have us<strong>ed</strong> test masses in theform of rectangular parallelepip<strong>ed</strong>s. The first such design, call<strong>ed</strong> the GRADIOaccelerometer, was for possible use in a propos<strong>ed</strong> gravity gradiometer mission(ARISTOTLES) to map the Earth’s gravity field, and has been the basis for anumber of later designs.The test mass for the GRADIO accelerometer is 4 cm by 4 cm by 1 cm indimensions. The material us<strong>ed</strong> for the test mass is a Pt–Rh alloy with a densityof about 20 g cm −3 . The housing is made of ultra-low expansion glass (ULE),with a gold coating on the inside that is carefully pattern<strong>ed</strong> to form the capacitiveplate electrodes. The gaps between the plates are recess<strong>ed</strong> and kept very small


126 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionto r<strong>ed</strong>uce the charge buildup on the underlying material. The position sensing isdone with 100 kHz transformer bridge circuits. Extra pairs of plates are includ<strong>ed</strong>to permit relative rotation of the test mass with respect to the housing to be sens<strong>ed</strong>and controll<strong>ed</strong>.The gaps between the capacitive plates and the test mass is 30 µm for thelarge square faces and 300 µm for the smaller rectangular ones. With the testmass horizontal, the factor four smaller vertical dimension and the 30 µm gappermit the test mass to be electrostatically suspend<strong>ed</strong> for laboratory tests. Theentire accelerometer is contain<strong>ed</strong> in a vacuum enclosure. Extensive tests havebeen carri<strong>ed</strong> out at ONERA with two GRADIO accelerometers mount<strong>ed</strong> on asingle pendular support platform in a special sub-basement room to r<strong>ed</strong>uce theeffects of tilts in the floor. The results for the differential horizontal accelerationshave been very encouraging, but are still several orders of magnitude worse thanthe theoretical sensitivity of roughly 10 −12 ms −2 /rtHz from a few millihertz to1 Hz in a quiet and constant temperature zero-g environment.A later but lower sensitivity version of the GRADIO accelerometer has beenflown a few times on the space shuttle, and other versions will fly soon on othermissions. These include the German–French CHAMP mission sch<strong>ed</strong>ul<strong>ed</strong> forlaunch in 2000, a USA–German mission call<strong>ed</strong> GRACE sch<strong>ed</strong>ul<strong>ed</strong> for a <strong>2001</strong>launch, and a later ESA mission call<strong>ed</strong> GOCE. GRACE is a satellite-to-satellitetracking mission that will map time variations in the Earth’s gravity field over afive-year period with high accuracy, and an accelerometer on each satellite willmonitor drag and other non-gravitational forces on the satellites. GOCE is agravity gradiometer mission to map very short wavelength spatial structure in thegravity field, using the differential accelerations between pairs of accelerometersin the same spacecraft.For the free mass sensors for the LISA mission, the requirements are quit<strong>ed</strong>ifferent from those for accelerometers. The purpose of the accelerometers isto measure the non-gravitational accelerations of the spacecraft to which theyare attach<strong>ed</strong>, and thus the readout sensitivity has to be high. For the GRADIOaccelerometer, with 300 µm gaps along the two most sensitive axes, the positionmeasurement sensitivity is 6 × 10 −12 m/rtHz at frequencies above 5 mHz.However, the small gaps lead to increases in some of the time-varying spuriousaccelerations of the test mass, as discuss<strong>ed</strong> later. For LISA, it turns out thatthe necessary sensitivity of the capacitive position measurements is only about10 −9 m/rtHz, so gaps of 2 mm or larger and different values of some of the othermeasurement parameters can be us<strong>ed</strong> to r<strong>ed</strong>uce the spurious accelerations.An initial ONERA study of a propos<strong>ed</strong> free mass sensor for LISA, call<strong>ed</strong>CAESAR, was publish<strong>ed</strong> by Touboul, Rodrigues and Le Clerc in 1996 [18]. Thetest mass is a cube 4 cm on a side, made of a 10% Pt–90% Au alloy to achieve lowmagnetic susceptibility and a density of 20 g cm −3 (see figures 10.6 and 10.7).The gaps assum<strong>ed</strong> were 4 mm along the most sensitive axis and 1 mm for theother two axes. An error budget for the sensor was develop<strong>ed</strong>, bas<strong>ed</strong> on the highthermal stability expect<strong>ed</strong> for the optical bench during the LISA mission.


Description of the LISA mission 127Figure 10.6. Early ONERA design of LISA free mass sensor call<strong>ed</strong> CAESAR. Up to fourcapacitor plates on each face of the housing around the test mass allow the relative positionand angular orientation to be measur<strong>ed</strong> accurately. Electrical forces also can be appli<strong>ed</strong>, ifdesir<strong>ed</strong>.After completion of the initial CAESAR design, additional studies of freemass sensor designs and of the LISA requirements have been carri<strong>ed</strong> out,influenc<strong>ed</strong> strongly by the CAESAR design. A number of the factors influencingthe design were discuss<strong>ed</strong> by Speake [19] and by Vitale and Speake [20]. A designwith a considerably different geometry for the capacitor plates was investigat<strong>ed</strong> byJosselin, Rodrigues and Touboul [21], and has since been construct<strong>ed</strong> and test<strong>ed</strong>in the laboratory and the LISA free mass sensor requirements were simplifi<strong>ed</strong> andsomewhat relax<strong>ed</strong> [6].One current version of the error allocation budget for the free mass sensoris describ<strong>ed</strong> in section 4.2.3 of [6]. The total acceleration error allow<strong>ed</strong> forone sensor is 3 × 10 −15 ms −2 /rtHz over the frequency range 0.1–30 mHz.The six largest allocations to individual error sources are as follows: thermaldistortion of the main spacecraft; temperature difference variations across the


128 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionFigure 10.7. Ultra-low expansion glass housing for the proof (test) mass in the earlyCAESAR design.test mass housing; electrical force on the charge on the test mass; Lorentzforce on the charg<strong>ed</strong> test mass from the fluctuating interplanetary magnetic field;residual gas impacts on the test mass; and fluctuating forces due to electrical fielddissipation in the test mass housing. These sources are each allocat<strong>ed</strong> an error of1 × 10 −15 ms −2 /rtHz.In view of the performance level desir<strong>ed</strong> for the LISA free mass sensors, it isplann<strong>ed</strong> to have a technology demonstration flight for them at the earliest possibl<strong>ed</strong>ate. Efforts to arrange for such a flight currently are under way both in Europeand the USA. Two sensors would be flown on a thermally isolat<strong>ed</strong> optical benchon a small spacecraft, with an interferometer to measure changes in the separationof the two test masses. To keep the rate of charging of the test masses comparablewith what it would be well away from the Earth, and to r<strong>ed</strong>uce the effect of theEarth’s gravity gradient, a perigee altitude of 10 000 km or higher is desir<strong>ed</strong>. Tor<strong>ed</strong>uce the cost, a tentative goal of demonstrating only 3 × 10 −14 ms −2 /rtHzperformance from 1–10 mHz currently is assum<strong>ed</strong>. It is believ<strong>ed</strong> that cautious


Description of the LISA mission 129engineering extrapolation from tests at this level plus some in-flight tests withintentionally increas<strong>ed</strong> disturbances will provide a sound basis for proce<strong>ed</strong>ingwith the LISA mission.10.1.5 Micronewton thrustersThe main non-gravitational force on a LISA spacecraft is expect<strong>ed</strong> to be fromthe solar radiation pressure and to have a magnitude of about 20 µN. If notcompensat<strong>ed</strong> for, it would cause an acceleration of the spacecraft of roughly10 −7 ms −2 . Since the test masses are shield<strong>ed</strong> from this source of acceleration, itis necessary to apply force to the spacecraft to keep it from moving with respectto them. The spectral amplitude of the fractional fluctuations in the solar pressureforce [22] over the 0.1–10 mHz frequency range is approximately1.3 × 10 −3 ( f/1 mHz) −1/3 .Comparable fluctuations in force but with a more r<strong>ed</strong>den<strong>ed</strong> spectrum and amuch lower dc level will be present from the solar wind.The type of thrusters that are plann<strong>ed</strong> for use on the LISA mission are fieldemission electric propulsion (FEEP) thrusters. They operate by accelerating ionsthrough a potential drop of 5–10 kV and ejecting them to provide thrust. Theejection velocity is roughly 60–100 km s −1 , corresponding to a specific impulseof 6000 to 10 000 s. In view of this high specific impulse and the low thrust levelne<strong>ed</strong><strong>ed</strong>, the fuel requir<strong>ed</strong> per thruster for a ten year extend<strong>ed</strong> mission lifetime isonly a few grams.Historically, most of the development of FEEP thrusters has been bas<strong>ed</strong>on the use of Cs ions. This work has been done mainly at the EuropeanSpace Research and Technology Centre in Noordwijk, The Netherlands and atCentrospazio in Pisa, Italy (see section 7.3 of [6]).A schematic drawing of a thruster is shown in figure 10.8. In each thruster,liquid Cs metal at a temperature somewhat above the melting point of 29 ◦ Cis contain<strong>ed</strong> in a small reservoir. It is drawn by capillary forces through anarrow channel between two polish<strong>ed</strong> metal plates spac<strong>ed</strong> 1 or 2 µm apart. Theaccelerating voltage is appli<strong>ed</strong> by a plate with a slot in it, locat<strong>ed</strong> at the outer <strong>ed</strong>geof the channel.The high field at the surface of the Cs metal causes an instability, and Taylorcones roughly a micrometre in diameter and a few micrometres apart form on thesurface. The tips of these cones are very sharp, and the high field around the tipscauses Cs ions to be drawn out by field emission and accelerat<strong>ed</strong> away in a beamperhaps 30 ◦ wide. The one substantial drawback to the use of Cs ions is thatany water vapour that is present will react with the Cs to form CsOH. Thus, thethrusters are kept in vacuum containers roughly 5 cm in dimensions until in spaceand ready for use. A spring-load<strong>ed</strong> cover with an O-ring seal is then releas<strong>ed</strong>.An alternative to Cs is the use of In. Field emission of In ions for theneutralization of positive charge on spacecraft has been pioneer<strong>ed</strong> by the Austrian


130 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionFigure 10.8. Schematic drawing of cesium Field Emission Electric Propulsion (FEEP)thruster. The thrust can be controll<strong>ed</strong> by changing the accelerator voltage smoothly. A setof six or more thrusters is us<strong>ed</strong> to control the spacecraft position and attitude.Research Centre Seibersdorf, and such devices have been flown on severalmissions. The geometry us<strong>ed</strong> consists of a sharpen<strong>ed</strong> tungsten ne<strong>ed</strong>le with a tipa few micrometres in diameter mount<strong>ed</strong> in the centre of a heat<strong>ed</strong> In reservoir.Capillary forces bring the In to the tip, and again an accelerator plate applies ahigh voltage. A single Taylor cone is form<strong>ed</strong>, and the In ions are produc<strong>ed</strong> byfield emission from the sharp tip of the cone.If such devices are us<strong>ed</strong> as thrusters, there is a limit to the thrust achievableper tungsten ne<strong>ed</strong>le. If too much current is drawn, small droplets form andgive much less thrust per unit mass than the ions do. However, it has beendemonstrat<strong>ed</strong> that a number of ne<strong>ed</strong>les can be operat<strong>ed</strong> in parallel to meet thequite low thrust requirements of the LISA mission during its operational phase.The results of research on such thrusters have been report<strong>ed</strong> by the Seibersdorfgroup [23]. The possibility of using In instead of Cs with the channel capillaryflow approach is now being investigat<strong>ed</strong> by Centrospazio, but no decision hasbeen made on the overall merits of the two materials with this approach.The advantages of FEEP thrusters, in addition to their high specific impulse,is that they can run continuously and be controll<strong>ed</strong> easily by small changesin the acceleration voltage. Work on their further development and testing isbeing support<strong>ed</strong> by ESA because of their potential for use on a number of futuremissions. If In ions are us<strong>ed</strong>, the thrusters have to run hotter because of the 156 ◦ Cmelting temperature for In. However, the power requir<strong>ed</strong> is still low, and doesnot appear to be a problem. To control both attitude and translation and allowsufficient r<strong>ed</strong>undancy, probably between 12 and 24 individual thrusters pointingin different directions will be includ<strong>ed</strong> per spacecraft.


Description of the LISA mission 13110.1.6 Mission scenarioA single launch with the Delta II launch vehicle is assum<strong>ed</strong> in current studiesof the LISA mission [7]. Each of the three spacecraft has a thin cylindricalpropulsion module attach<strong>ed</strong> to it. The three composite vehicles are stack<strong>ed</strong> on topof each other for launch, and then separate from the launch vehicle and from eachother a short time later. Since the spacecraft will be separat<strong>ed</strong> by 5000 000 kmand have different orbit planes in their final configuration, they travel separatelyand each has its own onboard guidance system.It is currently plann<strong>ed</strong> to use solar electric propulsion for the propulsionmodules. The top area of the main spacecraft is cover<strong>ed</strong> with solar cells, andenough power is generat<strong>ed</strong> to provide 15 or 20 mN of thrust with either Hall effector another type of Xe ion thruster. The time necessary for the three spacecraft toreach their proper locations and velocities about 20 ◦ behind the Earth in orbitabout the Sun is 14 or 15 months.After the nominal orbits have been achiev<strong>ed</strong>, the spacecraft will be track<strong>ed</strong>for a week or two with NASA’s Deep Space Tracking Network. Then, any desir<strong>ed</strong>minor corrections to the orbits can be made. The next step is separation of thepropulsion modules from the final spacecraft. After that, the orbits will be almostcompletely gravitational, with each spacecraft servocontroll<strong>ed</strong> to follow either theaverage of the two test masses inside it or just one of the two. The difference ofthe spurious accelerations of the two test masses will be roughly 10 −10 ms −2 orless, and can be correct<strong>ed</strong> for by applying weak and stable electrical forces to thetest masses at frequencies below the measurement range via the capacitor plates.The next step is for each spacecraft to acquire the other two optically. Startrackers aboard each spacecraft give the attitude to a few arcseconds, and thetransmitt<strong>ed</strong> laser beams are defocus<strong>ed</strong> by roughly a factor ten so that they can b<strong>ed</strong>etect<strong>ed</strong> on the other end of an arm even with relatively poor beam pointing. Thereceiv<strong>ed</strong> laser beam normally is bright enough to look like about a magnitude-3star, so there will be a substantial brightness level even with some defocusing. Ifnecessary, an angular scan pattern on the transmitt<strong>ed</strong> beam could be us<strong>ed</strong> to findthe beam, since the stability of the star tracker will be considerably better than itsabsolute accuracy.As soon as a receiving spacecraft finds the beam, the output from the CCDarray or quadrant diode in its angle tracker can be us<strong>ed</strong> to lock the attitude to th<strong>ed</strong>irection to the distant spacecraft, and the defocus in the transmitt<strong>ed</strong> beam canbe remov<strong>ed</strong> to give a high S/N ratio. The local laser then is turn<strong>ed</strong> on, and itsfrequency adjust<strong>ed</strong> to give a fairly low beat frequency with the receiv<strong>ed</strong> beam.The resulting heterodyne signal back at the original transmitting end provides ahigher S/N ratio, and makes detecting the defocus<strong>ed</strong> return beam easier. After thebeam pointing signals are lock<strong>ed</strong> in and the focus is correct<strong>ed</strong> on both ends ofeach arm, the system is ready for operation.Even with careful phase locking of the master laser to a stable cavity, orpossibly the locking of more lasers, the phase noise will still be significant up to


132 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionfrequencies of a few hundr<strong>ed</strong> hertz to 1 kHz. Also, the amplitude of the phasenoise at frequencies near 1 Hz and below will be large. To avoid aliasing thehigher frequency noise into the band of interest, below perhaps 3 Hz, it may b<strong>ed</strong>esirable to make the phase measurements at a rate approaching a kilohertz, andthen filter the data in software. For low-frequency noise, it is plann<strong>ed</strong> to combineall the signals at a single spacecraft and perform the algorithm for correcting forlaser phase noise there. The resulting signals then can be compress<strong>ed</strong> before beingsent to the ground.The number and amplitude of mechanical motions aboard the spacecraft arekept as small as possible to r<strong>ed</strong>uce gravitational attraction changes and excitationof vibrations. Roughly 30 cm diameter X-band antennae probably will be us<strong>ed</strong> tosend data down to 34 m Deep Space Network antennae on the ground, and willhave to be repoint<strong>ed</strong> about once a week. Step changes in the angle between theaxes of the two optical assemblies probably will be made at the same time withcoarse adjustment mechanisms, but smooth adjustments over roughly a range of5 arcmin are requir<strong>ed</strong> between the steps. It is expect<strong>ed</strong> that offsets in the position,velocity, and dc acceleration of the test masses will have to be solv<strong>ed</strong> for at thetimes of the steps.The nominal mission lifetime after the antenna geometry is establish<strong>ed</strong>probably will be three years. However, there are no expendables requir<strong>ed</strong> exceptfor the In or Cs fuel for the micronewton thrusters, and that will be made adequatefor a ten year or longer lifetime. Thus, a rolling three year approval periodhopefully can be establish<strong>ed</strong>, provid<strong>ed</strong> that the antenna continues to operateproperly.10.2 Expect<strong>ed</strong> gravitational-wave results from LISA10.2.1 LISA sensitivity and galactic sourcesThe two types of noise sources for the LISA antenna have been discuss<strong>ed</strong>earlier. The one that dominates at frequencies below about 3 mHz is spuriousaccelerations of the test masses, and the level of error allocat<strong>ed</strong> for frequenciesf between 0.1 and 30 mHz is 3 × 10 −15 ms −2 /rtHz for each free mass sensor.Considering only two arms of the antenna for simplicity, and adding the errorsfrom the four sensors quadratically, the resulting noise level for the difference in(one-way) length of the two arms is(L2 − L1) = 1.520 × 10 −16 (1Hz/f ) 2 m/rtHz.This should be combin<strong>ed</strong> quadratically with the error allocation of 2 ×10 −11 m/rtHz for measuring the difference in the distances between the testmasses along the two arms.For low frequencies, where the wavelength is long compar<strong>ed</strong> with the armlength, the response of the antenna to a gravitational wave [24] can be given


Expect<strong>ed</strong> gravitational-wave results from LISA 133Figure 10.9. LISA sensitivity and galactic sources. The instrumental and binary confusionthreshold sensitivity curves are shown for one year of observations and an S/N ratio of 5.Estimat<strong>ed</strong> signal strengths and frequencies for galactic binaries are indicat<strong>ed</strong> also.simply for the optimum direction of propagation and polarization of the wave.The change in L2 − L1 is just the amplitude h of the wave times the cosineof the angle between the two arms. Averaging over the direction of propagationand the polarization gives a factor 5 0.5 lower signal amplitude. However, at higherfrequencies, the signal is r<strong>ed</strong>uc<strong>ed</strong> by a factor that depends in a complicat<strong>ed</strong> way onthe direction of propagation and the polarization. The transfer function giving therms value of this r<strong>ed</strong>uction factor, and including the cosine of the angle betweenthe two arms, has been given by Schilling [25]. It was us<strong>ed</strong> in calculating the LISAthreshold sensitivity, as shown in figure 10.9 and discuss<strong>ed</strong> below. Essentially thesame curve has been calculat<strong>ed</strong> independently by Armstrong, Tinto and Estabrook[13], but for a slightly different choice of arm length.For almost all of the gravitational-wave signals LISA will see, the frequency,amplitude and polarization of the wave will be very stable, and will not changesubstantially over a few years of observation. Thus, it is desirable to plot theLISA threshold sensitivity as a curve such that there is a good chance of seeinga source in a reasonable observing time if its rms amplitude lies above the curve.To accomplish this, the threshold sensitivity curves in figures 10.9 and 10.11 areplott<strong>ed</strong> as the rms amplitude of a stable signal ne<strong>ed</strong><strong>ed</strong> in order to have a S/Nratio of five for one year of observation, when averag<strong>ed</strong> over source directionand polarization. The S/N ratio of five is chosen because almost all the sourceswill be unknown from optical observations, and roughly this S/N ratio is ne<strong>ed</strong><strong>ed</strong>to determine the reality of sources with unknown frequencies and with only


134 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionstatistical information on their directions in the sky. When averag<strong>ed</strong> over a year,the LISA antenna response is remarkably independent of the source direction inthe sky [26], so no allowance for the ecliptic latitude or longitude of the source isne<strong>ed</strong><strong>ed</strong> in using the LISA threshold sensitivity curve.The dominant type of signal LISA will see is gravitational <strong>waves</strong> from binarystars systems in our galaxy. Mironowskii [27] point<strong>ed</strong> out in 1966 that therewould be roughly 10 8 W UMa type binaries in the galaxy giving substantialsignals. These binaries consist of two main sequence stars so close together thattheir Roche lobes are in contact, and material can flow between them. Theirfrequencies lie in a fairly narrow band near 0.1 mHz. With so many sources atlow frequencies, it is clear that even many years of observations would not permitindividual signals to be resolv<strong>ed</strong>, except for a few that happen to be close to theSun and thus have high signal strengths. There are even larger numbers of noncontactbinaries compos<strong>ed</strong> of normal stars at lower frequencies.In 1984 Iben [28] point<strong>ed</strong> out the expect<strong>ed</strong> existence of a very large numberof close white dwarf binaries (CWDBs) in our galaxy, and the strength of signalsexpect<strong>ed</strong> from them. Because of their radii being only about 10 000 km, they cangive gravitational-wave signals at frequencies up to roughly 30 mHz. However,they had not been observ<strong>ed</strong> directly until the last few years, and their space densityin the galaxy is only poorly known. On the other hand, since the orbits of theinteresting CWDBs evolve mainly by gravitational radiation, the number per hertzwill decrease as the 11/3 power of the frequency, and thus the frequency abovewhich most of them can be resolv<strong>ed</strong> is only fairly weakly dependent on the spac<strong>ed</strong>ensity.Recent observations by Marsh and collaborators [29–32] have found anumber of examples of CWDBs, so their abundance must be moderately closeto the astrophysical estimates. However, the criteria us<strong>ed</strong> in selecting their targetsof observation make it difficult to derive a space density. Bas<strong>ed</strong> on the bestastrophysical studies available at present [33,34], it is expect<strong>ed</strong> that most CWDBswill be resolv<strong>ed</strong> at frequencies above roughly 3 mHz, even for those near thegalactic centre. The signals are strong enough so that the frequencies of anddirections to a few thousand such sources will be determin<strong>ed</strong> by LISA.The mean strength of signals from CWDBs at the galactic center as afunction of frequency is shown in figure 10.9 as a solid curve. It hookssharply upward at about 15 mHz because the majority of the CWDBs below thatfrequency contain either one or two He white dwarfs, and coalesce just above15 mHz. This leaves only binaries containing mainly white dwarfs compos<strong>ed</strong>almost completely of carbon and/or oxygen, which are call<strong>ed</strong> CO white dwarfs.The CO white dwarfs are roughly a factor two more massive and are smaller inradius, so they can reach somewhat higher frequencies before coalescing.Above and below the curve for the galactic centre are dot-dash<strong>ed</strong> curveslabell<strong>ed</strong> 5% and 95%. The 5% curve shows the mean strength of the signals fromCWDBs at a distance from the Earth such that 5% of all the ones in the galaxyare closer, and the definition for the 95% curve is similar. Thus roughly 90%


Expect<strong>ed</strong> gravitational-wave results from LISA 135of the CWDBs in the galaxy will give signal strengths in the region between thetwo curves. Since this region is substantially above the LISA threshold sensitivitycurve for S/N = 5, most of these sources will have S/N ratios of 10–50.Several other types of binaries also can contribute signals above theinstrumental sensitivity level for LISA [35–43]. When all of these types areconsider<strong>ed</strong> together, there is some frequency at which the expect<strong>ed</strong> number ofgalactic sources per frequency bin for one year of observations drops below avalue of very roughly 1/3. Below that frequency, very little information canbe obtain<strong>ed</strong> about the scientifically very important extragalactic sources, unlessthey are stronger than the sum of signals from the galactic binaries. Abovethat frequency, some information is lost because of having to fit out the galacticsignals, but some survives [44–46]. The effective binary noise level then drops bya factor of about 3–10, at which point it is coming from all the binaries in othergalaxies throughout most of the universe.A simple calculation shows that the extragalactic contribution from eachshell around our galaxy of a given thickness will give the same averagecontribution out until r<strong>ed</strong>shifts comparable with one, provid<strong>ed</strong> the effects ofgalaxy evolution are not too large [47]. A recent estimate of the effective noiselevel introduc<strong>ed</strong> by all of the galactic and extragalactic binary star systems, butnot allowing for evolution [46], is shown in figure 10.9. It has been adjust<strong>ed</strong> tocorrespond to S/N = 5. This confusion noise curve lies above the instrumentalsensitivity curve for frequencies from about 0.1–3 mHz, and has a major effect ondetermining whether LISA can reliably detect particular sources from one year ofobservations. This affects some galactic sources, as well as extragalactic ones.A few individual known sources also are shown on figure 10.9. Since thefrequencies, as well as directions in the sky, are known for most of these sources,the S/N ratio ne<strong>ed</strong><strong>ed</strong> to detect them is only about two. Thus, the ones somewhatbelow the threshold sensitivity or confusion noise curves probably are detectablefrom one to two years of observations. i Boo is a non-contact binary compos<strong>ed</strong>of normal stars, and WZ Sge is a very short period dwarf nova binary. The fourAm CVn binaries shown are each believ<strong>ed</strong> to consist of an extremely low masshelium dwarf losing mass to a CO dwarf primary [39]. They also are call<strong>ed</strong> heliumcataclysmics, and effects due to such binaries and their possible progenitors havebeen includ<strong>ed</strong> in the recent confusion noise estimates [46]. The final knownbinary shown is 4U 1820-30, an x-ray binary believ<strong>ed</strong> to contain a neutron star orstellar mass black hole.Before discussing the important questions about massive black holes andabout relativity that LISA hopefully will contribute unique new information on,it is useful to say something about how the data will be analys<strong>ed</strong> [48]. First,any large and easily recognizable signals will be fitt<strong>ed</strong> and remov<strong>ed</strong> from th<strong>ed</strong>ata. Next, a convenient data set such as a one year data record will be analys<strong>ed</strong>carefully to determine as many as possible of the galactic binary signals. Thissearch will be easy for frequencies above perhaps 10 mHz, where the signalsare quite well separat<strong>ed</strong>. However, the sidebands on the signals due to rotation


136 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionand motion of the antenna will overlap strongly as the frequency decreases tonear 3 mHz, and a simultaneous fit of all the signals in each roughly 100 or200 cycle/year band is likely to be necessary in order to fit the signals well. Atstill lower frequencies, only the strongest sources whose signals rise substantiallyabove the large number from near the galactic center will be detectable. All ofthe identifi<strong>ed</strong> sources will then be remov<strong>ed</strong> from the data record, before lookingfor the swept frequency signals expect<strong>ed</strong> from extragalactic sources involvingmassive black holes.10.2.2 Origin of massive black holesAn important astrophysical question is how the se<strong>ed</strong> black holes that later grewto be the massive and supermassive black holes observ<strong>ed</strong> today were form<strong>ed</strong>. Toaid in identifying different mass ranges for black holes, we will refer to thosefrom roughly 1.5 to 30 solar mass (M ⊙ ) that are thought to be capable of beingform<strong>ed</strong> by the evolution of very massive stars as stellar mass black holes, largerones up to about 3 × 10 7 M ⊙ as massive black holes (MBHs), and still larger onesas supermassive black holes.As is well known, supermassive black holes were invok<strong>ed</strong> first to providethe energy source necessary to explain quasars. However, more recent opticaland other observations have provid<strong>ed</strong> strong evidence for the existence of MBHsin AGNs much closer to the Earth [49]. In one case, the evidence is fromobservations of OH masers in Keplarian orbits around the MBH, and is essentiallyconclusive [50, 51].For some normal galaxies, the evidence also is very strong. One case comesfrom extremely high-resolution infrar<strong>ed</strong> observations of stars moving around a2.6 ± 0.2 × 10 6 M ⊙ object at the centre of our galaxy, that really can only be aMBH [52, 53]. Two other cases come from optical observations of the motionsof stars or gas around the centres of the M31 (Androm<strong>ed</strong>a) and M32 galaxies inthe Local Group at distances of about 0.6 Mparsec. The evidence for stellar massblack holes comes from observations of x-ray binaries.For MBH masses of 10 6 M ⊙ or less, the only fairly convincing observationsso far are two OH maser observations indicating about 10 6 M ⊙ central objects.Except for a few more observations of this kind, it is not apparent whetherelectromagnetic observations are likely to tell us much about the fraction ofnormal galaxies containing MBHs that are this low in mass. Because of this,the information potentially available from gravitational wave observations withLISA concerning the existence and masses of MBHs in other normal galaxies andhow they form<strong>ed</strong> is likely to be quite valuable.There are two main types of theories concerning how se<strong>ed</strong> black holes wereform<strong>ed</strong>. In one, collisions of stellar mass objects in dense galactic nuclei l<strong>ed</strong> to theformation of higher mass objects, and these sank down toward the centre (masssegregation), where their collision rates were enhanc<strong>ed</strong>. If the mass got to be afew hundr<strong>ed</strong> times the solar mass and an object was not already a black hole, it


Expect<strong>ed</strong> gravitational-wave results from LISA 137would evolve quickly to form one. When the largest objects got to be roughly athousand solar mass in size, they would then be able to continue growing fairlyrapidly by absorption of gas in the galactic nucleus and by tidal disruption of stars.At some point, the largest black hole would grow enough faster than the othersthat it would swallow up the ones of comparable size and become the se<strong>ed</strong> forgrowth of a perhaps 10 5 M ⊙ or larger MBH.The alternate type of theory involves the evolution of a dense cloud of gasand dust to the point where it becomes optically thick, and radiation pressureplus magnetic fields can prevent further fragmentation of the cloud to form stars.At that point, if energy and angular momentum can be dissipat<strong>ed</strong> fairly rapidly,there are two options. In one, a supermassive star possibly 10 5 –10 6 M ⊙ in sizeis form<strong>ed</strong>, and quickly evolves to the point of relativistic collapse and forms aMBH. In the other option, the cloud can become dense enough to reach the pointof relativistic instability and collapse directly to a MBH without going through thesupermassive star stage. There also is the possibility of a relativistic star clusterbecoming unstable and collapsing to a MBH.For the collisional growth scenario, quite detail<strong>ed</strong> calculations starting from1M ⊙ stars were carri<strong>ed</strong> out by Quinlan and Shapiro [54] (see this 1990 paper forearlier references). They found that roughly 100M ⊙ objects could form in a fewtimes 10 8 years, starting from plausible conditions in a dense galactic nucleus,and including the effects of mass segregation. However, it was not possible at thattime to follow the process further.In an alternate approach, Lee [55] start<strong>ed</strong> from assuming that 1% of the massin a dense galactic nucleus was in the form of 7M ⊙ black holes that result<strong>ed</strong> fromevolution of stars at the high end of the initial mass function (i.e. initial massdistribution). The rest of the material was in the form of 0.7M ⊙ normal stars.Dynamical friction l<strong>ed</strong> to segregation of the black holes to the core, and corecollapse among the black holes occurr<strong>ed</strong> on a time scale much shorter than fora single component cluster. For rms stellar velocities above 100 km s −1 and forplausible densities in the nucleus, it was shown that many black hole binariesform<strong>ed</strong> and merg<strong>ed</strong> to produce 14M ⊙ black holes within about two billion years.The process was not follow<strong>ed</strong> further, but it seems likely that most of the blackholes would have merg<strong>ed</strong> rapidly to form a substantial siz<strong>ed</strong> se<strong>ed</strong> MBH.The main objection rais<strong>ed</strong> to the collisional growth scenario is that it seemsdifficult to produce the se<strong>ed</strong> black holes and have them grow much further tothe supermassive black hole size before the appearance of quasars as early asa r<strong>ed</strong>shift of four [56–58]. Instead, it was suggest<strong>ed</strong> that the inefficiency ofstar formation would leave most of the material in a dense cloud in the formof gas and dust. The cloud would cool and condense toward the centre untilangular momentum support became important. <strong>Gravitational</strong> instabilities andother effects would help to remove energy and angular momentum and permitthe density to become high enough for collapse to a supermassive star or directlyto a MBH perhaps 10 5 M ⊙ or larger in size. A relat<strong>ed</strong> argument made is that aself-gravitating gaseous object of more than 10 8 M ⊙ does not appear to have any


138 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionstationary non-relativistic equilibrium state that can be support<strong>ed</strong> for very long.Under the cloud collapse scenario, an important question is whether signalsare likely to be produc<strong>ed</strong> that LISA could see. For example, if a supermassivestar forms and then evolves to the relativistic instability, the final collapse to aMBH could be slow enough that most of the gravitational-wave radiation wouldbe at such low frequencies that LISA would have poor sensitivity. Also, ifthe collapse were nearly spherically symmetric, the radiative efficiency wouldbe poor. However, recent fully relativistic calculations by Baumgarte andShapiro [59] of the evolution of a rotating supermassive star up to the onset ofcollapse provide some basis for a more optimistic view. The later evolution canbe determin<strong>ed</strong> reliably only by a numerical, three-dimensional hydrodynamicssimulation in general relativity. However, estimates of what will happen indicatethat most of the mass will go into a MBH, and that a bar instability which radiatesefficiently at frequencies observable by LISA may form.Even if the supermassive black holes in quasars at high r<strong>ed</strong>shifts are form<strong>ed</strong>initially by cloud collapse, it still seems quite possible that the collisional growthscenario may contribute substantially to the formation of se<strong>ed</strong> black holes formore modest siz<strong>ed</strong> MBHs, like the one in our galaxy. Under the collisionalgrowth scenario, if a number of 500M ⊙ se<strong>ed</strong> black holes are form<strong>ed</strong> beforerunaway growth occurs and the largest has already swallow<strong>ed</strong> the others, thenthe coalescence of two of these se<strong>ed</strong>s could be seen by LISA even at a substantialr<strong>ed</strong>shift. For such a coalescence at z = 5, it can be shown that the signal strengthas a function of frequency during the last year before coalescence, for a circularorbit, would stay just about at the LISA threshold sensitivity curve level, so theevent would be detectable with S/N = 5.A remaining question, even if many interm<strong>ed</strong>iate siz<strong>ed</strong> MBHs are produc<strong>ed</strong>by the collisional growth of se<strong>ed</strong>s, is whether the time of runaway growth ofthe largest se<strong>ed</strong> black hole would be delay<strong>ed</strong> to high enough mass for LISA toobserve the coalescences. As point<strong>ed</strong> out by Lee [60], the calculations of Quinlanand Shapiro are bas<strong>ed</strong> on the Fokker–Planck approach, and that approach doesnot allow for the proper statistical treatment of a runaway instability. This is truefor the calculations of Lee [55] also. In considering this issue, it should be not<strong>ed</strong>that only the chirp mass for the binary is important for determining the signalstrength and frequency as a function of time. Thus the coalescence of a 100M ⊙black hole with a 4000M ⊙ one would be as observable as for two 500M ⊙ blackholes. Further work on the runaway growth question, as well as on the overallcollisional growth and cloud collapse scenarios, certainly would be valuable.10.2.3 Massive black holes in normal galaxiesAnother important astrophysical question concerns the abundance of interm<strong>ed</strong>iatesize MBHs of roughly 10 5 –10 6 M ⊙ . From observations bas<strong>ed</strong> almost entirely ongalaxies containing larger MBHs and SMBHs, various authors have estimat<strong>ed</strong>that the mass of the central object is about 500 times smaller than the mass of the


Expect<strong>ed</strong> gravitational-wave results from LISA 139spheroid (central bulge) of the galaxy. An even tighter relationship to the velocitydispersion in the spheroid has been report<strong>ed</strong> recently [61]. However, the reasonsfor these relationships are not yet known [56, 62–64]. It appears likely that LISAdata can address whether a relation something like this extends to galaxies withsmaller spheroids, which constitute the majority of all galaxies.MBHs in galactic centres are expect<strong>ed</strong> to usually have an increas<strong>ed</strong> densityof stars around them in the region where the potential of the MBH dominates thatof the galaxy. This density cusp is usually taken to be a power law cusp, with a−7/4 power dependence on radius if the distribution of stellar motions is relax<strong>ed</strong>and a −3/2 power for some unrelax<strong>ed</strong> cusps. It is generally expect<strong>ed</strong> that therewill be large numbers of compact stars, i.e. white dwarfs and neutron stars, in thecusp. Occasionally, one of them that is on a nearly radial orbit and passes closeto the MBH may be deflect<strong>ed</strong> enough by the other stars so that it comes withinfive or so gravitational radii of the MBH and loses significant amounts of energyand angular momentum by gravitational radiation. If so, and if further deflectionsby the other stars are not important, the compact star orbit will continue to shrinkgradually until coalescence with the MBH occurs.Unfortunately, in almost all cases for white dwarfs and neutron stars, theabove gradual approach scenario is interferr<strong>ed</strong> with by interactions with otherstars. Hils and Bender [65] have simulat<strong>ed</strong> what happens for a particular modelwhich assumes 1M ⊙ for both the compact stars and the normal stars in the cusp.After the first pass near the MBH, the orbit of the compact star is modifi<strong>ed</strong> byinteractions with the other stars more rapidly than by the gravitational radiation,unless the compact star is bound very tightly to the MBH initially. Thus thecompact star usually will plunge rapidly into the MBH, or its point of closestapproach will wander far enough away to not give appreciable interaction. A rapidplunge will not provide enough integration time for detection. In the remainingfavourable cases, the signal typically will be observable by LISA from a one yeardata record starting up to roughly 100 years before coalescence.Despite a loss of several orders of magnitude in the event rate due toplunging, the study by Hils and Bender [65] gave some hope of LISA seeing suchsignals. However, studies by Sigurdsson and Rees [66] and by Sigurdsson [67],plus an unpublish<strong>ed</strong> extension of the above study by Hils and Bender, indicat<strong>ed</strong>that the prospects were considerably better for observing gradual approaches tocoalescence with galactic centre MBHs for roughly 5 or 10M ⊙ black holes. Theeffect of mass segregation was includ<strong>ed</strong>. Such events would be detectable at ar<strong>ed</strong>shift of z = 1 during the last year before coalescence for MBH masses from5 × 10 4 to 2 × 10 6 M ⊙ .Estimates of the event rate are certainly model dependent, but still offerencouragment that multiple signals of this kind will be observ<strong>ed</strong>. As an example,results obtain<strong>ed</strong> by Hils and Bender for one particular model are shown infigure 10.10. About 1% of the mass in the cusp was assum<strong>ed</strong> to be in 7M ⊙black holes, and mass segregation was includ<strong>ed</strong>. For each factor two range inthe mass of the central MBH about a nominal value M, and for each factor two


140 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionFigure 10.10. Expect<strong>ed</strong> signals from BH–MBH binaries. The instrumental and confusionnoise thresholds are shown for one year of observations, but for a S/N ratio of 10. Th<strong>ed</strong>ifferent symbols correspond to different MBH masses. For a given mass, the individualpoints correspond to factor of two different values of the r<strong>ed</strong>shift z.range in the r<strong>ed</strong>shift z, the signal strength and frequency of the strongest expect<strong>ed</strong>signal was calculat<strong>ed</strong>. The different symbols correspond to the different MBHmasses, ranging from 5 × 10 5 to 4 × 10 6 M ⊙ , and the highest and lowest valuesof z for each M are labell<strong>ed</strong>. It should be not<strong>ed</strong> that, only in this figure, thethreshold S/N ratio is taken to be ten rather than five. This is because the orbitsin this case generally will be quite complex because of relativistic effects, asdiscuss<strong>ed</strong> later, and thus require a higher S/N ratio for detection. With improv<strong>ed</strong>understanding of the conditions in the cusps, the distribution of MBH massesfor such events may give information that cannot be obtain<strong>ed</strong> in other ways onthe demographics of interm<strong>ed</strong>iate mass black holes in galactic centres throughoutmuch of the universe.A recent paper by Miralda-Escuda and Gould [68] looks specifically at thestellar mass black holes in the cusp around the MBH in the centre of our galaxy.With assum<strong>ed</strong> masses of 0.7M ⊙ for stars in the cusp and 7M ⊙ for the black holes,and with 1% of the mass in the black holes, they show that the black holes willrapidly sink down closer to the center. With this model, they estimate that about24 000 black holes would be concentrat<strong>ed</strong> within 0.7 parsec of the MBH, andcalculate a lifetime for loss of the black holes to the MBH of about 3×10 10 years.If a substantial fraction of the black holes approach coalescence gradually, andwith many galaxies like ours in the universe, the possible LISA event rate fromthis study also may be considerable. On the other hand, the authors suggest thatmost of the stars in the inner few parsecs of the cusp would be forc<strong>ed</strong> out to larger


Expect<strong>ed</strong> gravitational-wave results from LISA 141distances by interaction with the black holes. If so, the estimat<strong>ed</strong> observable eventrate for white dwarfs and neutron stars from [65] could be drastically r<strong>ed</strong>uc<strong>ed</strong>.10.2.4 Structure formation and massive black hole coalescenceThe third of the important astrophysical questions that LISA has a good chance ofgiving new information on concerns the development of structure in the universe.It is currently believ<strong>ed</strong> that objects considerably smaller than galaxies form<strong>ed</strong>first, and that continuing sequences of interactions and mergers l<strong>ed</strong> to the presentdistribution of galaxy types and sizes, and to galaxy clusters and superclusters. IfMBHs were already present in pre-galactic structures that merg<strong>ed</strong>, this could leadto the MBHs sinking down to the centre of the new structure by dynamical frictionand getting close enough together to coalesce due to gravitational radiation beforethe next merger [58, 69].The above scenario is quite attractive, and could give a high event rate forLISA. If so, valuable new constraints on the merger process and on the conditionsafter mergers would be obtain<strong>ed</strong>. However, there are a number of issues that affectthe event rate that have to be consider<strong>ed</strong>.One question is how effective dynamical friction will be in bringing MBHsof a given size to the centre of the new structure in less than the time betweenmergers. For galaxies like ours, the time requir<strong>ed</strong> is less than the Hubble time forMBHs of roughly 3×10 6 M ⊙ or larger (see, e.g., Zhu and Ostriker [70]). However,the stars in the cusp around a MBH before merger will stay with it for some time,and will affect the dynamics. Thus, whether the mergers of pre-galactic structuresin galaxies like ours with fairly modest spheroid and MBH masses are likely tohave produc<strong>ed</strong> coalescences of perhaps 10 5 or 10 6 M ⊙ MBHs appears to be anopen one.There also is a question about whether the MBHs will modify the stardistribution near the centre of the new structure enough so that the dynamicalfriction will be decreas<strong>ed</strong> considerably. Calculations indicate that this may occurwhen the MBHs are fairly close together, but not yet close enough to coalesce bygravitational radiation in less than a Hubble time (see, e.g., Makino [71]. On theother hand, conditions such as Brownian motion of the MBH and tri-axiality in thestructure may affect the results, and there also may be complex enough motionsgoing on soon after a merger to change the conclusions. Thus the effectiveness ofthis ‘hang-up’ in the coalescence process is not known.Another question arises if another merger occurs before the MBHs from anearlier merger have coalesc<strong>ed</strong>. In principle, one, two or all three of the MBHscould be thrown out of the new structure by the slingshot effect. However, thenew merger could even have a beneficial effect, if the resulting disturbances inthe central star density overcame the possible hang-up for the earlier two MBHsand allow<strong>ed</strong> them to coalesce.The signals arising from MBH coalescences after mergers of galaxies or ofpre-galactic structures would be very large. Figure 10.11 shows several cases


142 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionFigure 10.11. Strain amplitudes during the last year before MBH–MBH coalescence. Thecurves are for MBH–MBH binaries at z = 1 with circular orbits, and for one year ofobservations with S/N = 5. For a given mass combination, the first five points from the leftare for 1.0, 0.8, 0.6, 0.4, and 0.2 months before coalescence. The last two points are for0.5 week before and for very roughly the last stable orbit for Schwarzschild MBHs.of the signal strength as a function of frequency during the last year beforecoalescence for events at a r<strong>ed</strong>shift of z = 1 and for circular orbits. The squaresymbols show the time at 0.2 year intervals, so the first is one year before, thesecond is 0.8 year before, etc. The last symbol is shift<strong>ed</strong> slightly to be 0.5 weekbefore coalescence, instead of at that time. The cases shown are for equal MBHmasses, and only the spiral-in part of the event before the last stable orbit isreach<strong>ed</strong> is consider<strong>ed</strong>. However, it should be remember<strong>ed</strong> that the detectability ofthe signal for the spiral-in phase depends mainly on the chirp mass, so the curvesgiven can be us<strong>ed</strong> to estimate conditions under which unequal mass coalescenceswould be observable also. The case of possible coalescence of 500M ⊙ se<strong>ed</strong> blackholes during the early growth of MBHs that was discuss<strong>ed</strong> earlier is shown forz = 1 for comparison.Since the LISA threshold sensitivity curve and the confusion noise estimateare defin<strong>ed</strong> on the assumption that the frequency and signal strength are fix<strong>ed</strong> forthe one year period of the observation, and the MBH–MBH coalescence signalschange dramatically during the year, it is necessary to integrate the square of theamplitude S/N ratio during the year and then take the square root to obtain theeffective S/N ratio. The way in which the effective S/N ratio builds up duringthe year, with the LISA instrumental noise and the confusion noise combin<strong>ed</strong>quadratically, is shown in figure 10.12. What is shown for each of the cases from


Expect<strong>ed</strong> gravitational-wave results from LISA 143Figure 10.12. Cumulative weekly S/N ratios during the last year before MBH–MBHcoalescence at z = 1. Both the instrumental noise and the estimat<strong>ed</strong> binary confusionnoise are includ<strong>ed</strong>.figure 10.11 is the accumulat<strong>ed</strong> S/N ratio after the first week, the second week,etc, up until the end of the year. The large jump in the last week or so, in all butthe 500M ⊙ case, is due to the frequency of the signal having swept up to wherethe LISA sensitivity is much higher.It is clear by extrapolating roughly from figures 10.11 and 10.12 that MBH–MBH coalescences resulting from structure mergers could be seen clearly even atr<strong>ed</strong>shifts of 10 for 10 7 M ⊙ MBHs and 20 or more for less massive ones. Thus,any such events at any plausible occurrence time will be observable. Particularlybecause of the large uncertainties concerning what happens for roughly 10 6 M ⊙or smaller MBHs after mergers of structures, the event rate for LISA apparentlycould range from less than one per decade to quite a large value.10.2.5 Fundamental physics tests with LISAOf similar importance to the astrophysical questions about MBHs discuss<strong>ed</strong> inthe last three sections is the fundamental physics question of whether Einstein’sgeneral relativity theory is correct. Very strong tests of the theory can be provid<strong>ed</strong>either by MBH–MBH coalescences with high S/N ratio, if they are observ<strong>ed</strong>,or by highly unequal mass coalescences of 5 or 10M ⊙ black holes with MBHsin galactic centres. Only the latter case will be discuss<strong>ed</strong> here, since the muchlarger number of orbital periods observable and the substantial orbit eccentricityexpect<strong>ed</strong> are likely to be more important for testing relativity than a higher S/Nratio.


144 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionIt has been emphasiz<strong>ed</strong> by Cutler et al [72] that the strongest test of a theoryis likely to come from the observ<strong>ed</strong> phase of the signal, rather than the amplitude.This is because the cross-correlation of a theoretical template with the observ<strong>ed</strong>signal will be r<strong>ed</strong>uc<strong>ed</strong> substantially if the two get out of phase by even half a cycleor less during the entire data record. For our case of a perhaps 10M ⊙ black holeorbiting around a 10 5 or 10 6 M ⊙ MBH during the last year before coalescence,there will be roughly 10 5 cycles. Even a very small error in the metric will lead toa continuously increasing error in the orbit, and thus in the phase of the calculat<strong>ed</strong>signal. While some type of metric error could, in principle, affect the orbit inthe same way as slightly different values of parameters in the problem, such asthe spin and mass of the MBH, it seems unlikely that a conceptual breakdown inthe theory would be nearly equivalent to just having different parameter values.Thus, correctly pr<strong>ed</strong>icting the signal phase over 10 5 cycles would be an extremelystrong test of the theory.For this case, it is important that the orbit start<strong>ed</strong> out being nearly radial.The later evolution of the eccentricity from its initial value very close to unitycan be calculat<strong>ed</strong> from the rates of loss of energy and angular momentum due togravitational radiation. The results obtain<strong>ed</strong> by Tanaka et al for a SchwarzschildMBH [73] show for a typical case that the orbit never becomes circular, butinstead remains substantially eccentric (0.5 ?) up until the final plunge begins.However, rigorous calculations will be requir<strong>ed</strong> in order to fit the data well [74].During the last year, the periapsis distance changes only very slowly, but theapoapsis distance decreases substantially. At periapsis, the spe<strong>ed</strong> is about halfthat of light, so the dynamics are highly non-Newtonian. In fact, the precessionof periapsis during one radial motion period can be about a whole cycle. Withrelativistic beaming of the gravitational radiation and variation of the strengthof the radiation with time during a radial period, the amplitude observ<strong>ed</strong> in agiven direction can vary in a quite complex way. For an orbit plane that is notperpendicular to the spin axis for a rapidly rotating MBH, rapid Lense–Thirringprecession also will be present. In view of the complexity of the relativisticmotion and the large number of cycles over which the phase of the signal canbe follow<strong>ed</strong>, such a signal would give a nearly ideal test of the pr<strong>ed</strong>ictions ofgeneral relativity.It is of course necessary to be able to detect the signal in order to test thetheory. It was mention<strong>ed</strong> earlier that a S/N ratio of about ten would be ne<strong>ed</strong><strong>ed</strong>in order to detect such a complex signal. But there also is the question of howdifficult the search for such a signal would be. A very crude estimate of thenumber of templates ne<strong>ed</strong><strong>ed</strong> for a brute force search with one year of data gives,within a couple of orders of magnitude, something like 10 18 . Even with rapidadvances in computing power, such a search probably would not be possible.Some improvement could be made by a hierarchical search strategy, but it is notclear that this approach would be sufficient. On the other hand, more powerfulsearch algorithms such as ‘genetic algorithms’ and ‘stimulat<strong>ed</strong> annealing’ havebeen develop<strong>ed</strong> and demonstrat<strong>ed</strong> for substantially more challenging search


Expect<strong>ed</strong> gravitational-wave results from LISA 145problems. Consideration of such algorithms for use with LISA data is juststarting, but it currently appears unlikely that the difficulty of the search problemwill be a real limitation in the use of the LISA data, even for the case of highlyunequal mass binaries. For all other expect<strong>ed</strong> types of LISA sources, the searchproblem is much easier than for ground-bas<strong>ed</strong> detectors because of the roughly10 4 times lower number of data points that have to be handl<strong>ed</strong> for a year ofobservations.The other important question is whether the ability to calculate theoreticaltemplates will have improv<strong>ed</strong> enough by the time LISA data is available to carryout a thorough test of general relativity. For comparable mass MBHs, there is avery long way to go to accomplish this. However, for the highly unequal masscase, the chances for fairly rapid progress seem much better. It is hop<strong>ed</strong> thatnumerical methods can be develop<strong>ed</strong> that start from the test mass approximation,and converge moderately well. Still, since small changes in the initial conditionscan lead to very large changes in the motion a year later, it is important thatthis problem as well as the search problem for the highly unequal mass case bepursu<strong>ed</strong> vigorously in the next few years.For either detail<strong>ed</strong> tests of general relativity with LISA or for studies ofastrophysical questions concerning MBHs, the first requirement is that somesignals involving MBHs be seen. While it seems likely that several of the types ofMBH sources discuss<strong>ed</strong> earlier will be observ<strong>ed</strong>, this certainly is not guarante<strong>ed</strong>.Thus, the situation is somewhat like that for ground-bas<strong>ed</strong> observations, wherethe detection of signals within the next decade seems quite likely, but is notcertain. Still, a reasonable pr<strong>ed</strong>iction is that both ground-bas<strong>ed</strong> detectors in thenext decade and LISA not too much later will detect the desir<strong>ed</strong> types of signals,and extremely strong tests of general relativity will be among the most importantscientific results.Another fundamental physics test that LISA may contribute to concernsthe possible existence of a primordial gravitational-wave background [75, 76].Standard inflation theory pr<strong>ed</strong>icts a nearly scale-invariant spectrum, in which thespectral amplitude for the gravitational <strong>waves</strong> falls off at about the −1.5 powerof the frequency. However, the observ<strong>ed</strong> COBE microwave background spectrumis believ<strong>ed</strong> to be determin<strong>ed</strong> mainly by the large-scale density fluctuations at thetime of decoupling, and with a scale invariant spectrum would pr<strong>ed</strong>ict a very lowamplitude in the frequency range of LISA and of ground-bas<strong>ed</strong> detectors.There are theories that could give a non-scale-invariant spectrum anddetectable amplitudes for a cosmic background at LISA or ground-bas<strong>ed</strong>frequencies. LISA could detect an isotropic cosmic background near 10 mHz if itsenergy density were roughly 10 −11 of the closure density. A candidate for giving adetectable background is a phase transition at about the electro-weak energy scale,but it would have to be strongly first order in order to give enough amplitude. Butthis is not currently thought to be likely. Another possibility is associat<strong>ed</strong> withsuggest<strong>ed</strong> effects of extra dimensions, which may have become very small at atime which would produce a peak in the gravitational-wave spectrum near the


146 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionLISA frequency band. In any case, LISA will improve the limits on a possiblecosmic background in its frequency band, but the chances of seeing anythingappear to be very uncertain.10.2.6 Future prospectsProvid<strong>ed</strong> that signals involving massive black holes are inde<strong>ed</strong> seen by LISA,it is likely that there will be an opportunity for a later advanc<strong>ed</strong> gravitationalwavemission. Depending on what is seen, there are several directions in whichmajor instrumental improvements could be made. If improvements in sensitivityat frequencies above 3 mHz are of most interest, this can be achiev<strong>ed</strong> by increasingthe telescope size and the laser power, provid<strong>ed</strong> that comparable r<strong>ed</strong>uctions inother noise sources such as beam pointing jitter and phase measurement errorsalso can be made. Shortening the antenna arm length also would be desirable. Ifimprovements below about 0.1 mHz are the main objective, then improv<strong>ed</strong> freemass sensors and longer arm lengths would be ne<strong>ed</strong><strong>ed</strong>.As an example of a mission to achieve strongly increas<strong>ed</strong> high-frequencysensitivity, a goal of using 10 W lasers and 1 m diameter telescopes might bechosen, along with a r<strong>ed</strong>uction in the antenna arm length to 50 000 km. Theresulting factor of about 10 7 increase in the receiv<strong>ed</strong> power would r<strong>ed</strong>uce the shotnoise contribution to measuring changes in the arm length difference from about1 × 10 −11 m/rtHz to 3 × 10 −15 m/rtHz, but the signal strength for frequenciesbelow about 1 Hz would be r<strong>ed</strong>uc<strong>ed</strong> also because of the r<strong>ed</strong>uc<strong>ed</strong> arm lengths. Ifother errors such as from beam pointing jitter and from the phase measurementsonly double the distance measurement error, the level of the LISA thresholdsensitivity curve can be r<strong>ed</strong>uc<strong>ed</strong> by a factor ranging from about 20 at 10 mHzto 3000 at 1 Hz, and then retaining that value at higher frequencies. However,with the same assumptions about the confusion noise level due to extragalacticCWDBs as us<strong>ed</strong> for figures 10.8–10.11, and plausible estimates for extragalacticneutron star and black hole binaries, the overall sensitivity will be limit<strong>ed</strong> moreby the confusion noise than the instrumental noise up to at least 100 mHz.One strategy that is not plann<strong>ed</strong> for LISA but probably would be us<strong>ed</strong> in alater high-frequency mission with shorter arm lengths is to apply enough forceto each test mass to keep the rates of change of the arm lengths constant. Thisis because the force requir<strong>ed</strong> is about a factor 6000 less for 50 000 km arms, orroughly 4 × 10 −10 ms −2 . Also, the main emphasis is on the noise level above3 mHz, where keeping fluctuations in the appli<strong>ed</strong> voltages low is easier than atlower frequencies. This would make the phase measurements very much simplerthan they would be with substantial Doppler shifts.Quite a few types of measurements could be improv<strong>ed</strong> considerably withthe above sensitivity. For example, gradual coalescences of 10M ⊙ black holeswith MBHs with masses down to 10 3 M ⊙ could be observ<strong>ed</strong> virtually anywherein the universe, if they occur. And tests of general relativity for coalescenceswith highly unequal masses would have considerably higher S/N ratios, as well


Expect<strong>ed</strong> gravitational-wave results from LISA 147as large numbers of cycles. In addition, sensitivity to events involv<strong>ed</strong> in the initialformation of interm<strong>ed</strong>iate mass black holes would be improv<strong>ed</strong>.For improv<strong>ed</strong> measurements at low frequencies, below about 0.1 mHz, themost important goals are r<strong>ed</strong>uc<strong>ed</strong> noise levels from the free mass sensors andlonger arm lengths. Moderate goals might be a factor ten r<strong>ed</strong>uction in the freemass sensor noise level plus a factor three increase in the arm length. Basically thesame geometry could be us<strong>ed</strong> as for LISA, but the antenna probably would have tobe considerably farther from the Earth to keep the arm length changes from beingtoo large. This antenna would make it possible to observe MBH–MBH binarieswith both masses above about 10 5 M ⊙ much longer before coalescence, and thusincrease the number of such events observ<strong>ed</strong> by a large factor.A much more challenging goal would be to try for roughly 1 AU arm lengthsand a factor 100 r<strong>ed</strong>uction in the free mass sensor noise. This would permitimprov<strong>ed</strong> measurements mainly at frequencies below about 0.02 mHz, assumingthat the confusion noise level is nearly constant below this frequency, as given inHils, Bender and Webbink [39]. This would further push back the time beforecoalescence when MBH–MBH binaries could be observ<strong>ed</strong>, and probably tell usconsiderably more about such coalescences after mergers during the process ofgalaxy formation.One possible geometry would be to locate spacecraft near the L4 and L5points of the Earth–Sun system, 60 ◦ in front of and behind the Earth, and neareither the L1 orL2 point, within about 1500 000 km of the Earth. In this casethe third arm would be about 3 0.5 times the length of the other two. Because theL1 and L2 points are unstable, putting the third spacecraft near the L3 point onthe opposite side of the Sun from the Earth and various other possibilities wouldbe consider<strong>ed</strong>. One disadvantage of such geometries is that the antenna is in theecliptic plane, and the sensitivity to the ecliptic latitude of sources at low latitudesis r<strong>ed</strong>uc<strong>ed</strong>. However, the extra impulse requir<strong>ed</strong> to go considerably out of theecliptic is large.It is interesting to note that the requirement on the accuracy of distancemeasurements between the test masses for frequencies below 0.1 mHz are muchless severe than those for LISA, even for roughly 1 AU arm lengths. The distancemeasurement error would have to be about 1×10 −8 m/rtHz to equal the limitationdue to the confusion noise at 0.1 mHz, and more than a factor (0.1 mHz/ f ) 2larger at a lower frequency f . Thus the requir<strong>ed</strong> combination of laser power andtelescope size will be affect<strong>ed</strong> more by considerations such as staying sufficientlyabove possible noise in the scatter<strong>ed</strong> light level rather than by the desire to keepthe shot noise level low. In any case, 10 W of laser power and 1 m diametertelescopes would at least be adequate, since it would give the same receiv<strong>ed</strong> poweras for LISA.Achieving a factor 100 lower free mass sensor noise than for LISA wouldbe a major challenge. It might be desirable, for example, to consider havingeach free mass sensor in a small separate slave spacecraft containing as littleother equipment as possible in order to minimize disturbances. Several such slave


148 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionspacecraft could be includ<strong>ed</strong> if they could be made much smaller than the mainspacecraft. With the increas<strong>ed</strong> tolerance on the distance measurement uncertaintybelow 0.1 mHz, a small local diode laser system could track changes in theinternal geometry of each combination of a spacecraft plus its slaves. However,this approach would help only if the main noise sources in each free mass sensorwere independent. This would not be the case if common fluctuations in the solarintensity contribut<strong>ed</strong> substantially to the noise.In view of the apparent advantages of improving the sensitivity at both highand low frequencies after the LISA mission, it may be desirable to consider flyingtwo separate antennae, with one at least somewhat like the high frequency antennadiscuss<strong>ed</strong> earlier and the other like the moderately improv<strong>ed</strong> low frequencyantenna. However, questions such as this can be address<strong>ed</strong> much more easilyafter some data from LISA have been receiv<strong>ed</strong>.Acknowl<strong>ed</strong>gmentsIt is a pleasure to thank all of the people who have been involv<strong>ed</strong> in the variousstudies of LISA, including, in particular, Karsten Danzmann, Jim Hough, BernardSchutz, Stefano Vitale and the rest of our European colleagues. In the USA, RaiWeiss and Ron Drever and later Mark Vincent were very helpful in getting th<strong>ed</strong>iscussions start<strong>ed</strong>; Kip Thorne and Jan Hall have provid<strong>ed</strong> constant informationand encouragement throughout; and Bill Folkner, Sterl Phinney, Doug Richstone,Ron Webbink, John Armstrong, Ron Hellings and other members of the LISAMission Definition Team have been essential to the progress that has beenachiev<strong>ed</strong>. The 1999–2000 ESA Industrial Study has add<strong>ed</strong> a number of otherpeople who have made important contributions to future progress. At JILA,essential ideas and long-term creative efforts from Tuck Stebbins, Dieter Hilsand Jim Faller have been responsible for much of our present understanding ofthe opportunity present<strong>ed</strong> by the LISA mission.References[1] Press W H and Thorne K S 1972 Ann. Rev. Astron. Astrophys. 10 335–71[2] Moss G E, Miller L R and Forward R L 1971 Appl. Opt. 10 2495[3] Faller J E and Bender P L 1984 Precision Measurements and Fundamental ConstantsII (NBS Spec. Pub. 617) (Washington, DC: US Govt Printing Office) pp 689–90[4] Faller J E, Bender P L, Hall J L, Hils D and Vincent M A 1985 Proc. Colloq.on Kilometric Optical Arrays in Space, Cargese (Corsica), ESA Report SP-226pp 157–63[5] Laser Interferometer Space Antenna 1998 Proc. 2nd Int. LISA Symp. (AIP Conf. Proc.456) <strong>ed</strong> W M Folkner (Woodbury, NY: American Institute of Physics) pp 3–239[6] LISA 1998 Pre-Phase A Report MPQ-233, 2nd <strong>ed</strong>n, <strong>ed</strong> K V Danzmann (Garching:Max-Planck Institute for Quantum Optics) pp 1–191[7] 1997 Class. Quantum Grav. 14 1397–585


References 149[8] Folkner W M 1998 Laser Interferometer Space Antenna Proc. 2nd Int. LISA Symp.(AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner (Woodbury, NY: American Institute ofPhysics) pp 11–16[9] Stebbins R T 1998 Laser Interferometer Space Antenna Proc. 2nd Int. LISA Symp.(AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner (Woodbury, NY: American Institute ofPhysics) pp 17–23[10] Whalley M S, Turner R F and Sandford M C W 1998 Laser Interferometer SpaceAntenna Proc. 2nd Int. LISA Symp. (AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner(Woodbury, NY: American Institute of Physics) pp 24–30[11] Peskett S, Kent B, Whalley M and Sandford M 1998 Laser Interferometer SpaceAntenna Proc. 2nd Int. LISA Symp. (AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner(Woodbury, NY: American Institute of Physics) pp 31–8[12] Giampieri G, Hellings R W, Tinto M and Faller J E 1996 Opt. Commun. 123 669–78[13] Armstrong J W, Estabrook F B and Tinto M 1999 Ap. J. 527 814–26[14] Estabrook F B, Tinto M and Armstrong J W 1999 Ap. J. 527 814[15] Cutler C 1998 Phys. Rev. D 70 89–102[16] Staff of Space Science Department, Johns Hopkins University Appli<strong>ed</strong> PhysicsLaboratory and Staff of Guidance and Control Laboratory, Stanford University1974 AIAA J. Spacecraft 11 637–44[17] Lange B 1964 AIAA J. 2 1590–606[18] Touboul P, Rodrigues M and Le Clerc G M 1996 Class. Quantum Grav. 13 A259–70See also section 3.3 of LISA 1996 Pre-Phase A Report MPQ-208, 1st <strong>ed</strong>n, <strong>ed</strong>K V Danzmann (Garching: Max-Planck Institute for Quantum Optics) pp 60–75[19] Speake C C 1996 Class. Quantum Grav. 13 A291[20] Vitale S and Speake C C 1998 Laser Interferometer Space Antenna Proc. 2nd Int.LISA Symp. (AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner (Woodbury, NY: AmericanInstitute of Physics) pp 172–7[21] Josselin V, Rodrigues M and Touboul P 1998 Laser Interferometer Space AntennaProc. 2nd Int. LISA Symp. (AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner (Woodbury,NY: American Institute of Physics) pp 236–9[22] Wollard M 1984 Short-period oscillations in the total solar irradiance ThesisUniversity of California, San Diego, CA[23] Fehringer M, Ru<strong>ed</strong>enauer F and Steiger W 1998 Laser Interferometer Space AntennaProc. 2nd Int. LISA Symp. (AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner (Woodbury, NY:American Institute of Physics) pp 207–13[24] Thorne K 1987 300 Years of Gravitation <strong>ed</strong> S W Hawking and W Israel (Cambridge:Cambridge University Press) pp 330–458[25] Schilling R 1997 Class. Quantum Grav. 14 1513–19[26] Giampieri G 1997 Mon. Not. R. Astron. Soc. 289 185–95[27] Mironowskii V N 1966 Sov. Astron. 9 752–5[28] Iben I 1984 Paper Present<strong>ed</strong> at Workshop on Grav. Waves and Axions (Santa Barbara,CA: Institute for Theoretical Physics)[29] Marsh T R 1995 Mon. Not. R. Astron. Soc. 275 L1–5[30] Marsh T R, Dhillon V S and Duck S R 1995 Mon. Not. R. Astron. Soc. 275 828–40[31] Moran C, Marsh T R and Bragaglia A 1997 Mon. Not. R. Astron. Soc. 288 538–44[32] Maxt<strong>ed</strong> P F L and Marsh T R 1999 Mon. Not. R. Astron. Soc. 307 122–30[33] Webbink R F and Han Z 1998 Laser Interferometer Space Antenna Proc. 2nd Int.LISA Symp. (AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner (Woodbury, NY: American


150 LISA: A propos<strong>ed</strong> joint ESA–NASA gravitational-wave missionInstitute of Physics) pp 61–7[34] Nelemans G, Verbunt F, Yungelson L R and Portegies Zwart S F 2000 Astron.Astrophys. 360 1011–18 (astro-ph/0006216)[35] Hils D, Bender P L, Faller J E and Webbink R F 1986 11th Int. Conf. On GeneralRelativity and Gravitation: Abstracts of Cont. Papers vol 2 (University ofStockholm) p 509[36] Lipunov V M, Postnov K A and Prokhorov M E 1987 Astron. Astrophys. 176 L1–4[37] Lipunov V M and Postnov K A 1987 Sov. Astron. 31 228–30[38] Evans C R, Iben I and Smarr L 1987 Ap. J. 323 129–39[39] Hils D, Bender P L and Webbink R F 1990 Ap. J. 360 75–94[40] Warner B 1995 Astrophys. Space Sci. 225 249–70[41] Verbunt F 1997 Class. Quantum Grav. 14 1417–23[42] Verbunt F 1998 Proc. 2nd Amaldi Conf. on <strong>Gravitational</strong> Waves <strong>ed</strong> E Coccia,G Pizella and G Veneziano (Singapore: World Scientific) pp 50–61[43] Nelemans G, Portegies Zwart S F and Verbunt F 2000 <strong>Gravitational</strong> <strong>waves</strong> andexperimental relativity Proc. 36th Rencontres de Moriond (Les Arc, Jan. 1999)(Hanoi: World Publishers) pp 119–24[44] Bender P L and Hils D 1997 Class. Quantum Grav. 14 1439–44[45] Hils D 1998 Laser Interferometer Space Antenna Proc. 2nd Int. LISA Symp. (AIPConf. Proc. 456) <strong>ed</strong>WMFolkner (Woodbury, NY: American Institute of Physics)pp 68–78[46] Hils D and Bender P L 2000 Ap. J. 537 334–41[47] Kosenko D I and Postnov K A 1998 Astron. Astrophys. 336 786–90[48] Stebbins R T, Bender P L and Folkner W M 1997 Class. Quantum Grav. 14 1499–505[49] Richstone D et al 1998 Nature 395 A14–19[50] Miyoshi M et al 1995 Nature 373 127–9[51] Maoz E 1998 Ap. J. Lett. 494 L181–4[52] Ghez A M, Klein B L, Morris M and Becklin E E 1998 Ap. J. 509 678–86[53] Genzel R, Pichon C, Eckart A, Gerhard O E and Ott T 2000 Mon. Not. R. Astron.Soc. 317 348–74 (astro-ph/0001428)[54] Quinlan G D and Shapiro S L 1990 Ap. J. 356 483–500[55] Lee H M 1995 Mon. Not. R. Astron. Soc. 272 605–17[56] Haehnelt M G and Rees M J 1993 Mon. Not. R. Astron. Soc. 263 168–78[57] Rees M J 1997 Rev. Mod. Astron. 10 179–89[58] Haehnelt M G 1998 Laser Interferometer Space Antenna Proc. 2nd Int. LISA Symp.(AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner (Woodbury, NY: American Institute ofPhysics) pp 45–9[59] Baumgarte T and Shapiro S L 1999 Ap. J. 526 941–52[60] Lee M H 1993 Ap. J. 418 147–62[61] Gebhardt K et al 2000 Ap. J. in press (astro-ph/0006289)[62] Haiman Z and Loeb A 1998 Ap. J. 503 505–17[63] Silk J and Rees M J 1998 Astron. Astrophys. 331 L1–4[64] Kauffmann G and Haehnelt M 2000 Mon. Not. R. Astron. Soc. 311 576–88[65] Hils D and Bender P L 1995 Ap. J. Lett. 445 L7–10[66] Sigurdsson S and Rees M J 1997 Mon. Not. R. Astron Soc. 284 318–26[67] Sigurdsson S 1998 Laser Interferometer Space Antenna Proc. 2nd Int. LISA Symp.(AIP Conf. Proc. 456) <strong>ed</strong> W M Folkner (Woodbury, NY: American Institute ofPhysics) pp 53–6


References 151[68] Miralda-Escuda J and Gould A 2000 Preprint astro-ph/0003269[69] Haehnelt M G 1994 Mon. Not. R. Astron. Soc. 269 199–208[70] Xu G and Ostriker J P 1994 Ap. J. 437 184–93[71] Makino J 1997 Ap. J. 478 58–65[72] Cutler C et al 1993 Phys. Rev. Lett. 70 2984–7[73] Tanaka T, Shibata M, Sasaki M, Tagoshi H and Nakamura T 1993 Prog. Theor. Phys.90 65–83[74] Shibata M 1994 Phys. Rev. D 50 6297–311[75] Allen B 1996 Les Houches School on Astrophysical Sources of <strong>Gravitational</strong> Waves<strong>ed</strong> J-A Marck and J-P Lasota (Cambridge: Cambridge University Press)[76] Hogan C J 1998 Laser Interferometer Space Antenna Proc. 2nd Int. LISA Symp. (AIPConf. Proc. 456) <strong>ed</strong>WMFolkner (Woodbury, NY: American Institute of Physics)pp 79–86


Chapter 11Detection of scalar gravitational <strong>waves</strong>Francesco FucitoINFN, sez. di Roma 2, Via della Ricerca Scientifica, 00133 Rome,ItalyE-mail: Fucito@roma2.infn.itIn this talk I review recent progress in the detection of scalar gravitational <strong>waves</strong>.Furthermore, in the framework of the Jordan–Brans–Dicke theory, I compute thesignal-to-noise ratio for a resonant mass detector of spherical shape and for binarysources and collapsing stars. Finally, I compare these results with those obtain<strong>ed</strong>from laser interferometers and from Einsteinian gravity.11.1 IntroductionThe efforts aim<strong>ed</strong> at the detection of gravitational <strong>waves</strong> (GW) start<strong>ed</strong> more thana quarter of a century ago and have been, so far, unsuccessful [1,2]. Resonant barshave prov<strong>ed</strong> their reliability, being capable of continous data gathering for longperiods of time [3, 4]. Their energy sensitivity has improv<strong>ed</strong> by more than fourorders of magnitude since Weber’s pioneering experiment. However, a furtherimprovement is still necessary to achieve successful detection. While furtherdevelopments of bar detectors are underway, two new generations of Earth-bas<strong>ed</strong>experiments have been propos<strong>ed</strong>: detectors bas<strong>ed</strong> on large laser interferometersare already under construction [5] and resonant detectors of spherical shape areunder study [2].In this lecture I report on a series of papers [6] in which the opportunity ofintroducing resonant mass detectors of spherical shape was studi<strong>ed</strong>. As a generalmotivation for their study, spherical detectors have the advantage over bar-shap<strong>ed</strong>detectors of a larger degree of symmetry. This translates into the possibility ofbuilding detectors of greater mass and consequently of higher cross section.Besides this obvious observation, the higher degree of symmetry enjoy<strong>ed</strong> bythe spherical shape puts such a detector in the unique position of being able to152


Introduction 153detect GWs with a spin content different from two. This is a means of testingnon-Einsteinian theories of gravity.I would now like to remind the reader of the very special position ofEinstein’s general relativity (GR) among the possible gravitational theories.Theories of gravitation, in fact, can be divid<strong>ed</strong> into two families: metric andnon-metric theories [7]. The former can be defin<strong>ed</strong> to be all theories obeyingthe following three postulates:• spacetime is endow<strong>ed</strong> with a metric;• the world lines of test particles are geodesic of the above-mention<strong>ed</strong> metric;• in local free-falling frames, the non-gravitational laws of physics are thoseof special relativity.It is an obvious consequence of these postulates that a metric theory obeys theprinciple of equivalence. More succinctly a theory is said to be metric if the actionof gravitation on the matter sector is due exclusively to the metric tensor. GR isthe most famous example of a metric theory. Kaluza–Klein-type theories, alsobelong to this class along with the Brans–Dicke theory. Different representativesof this class differ by their equations of motion which in turn can be d<strong>ed</strong>uc<strong>ed</strong>from a Lagrangian principle. Since there seems to be no compelling experimentalor theoretical reasons to introduce non-Einsteinian or non-metric theories, theyare sometimes consider<strong>ed</strong> a curiosity. This point should perhaps be reconsider<strong>ed</strong>.String theories are, in fact, the most serious candidate for a theory of quantumgravity, the standard cosmological model has been emend<strong>ed</strong> with the introductionof inflation and even the introduction of a cosmological constant (which seems tobe ne<strong>ed</strong><strong>ed</strong> to explain recent cosmological data) could imply the existence of othergravitationally coupl<strong>ed</strong> fields. In all of the above cit<strong>ed</strong> cases we are forc<strong>ed</strong> tointroduce fields which are non-metrically coupl<strong>ed</strong> in the sense explain<strong>ed</strong> above.In the first section of this lecture we will explain that a spherical detector isable to detect any spin component of an impinging GW. Moreover, its vibrationaleigenvalues can be divid<strong>ed</strong> into two sets call<strong>ed</strong> spheroidal and toroidal. Onlythe first set couples to the metric. This leads to the opportunity of using such adetector as a veto for non-Einsteinian theories. In the second section we take asa model the Jordan–Brans–Dicke (JBD), in which along with the metric we alsohave a scalar field which is metrically coupl<strong>ed</strong>. We are then able to study thesignal-to-noise ratio for sources such as binary systems and collapsing stars andcompare the strength of the scalar signal with respect to the tensor one. Finally,in the third and last section we repeat this computation in the case of the hollowsphere which seems to be the detector which is most likely to be built.


154 Detection of scalar gravitational <strong>waves</strong>11.2 Testing theories of gravity11.2.1 Free vibrations of an elastic sphereBefore discussing the interaction with an external GW field, let us considerthe basic equations governing the free vibrations of a perfectly homogeneous,isotropic sphere of radius R, made of a material having density ρ and Lamécoefficients λ and µ [8].Following the notation of [9], let x i , i = 1, 2, 3 be the equilibrium position ofthe element of the elastic sphere and xi ′ be the deform<strong>ed</strong> position, then u i = xi ′ −x iis the displacement vector. Such vector is assum<strong>ed</strong> small, so that the linear theoryof elasticity is applicable. The strain tensor is defin<strong>ed</strong> as u ij = (1/2)(u i, j + u j,i )and is relat<strong>ed</strong> to the stress tensor by σ ij = δ ij λu ll + 2µu ij . The equations ofmotion of the free vibrating sphere are thus [8]ρ ∂2 u i∂t 2 = ∂∂x j (δ ijλu ll + 2µu ij ) (11.1)with the boundary condition:n j σ ij = 0 (11.2)at r = R where n i ≡ x i /r is the unit normal. These conditions simplystate that the surface of the sphere is free to vibrate. The displacement u i is atime-dependent vector, whose time dependence can be factoriz<strong>ed</strong> as u i (⃗x, t) =u i (⃗x) exp(iωt), where ω is the frequency. The equations of motion then become:µ∇ 2 u i + (λ + µ)∇ i (∇ j u j ) =−ω 2 ρu i . (11.3)Their solutions can be express<strong>ed</strong> as a sum of a longitudinal and two transversevectors [10]:⃗u(⃗x) = C 0⃗∇φ(⃗x) + C 1⃗Lχ(⃗x) + C 2⃗∇×⃗Lχ(⃗x) (11.4)where C 0 , C 1 , C 2 are dimension<strong>ed</strong> constants and ⃗L ≡ ⃗x × ⃗∇ is the angularmomentum operator. Regularity at r = 0 restricts the scalar functions φ and χ tobe express<strong>ed</strong> as φ(r,θ,ϕ) ≡ j l (qr)Y lm (θ, ϕ) and χ(r,θ,ϕ) ≡ j l (kr)Y lm (θ, ϕ).Y lm (θ, ϕ) are the spherical harmonics and j l the spherical Bessel functions [11]:( ) 1 d l ( ) sin xj l (x) =x dx x(11.5)q 2 ≡ ρω 2 /(λ + 2µ) and k 2 ≡ ρω 2 /µ are the longitudinal and transversewavevectors, respectively.Imposing the boundary conditions (11.2) at r = R yields two families ofsolutions:


Testing theories of gravity 155• Toroidal modes: these are obtain<strong>ed</strong> by setting C 0 = C 2 = 0, and C 1 ̸= 0. Inthis case the displacements in (11.4) can be written in terms of the basis:⃗ψ T nlm (r,θ,ϕ)= T nl(r) ⃗LY lm (θ, ϕ) (11.6)with T nl (r) proportional to j l (k nl r). The eigenfrequencies are determin<strong>ed</strong> bythe boundary conditions (11.2) which read [10]wheref 1 (kR) = 0 (11.7)f 1 (z) ≡ d dz[jl (z)z]. (11.8)• Spheroidal modes: these are obtain<strong>ed</strong> by setting C 1 = 0, C 0 ̸= 0 andC 2 ̸= 0. The displacements of (11.4) can be expand<strong>ed</strong> in the basis⃗ψ S nlm (⃗x) = A nl(r)Y lm (θ, ϕ)⃗n − B nl (r)⃗n × ⃗LY lm (θ, ϕ) (11.9)where A nl (r) and B nl (r) are dimensionless radial eigenfunctions [9], whichcan be express<strong>ed</strong> in terms of the spherical Bessel functions and theirderivatives. The eigenfrequencies are determin<strong>ed</strong> by the boundary conditions(11.2) which read [9](f2 (qR) −2µ λdetq2 R 2 )f 0 (qR) l(l + 1) f 1 (kR)1f 1 (qR)2f 2 (kR) + [ l(l+1)= 02− 1] f 0 (kR)(11.10)wheref 0 (z) ≡ j l(z)z 2 , f 2(z) ≡ d2dz 2 j l(z). (11.11)The eigenfrequencies can be determin<strong>ed</strong> numerically for both toroidal andspheroidal vibrations. Each mode of order l is (2l + 1)-fold degenerate. Theeigenfrequency values can be obtain<strong>ed</strong> from√ µω nl =ρ(kR) nlR . (11.12)11.2.2 Interaction of a metric GW with the sphere vibrational modesThe detector is assum<strong>ed</strong> to be non-relativistic (with sound velocity v s ≪ c andradius R ≪ λ the GW wavelength) and endow<strong>ed</strong> with a high quality factor(Q nl = ω nl τ nl ≫ 1, where τ nl is the decay time of the mode nl). Th<strong>ed</strong>isplacement ⃗u of a point in the detector can be decompos<strong>ed</strong> in normal modesas:⃗u(⃗x, t) = ∑ A N (t) ⃗ψ N (⃗x) (11.13)N


156 Detection of scalar gravitational <strong>waves</strong>where N collectively denotes the set of quantum numbers identifying the mode.The basic equation governing the response of the detector is [12]Ä N (t) + τN −1 ȦN (t) + ω 2 N A N (t) = f N (t). (11.14)We assume that the gravitational interaction obeys the principle of equivalencewhich has been experimentally support<strong>ed</strong> to high accuracy. In terms of the socall<strong>ed</strong>electric components of the Riemann tensor E ij ≡ R 0i0 j , the driving forcef N (t) is then given by [13]∫f N (t) =−M −1 E ij (t)ψ i∗N (⃗x)x j ρ d 3 x (11.15)where M is the sphere mass and we consider the density ρ as a constant. In anymetric theory of gravity E ij is a (3×3) symmetric tensor, which depends on time,but not on spatial coordinates.Let us now investigate which sphere eigenmodes can be excit<strong>ed</strong> by a metricGW, i.e. which sets of quantum numbers N give a non-zero driving force.(a) Toroidal modesThe eigenmode vector, ψnlm T can be express<strong>ed</strong> as in equation (11.6). Up to anadimensional normalization constant C, the driving force is∫f (T)N (t) =−e−iω N t 3C R∫ π4π R 3 drr 3 j l (k (T)nlr) dθ sin θ00∫ 2π{ (Eyy − E xx× dφsin θ sin 2φ ∂Y lm∗02∂θ+ cos θ cos 2φ ∂Y lm∗ )∂φ(+ E xy sin θ cos 2φ ∂Y lm∗− cos θ sin 2φ ∂Y ∗ )lm∂θ∂φ[+ E xz − sin φ cos θ ∂Y lm∗+ (sin θ cos φ − cos2 θ∗ ]∂θsin θ cos φ)∂Y lm∂φ[+ E yz cos φ cos θ ∂Y lm∗+ (sin θ sin φ − cos2 ∗ ]θlmsin φ)∂Y∂θsin θ ∂φ()+cos θ ∂Y ∗ }. (11.16)Using the equationsE zz − E xx + E yy2lm∂φand∂Y ∗lm∂θ∂Y ∗lm= (−) m [ 2l + 14π] 1(l − m)! 2 ∂ Pm(l + m)![ 2l + 1 (l − m)!∂φ =−im(−)m 4π (l + m)!l(cos θ)e −imφ (11.17)∂θ] 12Pml (cos θ)e −imφ (11.18)


0Testing theories of gravity 157the integration over φ can be perform<strong>ed</strong>. Equation (11.16) then contains integralsover θ of the form:∫ π[(sin 2 θ − cos 2 θ)Pl ±1 (cos θ)− sin θ cos θ ∂ ]P±1 l(cos θ)dθ (11.19)∂θand∫ π0[2 sin θ cos θ Pl ±2 (cos θ)+ sin 2 θ ∂ P±2∂θl(cos θ)]dθ. (11.20)After integration by parts, the derivative terms in equations (11.19) and (11.20)exactly cancel the non-derivative ones. The remaining boundary terms vanish too,thanks to the periodicity of the trigonometric functions and to the regularity of theassociat<strong>ed</strong> Legendre polynomials. The vanishing of these integrals has a profoundphysical consequence. It means that in any metric theory of gravity the toroidalmodes of the sphere cannot be excit<strong>ed</strong> by GW and can thus be us<strong>ed</strong> as a veto inthe detection.(b) Spheroidal modesThe forcing term is given by:∫f (S)N(t) =−M−1 E ij (t)− B N (r)e ink x nr L kY lm (θ, ϕ)( xx j ir A N (r)Y lm (θ, ϕ))ρ d 3 x. (11.21)One is thus l<strong>ed</strong> to compute integrals of the following types∫x j x i Y lm (θ, ϕ) d 3 x (11.22)and∫x j x i L k Y lm (θ, ϕ) d 3 x (11.23)Since the product x i x j can be express<strong>ed</strong> in terms of the spherical harmonics withl = 0, 2 and the angular momentum operator does not change the value of l,one imm<strong>ed</strong>iately concludes that in any metric theory of gravity only the l = 0, 2spheroidal modes of the sphere can be excit<strong>ed</strong>. At the lowest level there are a totalof five plus one independent spheroidal modes that can be us<strong>ed</strong> for GW detectionand study.11.2.3 Measurements of the sphere vibrations and wave polarization statesFrom the analysis of the spheroidal modes active for metric GW, we now want toinfer the field content of the theory. For this purpose it is convenient to expressthe Riemann tensor in a null (Newman–Penrose) tetrad basis [7].


158 Detection of scalar gravitational <strong>waves</strong>To lowest non-trivial order in the perturbation the six independent ‘electric’components of the Riemann tensor may be express<strong>ed</strong> in terms of the Newman–Penrose (NP) parameters asE ij =( − Re 4 − 22 Im 4 −2 √ 2Re 3Im 4 Re 4 − 22 2 √ )2Im 3−2 √ 2Re 3 2 √ . (11.24)2Im 3 −6 2The NP parameters allow the identification of the spin content of the metrictheory responsible for the generation of the wave [7]. The classification can besummariz<strong>ed</strong> in order of increasing complexity as follows:• General relativity (spin 2): 4 ̸= 0 while 2 = 3 = 22 = 0.• Tensor–scalar theories (spin 2 and 0): 4 ̸= 0, 3 = 0, 2 ̸= 0 and/or 22 ̸= 0 (e.g. Brans–Dicke theory with 4 ̸= 0, 2 = 0, 3 = 0 and 22 ̸= 0).• Tensor–vector theories (spin 2 and 1): 4 ̸= 0, 3 ̸= 0, 22 = 2 = 0.• Most general metric theory (spin 2, 1 and 0): 4 ̸= 0, 2 ̸= 0, 3 ̸= 0 and 22 ̸= 0, (e.g. Kaluza–Klein theories with 4 ̸= 0, 3 ̸= 0, 22 ̸= 0 while 2 = 0).In equation (11.24), we have assum<strong>ed</strong> that the wave comes from a localiz<strong>ed</strong>source with wavevector ⃗k parallel to the z-axis of the detector frame. In thiscase the NP parameters (and thus the wave polarization states) can be uniquelydetermin<strong>ed</strong> by monitoring the six lowest spheroidal modes. If the direction of theincoming wave is not known two more unknowns appear in the problem, i.e. thetwo angles of rotation of the detector frame ne<strong>ed</strong><strong>ed</strong> to align ⃗k along the z-axis. Inorder to dispose of this problem one can envisage the possibility of combining thepieces of information from an array of detectors [14]. We restrict our attention tothe simplest case in which the source direction is known.In order to infer the value of the NP parameters from the measurementsof the excit<strong>ed</strong> vibrational modes of the sphere, we decompose E ij in terms ofspherical harmonics. In fact, the experimental measurements give the vibrationalamplitudes of the sphere modes which are also naturally expand<strong>ed</strong> in the abovebasis. The use of the same basis makes the connection between the NP parametersand the measur<strong>ed</strong> amplitudes straightforward. In formulaeE ij (t) = ∑ l,mc l,m (t)S (l,m)ij(11.25)where S (0,0)ij≡ δ ij / √ 4π (with δ ij the Kronecker symbol) and S (2,m)ij(m =−2,...,2) are five linearly independent symmetric and traceless matrices suchasS (l,m)ijn i n j = Y lm , l = 0, 2. (11.26)The vector n i in equations (11.26) has been defin<strong>ed</strong> after equation (11.2).


Taking the scalar product we find<strong>Gravitational</strong> wave radiation in the JBD theory 159c 0,0 (t) = 4π 3 S0,0 ijE ij (t)c 2,m (t) = 8π15 S2,m ijE ij (t) (11.27)and then for the NP parameters√ √51 22 =16π c 2,0(t) −4π c 0,0(t), 2 =− 1 √512 π c 2,0(t) − 1√112 π c 0,0(t)√ √1515Re 4 =−32π [c 2,2 + c 2,−2 ], Im 4 =−i32π [c 2,2 − c 2,−2 ]Re 3 = 1√1516 π [c 2,1 − c 2,−1 ], Im 3 = i√1516 π [c 2,1 + c 2,−1 ]. (11.28)Equations (11.28) relate the measurable quantities c l,m with the GW polarizationstates, describ<strong>ed</strong> by the NP parameters. Equations (11.28) can be put incorrespondence with the output of experimental measurements if c l,m aresubstitut<strong>ed</strong> with their Fourier components at the quadrupole and monopoleresonant frequencies which, for the sake of simplicity, we collectively denote byω 0 . c l,m (ω 0 ) can be determin<strong>ed</strong> in the following way: once the Fourier amplitudesA N (ω 0 ) are measur<strong>ed</strong>, by Fourier transforming (11.14) and (11.15) we get theRiemann amplitudes E ij (ω 0 ) which, using (11.27), yield the desir<strong>ed</strong> result.In order to determine the A N (ω 0 ) amplitudes from a given GW signal thefollowing two conditions must be fulfill<strong>ed</strong>:• the vibrational states of the five-fold degenerate quadrupole and monopolemodes must be determin<strong>ed</strong>. The quadrupole modes can be studi<strong>ed</strong> byproperly combining the outputs of a set of at least five motion sensorsplac<strong>ed</strong> in independent positions on the sphere surface. Explicit formulas forpractical and elegant configurations of the motion sensors have been report<strong>ed</strong>by various authors [15, 16]. The vibrational state of the monopole modeis provid<strong>ed</strong> directly by the output of any of the above-mention<strong>ed</strong> motionsensors. If resonant motion sensors are us<strong>ed</strong>, since the quadrupole andmonopole states resonate at different frequencies, a sixth sensor is ne<strong>ed</strong><strong>ed</strong>.• The spectrum of the GW signal must be sufficiently broadband to overlapwith the antenna quadrupole and monopole frequencies.11.3 <strong>Gravitational</strong> wave radiation in the JBD theoryIn this section we analyse the signal emitt<strong>ed</strong> by a compact binary system inthe JBD theory. We compute the scalar and tensor components of the powerradiat<strong>ed</strong> by the source and study the scalar waveform. Eventually we considerthe detectability of the scalar component of the radiation by interferometers andresonant-mass detectors.


160 Detection of scalar gravitational <strong>waves</strong>11.3.1 Scalar and Tensor GWs in the JBD TheoryIn the Jordan–Fierz frame, in which the scalar field mixes with the metric butdecouples from matter, the action reads [17]S = S grav [φ,g µν ] + S m [ψ m , g µν ]∫= c3d 4 x √ [−g16πφ R − ω BDφ]gµν ∂ µ φ∂ ν φ + 1 ∫cd 4 xL m [ψ m , g µν ](11.29)where ω BD is a dimensionless constant, whose lower bound is fix<strong>ed</strong> to be ω BD ≈600 by experimental data [18], g µν is the metric tensor, φ is a scalar field and ψ mcollectively denotes the matter fields of the theory.As a preliminary analysis, we perform a weak-field approximation aroundthe background given by a Minkowskian metric and a constant expectation valuefor the scalar fieldg µν = η µν + h µνϕ = ϕ 0 + ξ. (11.30)The standard parametrization ϕ 0 = 2(ω BD +2)/G(2ω BD +3), with G the Newtonconstant, reproduces GR in the limit ω BD → ∞, which implies ϕ 0 → 1/G.Defining the new fieldθ µν = h µν − 1 2 η µνh − η µνξϕ 0(11.31)where h is the trace of the fluctuation h µν , and choosing the gaugeone can write the field equations in the following formwhereTS =−2(2ω BD + 3)∂ µ θ µν = 0 (11.32)∂ α ∂ α θ µν = − 16π τ µνϕ 0(11.33)∂ α ∂ α 8πξ =2ω BD + 3 S (11.34)τ µν = 1 (T µν + t µν ) (11.35)ϕ 0[ 12 ∂ α(θ∂ α ξ)+ 2 ∂ a (ξ∂ ξ)]α .ϕ 0(1 − 1 2 θ − 2 ξ ϕ 0)− 116π(11.36)


<strong>Gravitational</strong> wave radiation in the JBD theory 161In equation (11.35), T µν is the matter stress–energy tensor and t µν is thegravitational stress–energy pseudotensor, that is a function of quadratic order inthe weak gravitational fields θ µν and ξ. The reason why we have written the fieldequations at the quadratic order in θ µν and ξ is that in this way, as we will seelater, the expressions for θ µν and ξ include all the terms of order (v/c) 2 , where vis the typical velocity of the source (Newtonian approximation).Let us now compute τ 00 and S at the order (v/c) 2 . Introducing theNewtonian potential U produc<strong>ed</strong> by the rest-mass density ρ∫ ρ(⃗x ′ , t)U(⃗x, t) =|⃗x −⃗x ′ | d3 x ′ (11.37)the total pressure p and the specific energy density (that is the ratio of energydensity to rest-mass density) we get (for a more detail<strong>ed</strong> derivation, see [7]):τ 00 = 1 ρ, (11.38)ϕ 0TS ≃−(1 − 1 2(2ω BD + 3) 2 θ − 2 ξ )ϕ 0(ρ=1 + − 3 p 2(2ω BD + 3ρ + 2ω )BD + 1ω BD + 2 U . (11.39)Far from the source, equations (11.33) and (11.34) admit wave-like solutions,which are superpositions of terms of the formθ µν (x) = A µν (⃗x,ω)exp(ik α x α ) + c.c. (11.40)ξ(x) = B(⃗x,ω)exp(ik α x α ) + c.c. (11.41)Without affecting the gauge condition (11.32), one can impose h =−2ξ/ϕ 0 (sothat θ µν = h µν ). Gauging away the superflous components, one can write theamplitude A µν in terms of the three degrees of fre<strong>ed</strong>om corresponding to stateswith helicities ±2 and 0 [19]. For a wave travelling in the z-direction, one thusobtains⎛⎞0 0 0 0⎜ 0 eA µν = 11 − b e 12 0 ⎟⎝⎠ , (11.42)0 e 12 −e 11 − b 00 0 0 0where b = B/ϕ 0 .11.3.2 Power emitt<strong>ed</strong> in GWsThe power emitt<strong>ed</strong> by a source in GWs depends on the stress–energy pseudotensort µν according to the following expression∫∫P GW = r 2 d = r 2 〈t 0k 〉ˆx k d (11.43)


162 Detection of scalar gravitational <strong>waves</strong>where r is the radius of a sphere which contains the source, is the solid angle, is the energy flux and the symbol 〈···〉implies an average over a region of sizemuch larger than the wavelength of the GW. At the quadratic order in the weakfields we find〈t 0z 〉=−ẑ ϕ [ ]0c 4 4(ωBD + 1)〈(∂ 0 ξ)(∂ 0 ξ)〉+〈(∂ 0 h αβ )(∂ 0 h αβ )〉 . (11.44)32πϕ 2 0Substituting (11.40) and (11.41) into (11.44), one gets〈t 0z 〉=−ẑ ϕ [0c 4 ω 2 2(2ωBD + 3)|B| 2 + A αβ∗ A αβ − 1 ]16π2 |Aα α| 2 , (11.45)and using (11.42)〈t 0z 〉=−ẑ ϕ 0c 4 ω 28πϕ 2 0[|e 11 | 2 +|e 12 | 2 + (2ω BD + 3)|b| 2] . (11.46)From (11.46) we see that the purely scalar contribution, associat<strong>ed</strong> with b and thetraceless tensorial contribution, associat<strong>ed</strong> with e µν , are completely decoupl<strong>ed</strong>and can thus be treat<strong>ed</strong> independently.11.3.3 Power emitt<strong>ed</strong> in scalar GWsWe now rewrite the scalar wave solution (11.41) in the following wayξ(⃗x, t) = ξ(⃗x,ω)e −iωt + c.c. (11.47)In vacuo, the spatial part of the previous solution (11.47) satisfies the Helmholtzequation(∇ 2 + ω 2 )ξ(⃗x,ω)= 0. (11.48)The solution of (11.48) can be written asξ(⃗x,ω)= ∑ jmX jm h (1)j(ωr)Y jm (θ, ϕ) (11.49)where h (1)j(x) are the spherical Hankel functions of the first kind, r is the distanceof the source from the observer, Y jm (θ, ϕ) are the scalar spherical harmonicsand the coefficients X jm give the amplitudes of the various multipoles whichare present in the scalar radiation field. Solving the inhomogeneous waveequation (11.34), we find∫X jm = 16πiω j l (ωr ′ )Ylm ∗ (θ, ϕ)S(⃗x,ω)dV (11.50)Vwhere j l (x) are the spherical Bessel functions and r ′ is a radial coordinate whichassumes its values in the volume V occupi<strong>ed</strong> by the source.


<strong>Gravitational</strong> wave radiation in the JBD theory 163Substituting (11.44) in (11.43), considering the expressions (11.47) and(11.49) and averaging over time, one finally obtainsP scal = (2ω BD + 3)c 48πϕ 0∑|X jm | 2 . (11.51)To compute the power radiat<strong>ed</strong> in scalar GWs, one has to determine thecoefficients X jm ,defin<strong>ed</strong> in (11.50). The detail<strong>ed</strong> calculations can be foundin appendix A of the third reference in [6], while here we only give the finalresults. Introducing the r<strong>ed</strong>uc<strong>ed</strong> mass of the binary system µ = m 1 m 2 /m and thegravitational self-energy for the body a (with a = 1, 2) a =− 1 ∫ρ(⃗x)ρ(⃗x ′ )2 V a|⃗x −⃗x ′ d 3 x d 3 x ′ (11.52)|one can write the Fourier components with frequency nω 0 in the Newtonianapproximation(X 00 ) n =− 16√ 2π iω 0 ϕ 0 mµ3 ω BD + 2 a nJ n(ne) (11.53)√2π 2iω 2 (0 ϕ 0 2(X 1±1 ) n =−− )1µa3 ω BD + 2 m 2 m 1[× ±J n ′ (ne) − 1 ]e (1 − e2 ) 1/2 J n (ne)(11.54)(X 20 ) n = 2 √ π iω 3 0 ϕ 03 5 ω BD + 2 µa2 nJ n (ne) (11.55)√ π iω 3 0 ϕ 0(X 2±2 ) n =∓230 ω BD + 2 µa2× 1 n {(e2 − 2)J n (ne)/(ne 2 ) + 2(1 − e 2 )J ′ n (ne)/ejm∓ 2(1 − e 2 ) 1/2 [(1 − e 2 )J n (ne)/e 2 − J n ′ (ne)/(ne)]}. (11.56)Substituting these expressions in (11.51), leads to the power radiat<strong>ed</strong> in scalarGWs in the nth harmonic(P scal ) n = P j=0n + P j=1n + P j=2n (11.57)where the monopole, dipole and quadrupole terms are, respectively,Pn j=0 64 m 3 µ 2 G 4=9(ω BD + 2) a 5 c 5 n 2 Jn 2 (ne)64 m 3 µ 2 G 4=9(ω BD + 2) a 5 c 5 m(n; e) (11.58)


164 Detection of scalar gravitational <strong>waves</strong>P j=1n =4 m 2 µ 2 G 3 (23(ω BD + 2) a 4 c 3 − ) 21m 2 m 1× n 2 [J ′2n (ne) + 1 e 2 (1 − e2 )J 2 n (ne) ]4 m 2 µ 2 G 3 (2=3(ω BD + 2) a 4 c 3 − ) 21d(n; e) (11.59)m 2 m 1Pn j=2 8 m 3 µ 2 G 4=15(ω BD + 2) a 5 c 5 g(n; e). (11.60)The total power radiat<strong>ed</strong> in scalar GWs by a binary system is the sum ofthree termsP scal = P j=0 + P j=1 + P j=2 (11.61)whereP j=0 16=9(ω BD + 2)P j=1 =2ω BD + 2G 4 m 2 1 m2 2 m( )e 2c 5 a 5 (1 − e 2 ) 7/2 1 + e24(2− ) (21 G 3 m 2 1 m2 2 1m 2 m 1 c 3 a 4 (1 − e 2 ) 5/2 1 + e22)(11.62)(11.63)P j=2 8 G 4 m 2 1=m2 2 m(115(ω BD + 2) c 5 a 5 (1 − e 2 ) 7/2 1 + 7324 e2 + 37 )96 e4 . (11.64)Note that P j=0 , P j=1 , P j=2 all go to zero in the limit ω BD →∞.11.3.4 Scalar GWsWe now give the explicit form of the scalar GWs radiat<strong>ed</strong> by a binary system. Tothis end, note that the major semi-axis, a, is relat<strong>ed</strong> to the total energy, E, ofthesystem through the following equationa =− Gm 1m 22E . (11.65)Let us consider the case of a circular orbit, remembering that in the last phaseof evolution of a binary system this condition is usually satisfi<strong>ed</strong>. Furthermorewe will also assume m 1 = m 2 . With these positions only the quadrupole term,(11.60), of the gravitational radiation is different from zero. The total powerradiat<strong>ed</strong> in GWs, averag<strong>ed</strong> over time, is then given by (11.62)–(11.64)P =8 G 4 m 2 1 m2 2 m15(ω BD + 2) c 5 d 5 [6(2ω BD + 3) + 1] (11.66)where d is the relative distance between the two stars. The time variation of d inone orbital period isḋ =− Gm 1m 22E 2 P (11.67)


<strong>Gravitational</strong> wave radiation in the JBD theory 165Finally, substituting (11.65), (11.66) in (11.67) and integrating over time, oneobtains( ) 1/42 12ω BD + 19 G 3 m 1 m 2 md = 215 ω BD + 2 c 5 τ 4 (11.68)where we have defin<strong>ed</strong> τ = t c − t, t c being the time of the collapse between thetwo bodies.From (11.49), (11.53)–(11.56) one can d<strong>ed</strong>uce the form of the scalar field(see appendix B of the third reference in [6] for details) which, for equal masses,is2µ [ξ(t) =−v 2 + m r(2ω BD + 3) d − (ˆn ·⃗v)2 + m ]d 3 (ˆn · ⃗d) (11.69)where r is the distance of the source from the observer, and ˆn is the versor of theline of sight from the observer to the binary system centre of mass. Indicatingwith γ the inclination angle, that is the angle between the orbital plane and thereference plane (defin<strong>ed</strong> to be a plane perpendicular to the line of sight), and withψ the true anomaly, that is the angle between d and the x-axis in the orbital planex–y, yields ˆn · ⃗d = d sin γ sin ψ. Then, from (11.69) one obtains2Gµmξ(t) =(2ω BD + 3)c 4 dr sin2 γ cos(2ψ(t)) (11.70)which can also be written asξ(τ) = ξ 0 (τ) sin(χ(τ) +¯χ) (11.71)where ¯χ is an arbitrary phase and the amplitude ξ 0 (τ) is given by2Gµmξ 0 (τ) =(2ω BD + 3)c 4 dr sin2 γ( )1 ωBD + 2 1/4 ( ) 15G 1/4 5/4M c=2(2ω BD + 3)r 12ω BD + 19 2c 11 τ 1/4 sin2 γ. (11.72)In the last expression, we have introduc<strong>ed</strong> the definition of the chirp massM c = (m 1 m 2 ) 3/5 /m 1/5 .11.3.5 Detectability of the scalar GWsLet us now study the interaction of the scalar GWs, a spherical GW detector.As usual, we characterize the sensitivity of the detector by the spectraldensity of strain S h ( f ) [Hz] −1 . The optimum performance of a detector isobtain<strong>ed</strong> by filtering the output with a filter match<strong>ed</strong> to the signal. The energysignal-to-noise ratio (SNR) of the filter output is given by the well known formula:SNR =∫ +∞−∞|H ( f )| 2S h ( f ) d f (11.73)


166 Detection of scalar gravitational <strong>waves</strong>where H ( f ) is the Fourier transform of the scalar gravitational waveform h s (t) =Gξ 0 (t).We must now take into account the astrophysical restrictions on the validityof the waveform (11.71) which is obtain<strong>ed</strong> in the Newtonian approximation forpoint-like masses. In the following, we will take the point of view that thisapproximation breaks down when there are five cycles remaining to collapse[20, 21].The five-cycles limit will be us<strong>ed</strong> to restrict the range of M c over which ouranalysis will be perform<strong>ed</strong>. From (11.68), one can obtain√Gmω g (τ) = 2ω 0 = 2d 3( ) 3/815c5 ( )ωBD + 2 3/81= 264G 5/3 12ω BD + 19 M5/8 τ 3/8 . (11.74)cIntegrating (11.74) yields the amount of phase until coalescenceχ(τ) = 16( ) 3/815c5 ( )ωBD + 2 3/8 ( ) τ 5/85 64G 5/3 . (11.75)12ω BD + 19 M cSetting (11.75) equal to the limit period, T 5 cycles = 5(2π), solving for τ andusing (11.74) leads to( )ωBD + 2 3/5M ⊙ω 5 cycles = 2π(6870 Hz). (11.76)12ω BD + 19 M cTaking ω BD = 600, the previous limit readsω 5 cycles = 2π(1547 Hz) M ⊙. (11.77)M cA GW excites those vibrational modes of a resonant body having the propersymmetry. In the framework of JBD theory the spheroidal modes with l = 2 andl = 0 are sensitive to the incoming GW. Thanks to its multimode nature, a singlesphere is capable of detecting GWs from all directions and polarizations. We nowevaluate the SNR of a resonant-mass detector of spherical shape for its quadrupolemode with m = 0 and its monopole mode. In a resonant-mass detector, S h ( f ) isa resonant curve and can be characteriz<strong>ed</strong> by its value at resonance S h ( f n ) and byits half height width [22]. S h ( f n ) can thus be written asS h ( f n ) = G c 34kTσ n Q n f n. (11.78)Here, σ n is the cross section associat<strong>ed</strong> with the nth resonant mode, T is thethermodynamic temperature of the detector and Q n is the quality factor of themode.


<strong>Gravitational</strong> wave radiation in the JBD theory 167The half height width of S h ( f ) gives the bandwidth of the resonant modef n =f nƔn −1/2 . (11.79)Q nHere, Ɣ n is the ratio of the wideband noise in the nth resonance bandwidthto the narrowband noise.From the resonant-mass detector viewpoint, the chirp signal can be treat<strong>ed</strong>as a transient GW, depositing energy in a timescale short with respect to th<strong>ed</strong>etector damping time. We can then consider constant the Fourier transform ofthe waveform within the band of the detector and write [22]SNR = 2πf n|H ( f n )| 2. (11.80)S h ( f n )The cross sections associat<strong>ed</strong> with the vibrational modes with l = 0 andl = 2, m = 0 are respectively [6]GMv2 sσ (n0) = H nc 3 (11.81)(ω BD + 2)σ (n2) = F n6GMv s2c 3 (ω BD + 2) . (11.82)All parameters entering the previous equation refer to the detector, M is its mass,v s the sound velocity and the constants H n and F n are given in [6]. The signal-tonoise ratio can be calculat<strong>ed</strong> analytically by approximating the waveform with atruncat<strong>ed</strong> Taylor expansion around t = 0, where ω g (t = 0) = ω nl [20, 23]h s (t) ≈ Gξ 0 (t = 0) sin[ω nl t + 1 2( dωdtUsing quantum limit<strong>ed</strong> readout systems, one finally obtains(SNR n ) l=0 =(SNR n ) l=2 =)t=0t 2 ]. (11.83)5 × 2 1/3 H n G 5/3 M 5/3 c Mv2 s32(ω BD + 2)(12ω BD + 19)¯hc 3 r 2 sin 4 γ (11.84)ω 4/3 n05 × 2 1/3 F n G 5/3 M 5/3 c Mv2 s192(ω BD + 2)(12ω BD + 19)¯hc 3 r 2 sin 4 γ (11.85)ω 4/3 n0which are, respectively, the SNR for the modes with l = 0 and l = 2, m = 0ofaspherical detector.It has been propos<strong>ed</strong> to realize spherical detectors with 3 m diameter, madeof copper alloys, with mass of the order of 100 tons [24]. This propos<strong>ed</strong> detectorhas resonant frequencies of ω 12 = 2π ×807 rad s −1 and ω 10 = 2π ×1655 rad s −1 .In the case of optimally orient<strong>ed</strong> orbits (inclination angle γ = π/2) and ω BD =600, the inspiralling of two compact objects of 1.4 solar masses each will thenbe detect<strong>ed</strong> with SNR = 1 up to a source distance r(ω 10 ) ≃ 30 kpc andr(ω 12 ) ≃ 30 kpc.


168 Detection of scalar gravitational <strong>waves</strong>11.4 The hollow sphereAn appealing variant of the massive sphere is a hollow sphere [25]. The latterhas the remarkable property that it enables the detector to monitor GW signals ina significantly lower frequency range—down to about 200 Hz—than its massivecounterpart for comparable sphere masses. This can be consider<strong>ed</strong> a positiveadvantage for a future worldwide network of GW detectors, as the sensitivityrange of such antenna overlaps with that of the large-scale interferometers, now ina rather advanc<strong>ed</strong> state of construction [6,7]. In this section we study the responseof such a detector to the GW energy emitt<strong>ed</strong> by a binary system constitut<strong>ed</strong> of starsof masses of the order of the solar mass. A hollow sphere obviously has the samesymmetry of the massive one, so the general structure of its normal modes ofvibration is very similar [25] to that of the solid sphere. In particular, the hollowsphere is very well adapt<strong>ed</strong> to sense and monitors the presence of scalar modes inthe incoming GW signal. The extension of the analysis of the previous sectionsto a hollow sphere is quite straightforward and in the following we will only givethe main results. Due to the different geometry, the vibrational modes of a hollowsphere differ from those studi<strong>ed</strong> in section 11.2. In the case of a hollow sphere,we have two boundaries given by the outer and the inner surfaces of the soliditself. We use the notation a for the inner radius and R for the outer radius. Theboundary conditions are thus express<strong>ed</strong> byσ ij n j = 0 atr = R and at r = a (R ≥ a ≥ 0), (11.86)(11.3) must now be solv<strong>ed</strong> subject to these boundary conditions. The solutionthat leads to spheroidal modes is still (11.9) where the radial functions A nl (r) andB nl (r) have rather complicat<strong>ed</strong> expressions:A nl (r) = C nl[ 1q S nl+ D nl1q S nlddr j l(qnl S r) − l(l + 1)K j l (knl S r)nlknl S rddr y l(qnl S r) − l(l + 1) y l (knl ˜D S r) ]nlknl S rB nl (r) = C nl[jl (q S nl r)q S nl r − K nl1k S nl r ddr {rj l(k S nl r)}(11.87)+ D nly l (q S nl r)q S nl r − ˜D nl1k S nl r ddr {ry l(k S nl r)} ]. (11.88)Here knl S R and qS nlR are dimensionless eigenvalues, and they are the solutionto a rather complicat<strong>ed</strong> algebraic equation for the frequencies ω = ω nl —see [25]for details. In (11.87) and (11.88) we have setK nl ≡ C tqnlSC l knlS , D nl ≡ qS nlknlS E, ˜D nl ≡ C t FqnlSC l knlS(11.89)


The hollow sphere 169and introduc<strong>ed</strong> the normalization constant C nl , which is fix<strong>ed</strong> by the orthogonalityproperties ∫(u S n ′ l ′ m ′ ) ∗ · (u S nlm )ϱ 0 d 3 x = Mδ nn ′δ ll ′δ mm ′ (11.90)Vwhere M is the mass of the hollow sphere:M = 4π 3 ϱ 0 R 3 (1 − ς 3 ), ς ≡ a R≤ 1. (11.91)Equation (11.90) fixes the value of C nl through the radial integral∫ Rς R[A 2 nl (r) + l(l + 1)B2 nl (r)]r 2 dr = 4π 3 ϱ 0(1 − ς 3 )R 3 (11.92)as can be easily verifi<strong>ed</strong> by using well known properties of angular momentumoperators and spherical harmonics. We shall later specify the values of th<strong>ed</strong>ifferent parameters appearing in the above expressions as requir<strong>ed</strong> in eachparticular case which will in due course be consider<strong>ed</strong>. As seen in [9], ascalar–tensor theory of GWs such as JBD pr<strong>ed</strong>icts the excitation of the sphere’smonopole modes as well as the m = 0 quadrupole modes. In order to calculatethe energy absorb<strong>ed</strong> by the detector according to that theory it is necessary tocalculate the energy deposit<strong>ed</strong> by the wave in those modes, and this in turnrequires that we solve the elasticity equation with the GW driving term includ<strong>ed</strong> inits right-hand side. The result of such a calculation was present<strong>ed</strong> in full generalityin [9], and is directly applicable here because the structure of the oscillationeigenmodes of a hollow sphere is equal to that of the massive sphere—only theexplicit form of the wavefunctions ne<strong>ed</strong>s to be chang<strong>ed</strong>. We thus haveE osc (ω nl ) = 1 2 Mb2 nll∑m=−l|G (lm) (ω nl )| 2 (11.93)where G (lm) (ω nl ) is the Fourier amplitude of the corresponding incoming GWmode, andb n0 = − ϱ ∫ R0A n0 (r)r 3 dr (11.94)M ab n2 = − ϱ ∫ R0[A n2 (r) + 3B n2 (r)]r 3 dr (11.95)M afor monopole and quadrupole modes, respectively, and A nl (r) and B nl (r) aregiven by (11.87). Explicit calculation yieldsb n0R = 3 C n04π 1 − ς 3 [(R) − ς 3 (a)] (11.96)b n2R = 3 C n24π 1 − ς 3 [(R) − ς 3 (a)] (11.97)


170 Detection of scalar gravitational <strong>waves</strong>with(z) ≡ j 2(q n0 z)q n0 R(z) ≡ j 2(q n2 z)q n2 R+ D y 2 (q n0 z)n0q n0 R− 3K n2j 2 (k n2 z)k n2 R(11.98)+ D y 2 (q n2 z)n2q n2 R − 3 y 2 (k n2 z)˜D n2k n2 R . (11.99)The absorption cross section, defin<strong>ed</strong> as the ratio of the absorb<strong>ed</strong> energyto the incoming flux, can be calculat<strong>ed</strong> thanks to an optical theorem, as prov<strong>ed</strong>,for example, by Weinberg [26]. According to that theorem, the absorption crosssection for a signal of frequency ω close to ω N , say, the frequency of the detectormode excit<strong>ed</strong> by the incoming GW, is given by the expressionσ(ω) = 10πηc2ω 2 Ɣ 2 /4(ω − ω N ) 2 + Ɣ 2 /4(11.100)where Ɣ is the linewidth of the mode—which can be arbitrarily small, as assum<strong>ed</strong>in the previous section—and η is the dimensionless ratioη = Ɣ gravƔ= 1 ƔP GWE osc(11.101)where P GW is the energy re-emitt<strong>ed</strong> by the detector in the form of GWs as aconsequence of it being set to oscillate by the incoming signal. In the followingwe will only consider the case P GW = P scalar−tensor with [6, 9]P scalar−tensor =2Gω 6 [5c 5 |Q kk (ω)| 2 + 1 ](2ω BD + 3)3 Q∗ ij (ω)Q ij(ω)(11.102)where Q ij (ω) is the quadrupole moment of the hollow sphere:∫Q ij (ω) = x i x j ϱ(x,ω)d 3 x (11.103)and ω BD is Brans–Dicke’s parameter.Antenna11.5 Scalar–tensor cross sectionsExplicit calculation shows that P scalar−tensor is made up of two contributions:P scalar−tensor = P 00 + P 20 (11.104)where P 00 is the scalar, or monopole contribution to the emitt<strong>ed</strong> power, while P 20comes from the central quadrupole mode which, as discuss<strong>ed</strong> in [6] and [9], isexcit<strong>ed</strong> together with monopole in JBD theory. One must, however, recall that


Scalar–tensor cross sections 171monopole and quadrupole modes of the sphere happen at different frequencies,sothat cross sections for them only make sense if defin<strong>ed</strong> separately. More precisely,σ n0 (ω) = 10πη n0c 2ω 2 Ɣ 2 n0 /4(ω − ω n0 ) 2 + Ɣ 2 n0 /4 (11.105)σ n2 (ω) = 10πη n2c 2ω 2 Ɣ 2 n2 /4(ω − ω n2 ) 2 + Ɣ 2 n2 /4 (11.106)where η n0 and η n2 are defin<strong>ed</strong> as in (11.101), with all terms referring to thecorresponding modes. After some algebra one finds thatGMvS2 Ɣn0 2 σ n0 (ω) = H /4n(ω BD + 2)c 3 (ω − ω n0 ) 2 + Ɣn0 2 /4 (11.107)GMvS2 Ɣn2 2 σ n2 (ω) = F /4n(ω BD + 2)c 3 (ω − ω n2 ) 2 + Ɣn2 2 (11.108)/4.Here, we have defin<strong>ed</strong> the dimensionless quantities4π 2H n =9(1 + σ P ) (k n0b n0 ) 2 (11.109)8π 2F n =15(1 + σ P ) (k n2b n2 ) 2 (11.110)where σ P represents the sphere material’s Poisson ratio (most often very close toa value of 1/3), and b nl are defin<strong>ed</strong> in (11.96); v S is the spe<strong>ed</strong> of sound in thematerial of the sphere.In tables 11.1 and 11.2 we give a few numerical values of the above crosssection coefficients.As already stress<strong>ed</strong> in [25], one of the main advantages of a hollow sphereis that it enables us to reach good sensitivities at lower frequencies than in a solidsphere. For example, a hollow sphere of the same material and mass as a solidone (ς = 0) has eigenfrequencies which are smaller byω nl (ς) = ω nl (ς = 0)(1 − ς 3 ) 1/3 (11.111)for any mode indices n and l. We now consider the detectability of JBD GWcoming from several interesting sources with a hollow sphere.The values of the coefficients F n and H n , together with the expressions(11.105) for the cross sections of the hollow sphere, can be us<strong>ed</strong> to estimatethe maximum distances at which a coalescing compact binary system and agravitational collapse event can be seen with such detector. We consider thesein turn.By taking as a source of GWs a binary system form<strong>ed</strong> by two neutronstars, each of them with a mass of m 1 = m 2 = 1.4M ⊙ . The chirp mass


172 Detection of scalar gravitational <strong>waves</strong>Table 11.1. Eigenvalues kn0 S R, relative weights D n0 and H n coefficients for a hollowsphere with Poisson ratio σ P = 1/3. Values are given for a few different thicknessparameters ς.ς n k S n0 R D n0 H n0.01 1 5.487 38 −0.000 143 328 0.909 291 12.233 2 −0.001 596 36 0.141 942 18.632 1 −0.005 589 61 0.059 264 24.969 3 −0.001 279 0.032 670.10 1 5.454 10 −0.014 218 0.895 301 11.924 1 −0.151 377 0.150 482 17.727 7 −0.479 543 0.049 224 23.541 6 −0.859 885 0.043 110.15 1 5.377 09 −0.045 574 0.860 762 11.387 9 −0.434 591 0.176 463 17.105 −0.939 629 0.056 744 23.605 −0.806 574 0.053 960.25 1 5.048 42 −0.179 999 0.737 272 10.651 5 −0.960 417 0.305 323 17.819 3 −0.425 087 0.042 754 25.806 3 0.440 100 0.063 470.50 1 3.969 14 −0.631 169 0.494 292 13.236 9 0.531 684 0.581 403 25.453 1 0.245 321 0.017 284 37.912 9 0.161 117 0.071 920.75 1 3.265 24 −0.901 244 0.430 702 25.346 8 0.188 845 0.662 843 50.371 8 0.093 173 0.003 414 75.469 0.061 981 0.074 800.90 1 2.981 41 −0.963 552 0.420 432 62.902 7 0.067 342 0.676 893 125.699 0.033 573 0.000 474 188.519 0.022 334 0.075 38corresponding to this system is M c ≡ (m 1 m 2 ) 3/5 (m 1 + m 2 ) −1/5 = 1.22M ⊙ , andν [5 cycles] = 1270 Hz. Repeating the analysis carri<strong>ed</strong> on in section 11.3 we finda formula for the minimum distance at which a measurement can be perform<strong>ed</strong>given a certain SNR, for a quantum limit<strong>ed</strong> detectorr(ω n0 ) =[5 × 21/3321G 5/3 Mc5/3 MvS2( BD + 2)(12 BD + 19) c 3¯hω 4/3n0 SNR H n] 1/2(11.112)


Scalar–tensor cross sections 173Table 11.2. Eigenvalues k S n2 R, relative weights K n2, D n2 , ˜D n2 and F n coefficients for ahollow sphere with Poisson ratio σ P = 1/3. Values are given for a few different thicknessparameters ς.ς n k S n2 R K n2 D n2˜D n2 F n0.10 1 2.638 36 0.855 799 0.000 395 −0.003 142 2.946 022 5.073 58 0.751 837 0.002 351 −0.018 451 1.169 343 10.960 90 0.476 073 0.009 821 −0.071 685 0.022 070.15 1 2.611 61 0.796 019 0.001 174 −0.009 288 2.869 132 5.028 15 0.723 984 0.007 028 −0.053 849 1.241 533 8.258 09 −2.010 150 −0.094 986 0.672 786 0.081 130.25 1 2.491 22 0.606 536 0.003 210 −0.024 94 2.552 182 4.912 23 0.647 204 0.019 483 −0.138 67 1.550 223 8.242 82 −1.984 426 −0.126 671 0.675 06 0.053 254 10.977 25 0.432 548 −0.012 194 0.022 36 0.035 030.50 1 1.943 40 0.300 212 0.003 041 −0.022 68 1.619 782 5.064 53 0.745 258 0.005 133 −0.028 89 2.295 723 10.111 89 1.795 862 −1.697 480 2.982 76 0.197 074 15.919 70 −1.632 550 −1.965 780 −0.309 53 0.171 080.75 1 1.449 65 0.225 040 0.001 376 −0.010 17 1.152 912 5.215 99 0.910 998 −0.197 532 0.409 44 1.822 763 13.932 90 0.243 382 0.748 219 −3.201 30 1.089 524 23.763 19 0.550 278 −0.230 203 −0.817 67 0.081 140.90 1 1.265 65 0.213 082 0.001 019 −0.007 55 1.038 642 4.977 03 0.939 420 −0.323 067 0.522 79 1.541 063 31.864 29 6.012 680 −0.259 533 4.052 74 1.464 864 61.299 48 0.205 362 −0.673 148 −1.043 69 0.134 70r(ω n2 ) =[5 × 21/31921G 5/3 Mc5/3 MvS2( BD + 2)(12 BD + 19) c 3¯hω 4/3n2 SNR F n] 1/2(11.113)For a CuAl sphere, the spe<strong>ed</strong> of sound is v S = 4700 m s −1 . We reportin table 11.3 the maximum distances at which a JBD binary can be seen with a100 ton hollow spherical detector, including the size of the sphere (diameter andthickness factor) for SNR = 1. The Brans–Dicke parameter BD has been givena value of 600. This high value has as a consequence that only relatively nearbybinaries can be scrutiniz<strong>ed</strong> by means of their scalar radiation of GWs. A slightimprovement in sensitivity is appreciat<strong>ed</strong> as the diameter increases in a fix<strong>ed</strong> massdetector. Vacancies in the tables mean the corresponding frequencies are higherthan ν [5 cycles] .The signal associat<strong>ed</strong> with a gravitational collapse can be modell<strong>ed</strong>, within


174 Detection of scalar gravitational <strong>waves</strong>Table 11.3. Eigenfrequencies, sizes and distances at which coalescing binaries can be seenby monitoring of their emitt<strong>ed</strong> JBD GWs. Figures correspond to a 100 ton CuAl hollowsphere.ς (m) ν 10 (Hz) ν 12 (Hz) r(ν 10 ) (kpc) r(ν 12 ) (kpc)0.00 2.94 1655 807 − 29.80.25 2.96 1562 771 − 30.30.50 3.08 1180 578 55 31.10.75 3.5 845 375 64 330.90 4.5 600 254 80 40JBD theory, as a short pulse of amplitude b, whose value can be estimat<strong>ed</strong> as [27]( )( )( )500b ≃ 10 −23 M∗ 10 Mpc(11.114) BD M ⊙ rwhere M ∗ is the collapsing mass.The minimum value of the Fourier transform of the amplitude of the scalarwave, for a quantum limit<strong>ed</strong> detector at unit SNR, is given by( ) 1/24¯h|b(ω nl )| min =MvS 2ω (11.115)nl K nwhere K n = 2H n for the mode with l = 0 and K n = F n /3 for the mode withl = 2, m = 0.The duration of the impulse, τ ≈ 1/f c , is much shorter than the decay timeof the nl mode, so that the relationship between b and b(ω nl ) isb ≈|b(ω nl )| f c at frequency ω nl = 2π f c (11.116)so that the minimum scalar wave amplitude detectable is|b| min ≈(4¯hω nlπ 2 Mv 2 S K n) 1/2. (11.117)Let us now consider a hollow sphere made of molybdenum, for which thespe<strong>ed</strong> of sound is as high as v S = 5600 m s −1 . For a given detector massand diameter, equation (11.117) tells us which is the minimum signal detectablewith such a detector. For example, a solid sphere of M = 31 tons and 1.8 m indiameter, is sensitive down to b min = 1.5 × 10 −22 . Equation (11.114) can thenbe invert<strong>ed</strong> to find which is the maximum distance at which the source can beidentifi<strong>ed</strong> by the scalar <strong>waves</strong> it emits. Taking a reasonable value of BD = 600,one finds that r(ν 10 ) ≈ 0.6 Mpc.


Scalar–tensor cross sections 175Table 11.4. Eigenfrequencies, sizes and distances at which coalescing binaries can be seenby monitoring of their emitt<strong>ed</strong> JBD GWs. Figures correspond to a3mexternal diameterCuAl hollow sphere.ς M (ton) ν 10 (Hz) ν 12 (Hz) r(ν 10 ) (kpc) r(ν 12 ) (kpc)0.00 105 1653 804 — 330.25 103.4 1541 760 — 310.50 92 1212 593 52 27.60.75 60.7 997 442 44.8 230.90 28.4 910 386 32 16.3Table 11.5. Eigenfrequencies, maximum sensitivities and distances at which agravitational collapse can be seen by monitoring the scalar GWs it emits. Figurescorrespond to a 31 ton Mb hollow sphere.ς φ (m) ν 10 (Hz) |b| min (10 −22 ) r(ν 10 ) (Mpc)0.00 1.80 3338 1.5 0.60.25 1.82 3027 1.65 0.50.50 1.88 2304 1.79 0.460.75 2.16 1650 1.63 0.510.90 2.78 1170 1.39 0.6Table 11.6. Eigenfrequencies, maximum sensitivities and distances at which agravitational collapse can be seen by monitoring the scalar GWs it emits. Figurescorrespond to a 1.8 m outer diameter Mb hollow sphere.ς M (ton) ν 10 (Hz) |b| min (10 −22 ) r(ν 10 ) (Mpc)0.00 31.0 3338 1.5 0.60.25 30.52 3062 1.71 0.480.50 27.12 2407 1.95 0.420.75 17.92 1980 2.34 0.360.90 8.4 1808 3.31 0.24Like before, we report, in tables 11.4–11.6, the sensitivities of the detectorand consequent maximum distance at which the source appears visible to th<strong>ed</strong>evice for various values of the thickness parameter ς. In table 11.5 a detector ofmass of 31 tons has been assum<strong>ed</strong> for all thicknesses, and in tables 11.4 and 11.6a constant outer diameter of 3 and 1.8 m has been assum<strong>ed</strong> in all cases.


176 Detection of scalar gravitational <strong>waves</strong>Acknowl<strong>ed</strong>gmentsWe wish to thank very much the organizers of the school for their kind invitationto present this lecture, and all of my collaborators for sharing their insights on thesubject with me.References[1] Thorne K S 1987 Three Hundr<strong>ed</strong> Years of Gravitation <strong>ed</strong> S W Hawking and W Israel(Cambridge: Cambridge University Press)[2] Coccia E, Pizzella G and Ronga F 1995 <strong>Gravitational</strong> wave experiments Proc. FirstEdoardo Amaldi Conf. (Frascati, 1994) (Singapore: World Scientific)[3] Astone P et al 1993 Phys. Rev. D 47 362[4] Johnson W W 1995 Proc. First Edoardo Amaldi Conf. (Frascati, 1994) (Singapore:World Scientific)[5] Abramovici A et al 1992 Science 26 325Bradaschia C et al 1990 Nucl. Instrum. Methods A 289 518[6] Bianchi M, Coccia E, Colacino C N, Fafone V and Fucito F 1996 Class. QuantumGrav. 13 2865Bianchi M, Brunetti M, Coccia E, Fucito F and Lobo J A 1998 Phys. Rev. D 57 4525Brunetti M, Coccia E, Fafone V and Fucito F 1999 Phys. Rev. D 59 044027Coccia E, Fucito F, Lobo J A and Salvino M 2000 Phys. Rev. D 62 044019[7] Will C M 1993 Theory and Experiment in <strong>Gravitational</strong> Physics (Cambridge:Cambridge University Press)[8] Jaerish P 1880 J.f. Math. (Crelle) 88Lamb H 1882 Proc. London Mathematical Society vol 13Love A E H 1927 A Treatise on the Mathematical Theory of Elasticity (London:Cambridge University Press)[9] Lobo J A 1995 Phys. Rev. D 52 591[10] Ashby N and Dreitlein J 1975 Phys. Rev. D 12 336[11] Jackson J D 1975 Classical Electrodynamics (New York: Wiley)[12] Misner C W, Thorne K S and Wheeler J A 1973 Gravitation (New York: Freeman)[13] Wagoner R V and Paik H J 1977 Experimental gravitation Proc. Int. Symp. held inPavia 1976 (Roma: Acc. Naz. dei Lincei)[14] Dhurandhar S V and Tinto M 1989 Proc. V Marcel Grossmann Meeting <strong>ed</strong> D G Blair,M J Buckingham and R Ruffini (Singapore: World Scientific)[15] Zhou C Z and Michelson P F 1995 Phys. Rev. D 51 2517[16] Johnson W W and Merkovitz S M 1993 Phys. Rev. Lett. 70 2367[17] Jordan P 1959 Z. Phys. 157 112Brans C and Dicke R H 1961 Phys. Rev. 124 925[18] Miller J, Schneider M, Soffel M and Ruder H 1991 Astrophys. J. Lett. 382 L101[19] Lee D L 1974 Phys. Rev. D 10 2374[20] Dewey D 1987 Phys. Rev. D 36 1577[21] Coccia E and Fafone V 1996 Phys. Lett. A 213 16[22] Astone P, Pallottino G V and Pizzella G 1997 Class. Quantum Grav. 14 2019[23] Clark J P A and Eardley D M 1977 Astrophys. J. 215 311[24] Frossati G and Coccia E 1994 Cryogenics 34 (ICEC Supplement) 9


References 177Frossati G and Coccia E 1994 Cryogenics 34 (ICEC Supplement) 304[25] Coccia E, Fafone V, Frossati G, Lobo J A and Ortega J A 1998 Phys. Rev. D 57 2051[26] Weinberg S 1972 Gravitation and Cosmology (New York: Wiley)[27] Shibata M, Nakao K and Nakamura T 1994 Phys. Rev. D 50 7304Saijo M, Shinkai H and Ma<strong>ed</strong>a K 1997 Phys. Rev. D 56 785


PART 3THE STOCHASTICGRAVITATIONAL-WAVEBACKGROUNDD Babusci 1 ,SFoffa 2,3,4 , G Losurdo 2,4 , M Maggiore 2,3 ,G Matone 1 and R Sturani 2,3,41 INFN, Laboratori Nazionali di Frascati2 INFN, Sezione di Pisa3 Dipartimento di Fisica, Universita’ di Pisa4 Scuola Normale Superiore di Pisa


Chapter 12Generalities on the stochastic GWbackground12.1 IntroductionThe stochastic gravitational-wave background (SGWB) is a random noise ofgravity <strong>waves</strong> with no evidence of any sharp specific characters in either the timeor frequency domain. The important fact to note is that with the exception of acomponent associat<strong>ed</strong> with the random superposition of many weak signals frombinary-star systems, the SGWB could be the result of processes that took plac<strong>ed</strong>uring the early stages of the evolution of the universe.This potential cosmological origin clearly emerges from the calculation ofthe time at which the graviton decouples from the evolution of the rest of theuniverse. For a given particle species the decoupling time t dec is defin<strong>ed</strong> as thetime when its interaction rate Ɣ equals the expansion rate of the universe, asmeasur<strong>ed</strong> by the Hubble parameter H . In fact, it can be shown [1] that, underassumptions usually met, the number of interactions that the particle species sufferfrom t dec onward is less than one. This implies that the spectrum of a particlespecies produc<strong>ed</strong> after the decoupling retains memory of the state of the universeat that time. The only change in the character of these particles is a r<strong>ed</strong>shift of themagnitude of their three-momentum due to the expansion of the universe.On purely dimensional grounds, at a given temperature T the interaction ratefor particles that interact only gravitationally is [1]:Ɣ ∼ G 2 N T 5 = T 5where G is the Newton constant and M Pl = 1/ √ G = 1.22 × 10 19 GeV is thePlanck mass (we always use units ¯h = c = k B = 1). Because in the radiationdominat<strong>ed</strong>era (before the time t eq of equal matter and radiation energy density)H ∼ T 2 /M Pl , one has:T dec ∼ M Pl .M 4 Pl181


182 Generalities on the stochastic GW backgroundHence, the gravitons are decoupl<strong>ed</strong> below the Planck scale. (At the Planckscale the above estimate of the interaction rate is not valid and nothing canbe said without a quantum theory of gravity). Just like the cosmic microwavebackground (CMB), the gravity-wave background is a randomly polariz<strong>ed</strong> relicof the Early universe. The important difference is that the electromagnetic (em)<strong>waves</strong> decoupl<strong>ed</strong> about 4 × 10 5 years after the big bang, while the gravitationalbackground could come from times as early as the Planck epoch at t Pl ≃5 × 10 −44 s. This means that the relic gravitational <strong>waves</strong> give information onthe state of the very early universe and, therefore, on physics at correspondinglyhigh energies, which cannot be access<strong>ed</strong> experimentally in any other way.Let us consider the standard Fri<strong>ed</strong>mann–Robertson–Walker (FRW)cosmological model, consisting of a radiation-dominat<strong>ed</strong> (RD) phase follow<strong>ed</strong>by the present matter-dominat<strong>ed</strong> (MD) phase, and let us call a(t) the FRW scalefactor. The RD phase goes backward in time until some new regime sets in. Thiscould be an inflationary epoch, for example, at the grand unification scale, or theRD phase could go back in time until Planckian energies are reach<strong>ed</strong> and quantumgravity sets in (t ∼ t Pl ). If the correct theory of quantum gravity is provid<strong>ed</strong> bystring theory, the characteristic mass scale is the string mass which is somewhatsmaller than the Planck mass and is presumably in the 10 17 –10 18 GeV region, andthe corresponding characteristic time is therefore one or two orders of magnitudelarger than t Pl . The transition between the RD and MD phases takes place att = t eq , when the temperature of the universe is of the order of only a few eV,so we are interest<strong>ed</strong> in graviton production which takes place well within the RDphase, or possibly at Planckian energies.A graviton produc<strong>ed</strong> with a frequency f ∗ at a time t = t 1 ∗ , within theRD phase has today (t = t 0 ) a r<strong>ed</strong>-shift<strong>ed</strong> frequency f 0 = f ∗ a ∗ /a 0 . Tocompute the ratio a ∗ /a 0 one uses the fact that during the standard RD and MDphases the universe expands adiabatically: the entropy per unit comoving volumeS = g S (T )a 3 (t)T 3 is constant, where g S (T ) counts the effective number ofspecies [1]. From this one has [2]( )gS (T 0 ) 1/3 ( )T 0100 1/3 ( ) 1 GeVf 0 = f ∗ ≃ 8.0 × 10 −14 f ∗ , (12.1)g S (T ∗ ) T ∗ g S (T ∗ ) T ∗where we us<strong>ed</strong> the fact that T 0 = 2.728 ± 0.002 K [3] and according to thestandard model g S (T 0 ) ≃ 3.91 [1].The frequency f ∗ of the graviton produc<strong>ed</strong> when the temperature was T ∗ isdetermin<strong>ed</strong> by the Hubble constant H ∗ , i.e. the size of the horizon, at the time t ∗ ofproduction. The horizon size, physically, is the length scale beyond which causalmicrophysics cannot operate (see chapter 8.4 of [1]), and therefore, for causalityreasons, we expect that the production of gravitons or any other particles, attime t ∗ , with a wavelength longer than H∗−1 , will be exponentially suppress<strong>ed</strong>.Therefore, we let λ ∗ = ɛ H∗−1,with ɛ ≤ 1. During RD, H ∗ 2 = (8π/3)Gρ rad. The1 Hereafter, the subscript α denotes the value assum<strong>ed</strong> by a generic quantity at t = t α .


Introduction 183contribution to the energy density from a single species of relativistic particlewith g i internal states (helicity, colour, etc) is g i (π 2 /30)T 4 for a boson and(7/8)g i (π 2 /30)T 4 for a fermion. Taking into account that the ith species hasin general a temperature T i ̸= T if it already dropp<strong>ed</strong> out of equilibrium, we candefine a function g(T ) from ρ rad = (π 2 /30)g(T )T 4 . Then [1]g(T ) =∑i=bosonsg i(TiT) 4+ 7 8∑i=fermions( ) 4 Tig i (12.2)Twhere the sum runs over relativistic species. This holds if a species is in thermalequilibrium at temperature T i . If instead it does not have a thermal spectrum(which in general is the case for gravitons) we can still use the above equation,where for this species T i does not represent a temperature but is defin<strong>ed</strong> (forbosons) from ρ i = g i (π 2 /30)Ti 4,where ρ i is the energy density of this species.The quantity g S (T ) us<strong>ed</strong> before for the entropy is given by the same expressionas g(T ), with (T i /T ) 4 replac<strong>ed</strong> by (T i /T ) 3 . We see that both g(T ) and g S (T )give a measure of the effective number of species. For most of the early historyof the universe, g(T ) = g S (T ), and in the standard model at T ² 300 GeV theyhave the common value g ∗ = 106.75, while today g 0 = 3.36 [1]. ThereforeH∗ 2 = 4π 3 g ∗ T∗445MPl2 , (12.3)and, using f ∗ ≡ H ∗ /ɛ, equation (12.1) can be written as [2]f 0 ≃ 1.65 × 10 −7 1 ( )T∗(g∗) 1/6Hz. (12.4)ɛ 1 GeV 100This simple equation allows us to understand a number of important pointsconcerning the energy scales that can be prob<strong>ed</strong> in GW experiments. The simplestestimate of f ∗ corresponds to taking ɛ = 1 in equation (12.4) [4]. In this case,we would find that a graviton observ<strong>ed</strong> today at the frequency f 0 = 100 Hz, thescale relevant for VIRGO, was produc<strong>ed</strong> when the universe had a temperatureT ∗ ∼ 6 × 10 8 GeV. Since in the RD phase one has [1]t = 1 ( )2H = 45 1/2M Pl16π 3 g(T ) T 2 ≃ 2.42 ( ) MeV 2g 1/2 s, (12.5)Tthis temperature corresponds to a production time t ∗ ∼ 7 × 10 −25 s. At this timethe graviton had an energy E ∗ ∼ 3 GeV.However, the fact that it makes sense to consider gravitons production onlyfor wavelengths with λ ∗ º H∗−1 does not necessarily mean, in general, that att = t ∗ the typical wavelength of GWs produc<strong>ed</strong> will be at λ ∗ ∼ H∗−1 (ɛ ∼ 1) evenas an order of magnitude estimate. In [5] this point is illustrat<strong>ed</strong> with two specificexamples, one in which the assumption λ ∗ ∼ H∗−1 turns out to be basically


184 Generalities on the stochastic GW backgroundcorrect, and one in which it can fail by several orders of magnitudes. Bothexamples will, in general, illustrate the fact that the argument does not involveonly kinematics, but also the dynamics of the production mechanism.From equation (12.4) we see that the temperatures of the early universeexplor<strong>ed</strong> detecting today relic GWs at a frequency f 0 are, for constant g ∗ , smallerby a factor approximately equal to ɛ, compar<strong>ed</strong> to those estimates with ɛ = 1.Equivalently, a signal produc<strong>ed</strong> at a given temperature T ∗ could in principle showup today in the VIRGO/LIGO frequency band when a naive estimate with ɛ = 1suggests that it falls at lower frequencies.There is, however, another effect, which instead gives hopes of exploringthe universe at much higher temperatures than naively expect<strong>ed</strong>, using GWexperiments. In fact, the characteristic frequency that we have discuss<strong>ed</strong> is thevalue of the cut-off frequency in the graviton spectrum. Above this frequency, thespectrum decreases exponentially [6], and no signal can be detect<strong>ed</strong>. Below thisfrequency, however, the form of the spectrum is not fix<strong>ed</strong> by general arguments.Thermal spectra have a low-frequency behaviour dρ/d f ∼ f 2 , but, since thegravitons below the Planck scale interact too weakly to thermalize, there is noa priori reason for a ∼ f 2 dependence. The gravitons will retain the form ofthe spectrum that they had at the time of production, and this is a very modeldependentfeature. However, as shown in [5], from a number of explicit examplesand general arguments we learn that spectra flat or almost flat over a large rangeof frequencies seem to be not at all unconceivable.This fact has potentially important consequences. It means that, even if aspectrum of gravitons produc<strong>ed</strong> during the Planck era has a cut-off at frequenciesmuch larger than the VIRGO/LIGO frequency band, we can still hope to observein the 10 Hz–1 kHz region the long low-frequency tails of these spectra.12.2 Definitions12.2.1 gw ( f ) and the optimal SNRThe intensity of a stochastic background of gravitational <strong>waves</strong> (GWs) can becharacteriz<strong>ed</strong> by the dimensionless quantity gw ( f ) = 1 ρ cdρ gwdln f , (12.6)where ρ gw is the energy density of the stochastic background of gravitational<strong>waves</strong>, f is the frequency and ρ c is the present value of the critical energy densityfor closing the universe. In terms of the present value of the Hubble constant H 0 ,the critical density is given byρ c = 3H 2 08πG . (12.7)


Definitions 185The value of H 0 is usually written as H 0 = h 0 × 100 km s −1 Mpc −1 , where h 0parametrizes the existing experimental uncertainty. A conservative estimate forthis parameter is in the range 0.50 < h 0 < 0.65 (see [5] and references quot<strong>ed</strong>therein).It is not very convenient to normalize ρ gw to a quantity, ρ c , which isuncertain: this uncertainty would appear in all the subsequent formulae, althoughit has nothing to do with the uncertainties on the GW background. Therefore, werather characterize the stochastic GW background with the quantity h 2 0 gw( f ),which is independent of h 0 . All theoretical computations of a relic GW spectrumare actually computations of dρ gw /dln f and are independent of the uncertaintyon H 0 . Therefore, the result of these computations is express<strong>ed</strong> in terms of h 2 0 gw,rather than of 2 gw . Under the assumption that the stochastic background ofgravitational radiation is isotropic, unpolariz<strong>ed</strong>, stationary and Gaussian (see [4,7]for details), it is completely specifi<strong>ed</strong> by its spectrum gw ( f ).To detect a stochastic GW background the optimal strategy consists inperforming a correlation between two (or more) detectors, since, as we willdiscuss below, the signal will be far too low to exce<strong>ed</strong> the noise level in anyexisting or plann<strong>ed</strong> single detector (with the exception of the space interferometerLISA). The strategy has been discuss<strong>ed</strong> in [8–10], and is clearly review<strong>ed</strong> in [4,7].Let us recall the main points of the analysis.The cross-correlation between the outputs s 1 (t) and s 2 (t) of two detectors isdefin<strong>ed</strong> asS =∫ T/2−T/2dt∫ T/2−T/2dt ′ s 1 (t)s 2 (t ′ )Q(t, t ′ ), (12.8)where T is the total integration time (e.g. one year) and Q is a filter function.The output of a single detector characteriz<strong>ed</strong> by an intrinsic noise n(t) and onwhich acts a gravitational strain h(t) is of the form s(t) = n(t) + h(t). Thegravitational strain can be express<strong>ed</strong> in terms of the amplitudes h +,× of the wavein the following way [11]h(t) = F + h + (t) + F × h × (t), (12.9)where the detector pattern functions F +,× are introduc<strong>ed</strong>, which depend on thelocation, orientation and geometry of the detector and the direction of arrivalof the GW and its polarization (see section 12.3). These functions have valuesin the range 0 ≤ |F +,× | ≤ 1. The noise intrinsic to the detector is assum<strong>ed</strong>stationary, Gaussian and statistically independent on the gravitational strain. As aconsequence of the assum<strong>ed</strong> stationariety of both the stochastic GW backgroundand noise, the filter function turns out to be Q(t, t ′ ) = Q(t −t ′ ). Furthermore, thenoises in the two detectors are assum<strong>ed</strong> uncorrelat<strong>ed</strong>, i.e. the ensemble average of2 This simple point has occasionally been miss<strong>ed</strong> in the literature, where one can find the statementthat, for small values of H 0 , gw is larger and therefore easier to detect. Of course, it is larger onlybecause it has been normaliz<strong>ed</strong> using a smaller quantity.


186 Generalities on the stochastic GW backgroundtheir Fourier components satisfies〈ñ ∗ i ( f )ñ j ( f ′ )〉=δ( f − f ′ 1)δ ij 2S n (i) (| f |), (12.10)where the function S n (| f |), with dimensions Hz −1 , is known as the noise powerspectrum 3 . The factor 1 2is conventionally insert<strong>ed</strong> in the definition so that the totalnoise power is obtain<strong>ed</strong> integrating S n ( f ) over the physical range 0 ≤ f < ∞,rather than from −∞ to ∞.As discuss<strong>ed</strong> in [4, 7–10], for any given form of the signal, i.e. for any givenfunctional form of h 2 0 gw( f ), it is possible to find explicitly the filter functionQ(t) which maximizes the signal-to-noise ratio (SNR). It can be shown [7] thatunder the above-mention<strong>ed</strong> assumptions, the Fourier transform of this optimalfilter and the corresponding value of the optimal SNR turn out to be, respectively:γ(| f |) gw (| f |)Q( f ) = λ| f | 3 S n (1) (| f |)S n (2) (| f |) , (12.11)[ ( )9H 4 ∫ ∞0SNR =8π 4 F 2 γ 2 (| f |) 2 ]gw (| f |) 1/4T d f0 f 6 S n (1) (| f |)S n (2) , (12.12)(| f |)where F is a normalization factor less than one depending only upon the geometryof the detectors, and λ is a real overall normalization constant that, assuming forthe spectrum a power-law gw ( f ) = β f β (with β = constant), is fix<strong>ed</strong> bythe condition 〈S〉 = β T . The function γ(f ) appearing in both formulae isthe overlap r<strong>ed</strong>uction function introduc<strong>ed</strong> in [10], which takes into account th<strong>ed</strong>ifference in location and orientation of the two detectors. At this stage let usonly remark that γ(f ) is maximum and equal to one in the case of two detectorswith the same location and orientation. The detail<strong>ed</strong> expression for F and γ(f )will be discuss<strong>ed</strong> in section 12.3. Finally, let us note that in equation (12.12) wehave taken into account the fact that what has been call<strong>ed</strong> S in equation (12.8)is quadratic in the signals and, with usual definitions, it contributes to the SNRsquar<strong>ed</strong>. This differs from the convention us<strong>ed</strong> in [4, 7].In principle the expression for the SNR, equation (12.12), is all that wene<strong>ed</strong> in order to discuss the possibility of detection of a given GW background.However, it is useful, for order of magnitude estimates and for intuitiveunderstanding, to express the SNR in terms of a characteristic amplitude of thestochastic GW-background and of a characteristic noise level, although, as wewill see, the latter is a quantity that describes the noise only approximately, incontrast to equation (12.12) which is exact. We will introduce these quantities inthe next two subsections.3 Unfortunately there is not much agreement about notations in the literature. The noise powerspectrum, that we denote by S n ( f ) following, for example, [10], is call<strong>ed</strong> P( f ) in [4]. Other authorsuse the notation S h ( f ), which we instead reserve for the power spectrum of the signal. To make thingsworse, S n is sometimes defin<strong>ed</strong> with or without the factor 1 2 in equation (12.10).


Definitions 18712.2.2 The characteristic amplitudeIn a transverse-traceless (TT) gauge, the perturbation to the Minkowski metric ofspacetime introduc<strong>ed</strong> by the GW background can be written in terms of a planewave expansion as [4]h ij (t, ⃗r) =∑ ∫ ∞ ∫ ∫ 2πd f d ˆ dψ h A ( f, ˆ, ψ)e i2π f (t− ˆ·⃗r) εA=+,×−∞ S 2 ij A ( ˆ, ψ),0(12.13)where, since h ij is real, the Fourier amplitudes are complex arbitrary functionsthat satisfy the condition h A (− f, ˆ, ψ) = h ∗ A ( f, ˆ, ψ). The unit vector ˆ isalong the propagation direction of the wave, and, in terms of the standard polar(θ) and azimuthal (φ) angles on the 2-sphere, is:ˆ ≡ (cos φ sin θ,sin φ sin θ,cos θ), (12.14)with d ˆ = d cos θ dφ. The angle ψ describes the polarization of the wave.By introducing the following pair of orthogonal unit vectors lying in the planeperpendicular to ˆˆm( ˆ) ≡ (cos φ cos θ,sin φ cos θ,− sin θ), ˆn( ˆ) ≡ (sin φ,− cos φ,0),(12.15)ψ is the angle of which is rotat<strong>ed</strong> the intrinsic frame of the wave (where, in aTT gauge, h x ′ y ′ =−h y ′ x ′) respect to the frame ( ˆm, ˆn) (see [11], p 367). Thepolarization tensors can be written aswhereε + ( ˆ, ψ) = e + ( ˆ) cos 2ψ − e × ( ˆ) sin 2ψε × ( ˆ, ψ) = e + ( ˆ) sin 2ψ + e × ( ˆ) cos 2ψ, (12.16)e + ( ˆ) = ˆm( ˆ)⊗ ˆm( ˆ)−ˆn( ˆ)⊗ˆn( ˆ),with the normalizatione × ( ˆ) = ˆm( ˆ)⊗ˆn( ˆ)+ˆn( ˆ)⊗ ˆm( ˆ)(12.17)Tr{e A ( ˆ)e A′ ( ˆ)} =2δ AA′ .In the case of a stochastic background, we treat the complex Fourieramplitude h A as a random variable with zero mean value. If this backgroundis isotropic, unpolariz<strong>ed</strong> and stationary, the ensemble average of the product oftwo Fourier amplitudes satisfies:〈h ∗ A ( f, ˆ, ψ)h A ′( f ′ , ˆ ′ ,ψ ′ )〉=δ AA ′δ( f − f ′ ) δ2 ( ˆ, ˆ ′ )4πδ(ψ − ψ ′ )2π12 S h( f ),(12.18)


188 Generalities on the stochastic GW backgroundwhere δ 2 ( ˆ, ˆ ′ ) = δ(φ−φ ′ )δ(cos θ −cos θ ′ ). The function S h ( f ) defin<strong>ed</strong> by theabove equation has dimensions Hz −1 and satisfies S h ( f ) = S h (− f ). With thisnormalization, S h ( f ) is the quantity to be compar<strong>ed</strong> with the noise level S n ( f )defin<strong>ed</strong> in equation (12.10). Using equations (12.13) and (12.18) we get〈h ij (t, ⃗r)h ij (t, ⃗r)〉 =2= 4∫ ∞−∞∫ f =∞f =0d fS h ( f ) = 4We now define the characteristic amplitude h c ( f ) from〈h ij (t, ⃗r)h ij (t, ⃗r)〉 =2∫ ∞0d fS h ( f )d(ln f ) fS h ( f ). (12.19)∫ f =∞f =0d(ln f ) h 2 c ( f ). (12.20)Note that h c ( f ) is dimensionless, and represents a characteristic value of theamplitude, per unit logarithmic interval of frequency. The origin of the factorof two on the right-hand side of equation (12.20) is carefully explain<strong>ed</strong> in [5].Comparing equations (12.19) and (12.20), we geth 2 c ( f ) = 2 fS h( f ). (12.21)We now wish to relate h c ( f ) and h 2 0 gw( f ). The starting point is the expressionfor the energy density of gravitational <strong>waves</strong>, given by the 00-component of theenergy-momentum tensor. The energy-momentum tensor of a GW cannot belocaliz<strong>ed</strong> inside a single wavelength (see, e.g., sections 20.4 and 35.7 in [12] fora careful discussion) but it can be defin<strong>ed</strong> with a spatial averaging over severalwavelengths:ρ gw = 132πG 〈ḣ ij ḣ ij 〉. (12.22)For a stochastic background, the spatial average over a few wavelengths is thesame as a time average at a given point, which, in Fourier space, is the ensembleaverage perform<strong>ed</strong> using equation (12.18). By inserting equation (12.13) intoequation (12.22) and using equation (12.18) one hasand, thusρ gw = 432πGdρ gwdln f∫ f =∞f =0d(ln f ) f (2π f ) 2 S h ( f ), (12.23)= π2G f 3 S h ( f ). (12.24)Comparing equations (12.24) and (12.21) we get the important relationdρ gwdln f= π4G f 2 h 2 c ( f ), (12.25)


Definitions 189or, dividing by the critical density ρ c , gw ( f ) = 2π 23H 2 0f 2 h 2 c ( f ). (12.26)Inserting the numerical value of H 0 ,wefind ( [11], equation (65))h c ( f ) ≃ 1.26 × 10 −18 ( 1HzfFrom equations (12.21) and (12.26) one has) √h 2 0 gw( f ). (12.27) gw ( f ) = 4π 23H 2 0f 3 S h ( f ), (12.28)and defining S n ( f ) = (S n (1) ( f )S n(2) ( f )) 1/2 , equation (12.12) can be written in thefollowing more transparent form:SNR =[2F 2 T∫ ∞0d f γ 2 ( f ) S2 h ( f )S 2 n ( f ) ] 1/4. (12.29)The factor of two in front of the integral can be understood from ∫ ∞−∞ d f =2 ∫ ∞0d f .Finally, we mention another useful formula which expresses h 2 0 gw( f ) interms of the number of gravitons per cell of the phase space, n(⃗x, ⃗k). For anisotropic stochastic background n(⃗x, ⃗k) = n f depends only on the frequency f =|⃗k|/(2π) and ρ gw = ∫ n f 2π f d 3 k/(2π) 3 = 8π 2 ∫ ∞0d(ln f )n f f 4 . Therefor<strong>ed</strong>ρ gw /dln f = 8π 2 n f f 4 , and( n ) ( )h 2 0 f f 4gw( f ) ≃ 1.810 37 . (12.30)1 kHzAs we will discuss below, to be observable at the VIRGO/LIGO interferometers,we should have h 2 0 gw ∼ 10 −6 between 1 Hz and 1 kHz, corresponding to n fof order 10 31 at 1 kHz and n f ∼ 10 43 at 1 Hz. A detectable stochastic GWbackground is therefore exce<strong>ed</strong>ingly classical, n k ≫ 1.12.2.3 The characteristic noise levelWe have seen in the previous section that there is a very natural definition ofthe characteristic amplitude of the signal, given by h c ( f ), which contains all theinformation on the physical effects, and is independent of the apparatus. Wecan, therefore, associate with h c ( f ) a corresponding noise amplitude h n ( f ), thatembodies all the informations on the apparatus, defining h c ( f )/h n ( f ) in termsof the optimal SNR.


190 Generalities on the stochastic GW backgroundIf, in the integral giving the optimal SNR, equation (12.12) orequation (12.29), we consider only a range of frequencies f such that theintegrand is approximately constant, we can writeSNR ≃[2F 2 T f γ 2 ( f )S 2 h ( f )S 2 n ( f ) ] 1/4=[F 2 T f γ 2 ( f )h 4 c ( f )2 f 2 S 2 n ( f ) ] 1/4. (12.31)The right-hand side of equation (12.31) is proportional to h c ( f ), and we cantherefore define h n ( f ) equating the right-hand side of equation (12.31) toh c ( f )/h n ( f ), so that[ ]1 fSn ( f ) 1/2h n ( f ) =( 1 2 T f . (12.32))1/4 F|γ(f )|From the derivation of equation (12.32) we can understand the limitationsimplicit in the use of h n ( f ). It gives a measure of the noise level onlyunder the approximation that leads from equation (12.29), which is exact, toequation (12.31). This means that f must be small compar<strong>ed</strong> to the scaleon which the integrand in equation (12.29) changes, so that γ(f )S h ( f )/S n ( f )must be approximately constant. In a large bandwidth this is non-trivial, andof course depends also on the form of the signal; for instance, if h 2 0 gw is flat,then S h ( f ) ∼ 1/f 3 . For accurate estimates of the SNR there is no substitutefor a numerical integration of equation (12.12) or equation (12.29), unless thefrequency range f in which we are interest<strong>ed</strong> is sufficiently small. However, fororder of magnitude estimates, equation (12.27) for h c ( f ) and equation (12.32) forh n ( f ) are simpler to use, and they have the advantage of clearly separating thephysical effect, which is describ<strong>ed</strong> by h c ( f ), from the properties of the detectors,that enter only in h n ( f ).Equation (12.32) also shows very clearly the advantage of correlating twodetectors compar<strong>ed</strong> with the use of a single detector. With a single detector, theminimum observable signal, at SNR = 1, is given by the condition S h ( f ) ≥S n ( f ). This means, from equation (12.21), a minimum detectable value for h c ( f )given byh 1dmin ( f ) = (2 fS n( f )) 1/2 , (12.33)where the superscript 1d reminds us that this quantity refers to a single detector.From equation (12.32), by indicating with ¯h 1dminthe minimum detectable signalfor a detector which noise level equals the geometric average of the noise levelsof two detectors in coincidence, the minimum detectable signal for this detectorpair is:h 2dmin ( f ) = 1 ¯h 1dmin ( f )(2T f ) 1/4 [F|γ(f )|] 1/2≃ 1.1 × 10 −2 ( 1Hzf) 1/4 ( ) 1 year 1/4 ¯h 1dTmin ( f ) . (12.34)[F|γ(f )|] 1/2


The overlap r<strong>ed</strong>uction function 191Of course, the r<strong>ed</strong>uction factor in the noise level is larger if the integration timeis larger, and if we increase the bandwidth f over which a useful coincidenceis possible. Note that h 2 0 gw is quadratic in h c ( f ), so that an improvement insensitivity by two orders of magnitudes in h c means four orders of magnitude inh 2 0 gw.12.3 The overlap r<strong>ed</strong>uction functionWhen we consider the correlation between the signals of two GW detectors wehave a r<strong>ed</strong>uction in sensitivity due to the fact that, in general, these detectorswill not be either coincident or coalign<strong>ed</strong>. This effect is quantifi<strong>ed</strong> by theoverlap r<strong>ed</strong>uction function γ(f ) appearing in the expression of the SNR (seeequation (12.12)). This is a dimensionless function of frequency f which takesinto account the relative positions and orientations of the two detectors and for anisotropic and unpolariz<strong>ed</strong> background is defin<strong>ed</strong> as [10]γ(f ) = 1 ∑〈e i2π f ˆ·⃗r F1 AF(ˆr 1, ˆ, ψ)F2 A (ˆr 2, ˆ, ψ)〉 ˆ,ψ ∼ Ɣ( f )FA(12.35)where ⃗r =⃗r 1 −⃗r 2 is the separation vector between the two detector sites, F Aiis the pattern function characterizing the response of the ith detector (i = 1, 2) tothe A =+, × polarization, and the following notation〈···〉ˆ,ψ = ∫S 2 d ˆ4π∫ 2π0dψ(···) (12.36)2πhas been introduc<strong>ed</strong> to indicate the average over the propagation direction (θ, φ)and the polarization angle ψ of the GW. The normalization factor F is given by:F = ∑ A〈F A1 (ˆr 1, ˆ, ψ)F A2 (ˆr 2, ˆ, ψ)〉 ˆ,ψ | 1≡2, (12.37)where the notation 1 ≡ 2 is a compact way to indicate that the detectors arecoincident and coalign<strong>ed</strong> and, if at least one of them is an interferometer, theangle between its arms is equal to π/2 (L-shap<strong>ed</strong> geometry). In this situation, bydefinition, γ(f ) = 1. When the detectors are shift<strong>ed</strong> apart (so there is a phaseshift between the signals in the two detectors), or rotat<strong>ed</strong> out of coalignment (sothe detectors have different sensitivity to the same polarization) it turns out that:|γ(f )| < 1.The pattern functions (or orientation factors) of a GW detector can be writtenin the following formF A (ˆr, ˆ, ψ) = Tr{D(ˆr)ε A ( ˆ, ψ)} (12.38)


192 Generalities on the stochastic GW backgroundwhere the symmetric, traceless tensor D(ˆr) describes the orientation andgeometry of the detector locat<strong>ed</strong> at ⃗r. In terms of this tensor the gravitational<strong>waves</strong>train sens<strong>ed</strong> by this detector (see equation (12.9)) is given by [10]h(t, ⃗r) = D ij (ˆr)h ij (t, ⃗r). (12.39)For an interferometer, indicating with û and ˆv the unit vectors in the directions ofits arms, one has:D(ˆr) = 1 2{û(ˆr) ⊗û(ˆr) −ˆv(ˆr) ⊗ˆv(ˆr)}. (12.40)For the lowest longitudinal mode of a cylindrical GW antenna with axis in th<strong>ed</strong>irection individu<strong>ed</strong> by the unit vector ˆl, one has [13]D(ˆr) = ˆl(ˆr) ⊗ ˆl(ˆr) − 1 3I, (12.41)where I is the unit matrix. Finally, for the lowest five degenerate quadrupolemodes (m =−2,...,+2) of a spherical detector, the corresponding tensors areD (0) (ˆr) = 12 √ 3 {e+ (ˆr) + 2g + (ˆr)} ∼ 12 √ 3 {2 f + (ˆr) − e + (ˆr)}D (+1) (ˆr) =− 1 2 g× (ˆr), D (−1) (ˆr) =− 1 2 f × (ˆr) (12.42)D (+2) (ˆr) = 1 2 e+ (ˆr), D (−2) (ˆr) =− 1 2 e× (ˆr)wheref + (ˆr) = ˆm(ˆr) ⊗ˆm(ˆr) −ˆr ⊗ˆr,g + (ˆr) =ˆn(ˆr) ⊗ˆn(ˆr) −ˆr ⊗ˆr,f × (ˆr) = ˆm(ˆr) ⊗ˆr +ˆr ⊗ˆm(ˆr)g × (ˆr) =ˆn(ˆr) ⊗ˆr +ˆr ⊗ˆn(ˆr),and e +,× (ˆr) are the tensors of equation (12.17) written in terms of the unit vectorsˆm(ˆr) and ˆn(ˆr) lying on the plane perpendicular to ˆr. From these expressions forthe tensors D ij and interpreting each of the five modes of a sphere as a singl<strong>ed</strong>etector, it is possible to show that in the case of coincident detectors one has:〈F1 A (ˆr, ˆ, ψ)F2 B (ˆr, ˆ, ψ)〉 ˆ,ψ ∼ c 12δ AB , (A, B =+, ×) (12.43)where c 12 depends only on the geometry and the relative orientations of thetwo detectors. The corresponding values of F (see equation (12.37)) for thethree different geometries consider<strong>ed</strong> (interferometer, cylindrical bar, sphere) aresummariz<strong>ed</strong> in table 12.1.By introducing the following notation⃗r = dŝ,η = 2π fd,where ŝ is the unit vector along the direction connecting the two detectors andd is the distance between them, it can be shown [10] that the overlap r<strong>ed</strong>uctionfunction assumes the following form (D k ≡ D(ˆr k )):γ(f ) = ρ 0 (η)D ij1 D 2ij + ρ 1 (η)D ij1 Dk 2i s j s k + ρ 2 (η)D ij1 Dkl 2 s is j s k s l (12.44)


The overlap r<strong>ed</strong>uction function 193Table 12.1. The normalization factor F for three different geometries of the detectors:interferometer (ITF), cylindrical bar (BAR) and sphere (SPH). A ⋆ denotes entries that canbe obtain<strong>ed</strong> from the symmetry of the table.SPHITF BAR m = 0 m =±1 m =±2ITF 2/5 ⋆ ⋆ ⋆ ⋆BAR 2/5 8/15 ⋆ ⋆ ⋆m = 0 0 2 √ 3/15 2/5 ⋆ ⋆SPH m =±1 0 0 0 2/5 ⋆m =±2 2/5 2/5 0 0 2/5where[ ρ0]ρ 1 (η) = 1[ 2η2 ][ ]−4η 2 j0ρFη 2 −4η 2 16η −20 j 1 (η), (12.45)2 η 2 −10η 35 j 2with j k (η) the standard spherical Bessel functions:j 0 (η) = sin ηη , j 1(η) = j 0(η) − cos η, j 2 (η) = 3 j 1(η)− j 0 (η).ηηIn the following, this formalism will be appli<strong>ed</strong> to treat the case in whichthe first detector is an interferometer and the other is, in turn, an interferometer, acylindrical bar and a sphere.12.3.1 Two interferometersFollowing equation (12.40), the detector tensor of the interferometer locat<strong>ed</strong> at agiven point ⃗r k can be written asD k = 1 4 {(cos 2ξ k − cos 2ζ k )e + (ˆr k ) − (sin 2ξ k − sin 2ζ k )e × (ˆr k )} (k = 1, 2)(12.46)where ξ k and ζ k are the orientations of the kth interferometer arms measur<strong>ed</strong>anticlockwise from the true north. The value of these angles, together withthe location of the central (corner) station on the Earth’s surface, for differentinterferometers are report<strong>ed</strong> in table 12.2. By inserting this expression inequation (12.44) it is possible to compute the overlap r<strong>ed</strong>uction function for ageneric pair of interferometers. The results of this calculation in the case in whichone of the interferometers is VIRGO are shown in figure 12.1.Some important conclusions can be drawn from the inspection of this figure.First of all, owing to the oscillating behaviour with η ∝ fdof the spherical Bessel


194 Generalities on the stochastic GW backgroundTable 12.2. Site and orientation of Earth-bas<strong>ed</strong> interferometric gravitational-wav<strong>ed</strong>etectors. The latitude is measur<strong>ed</strong> in degrees north from the equator, and the longitudein degrees east of Greenwich, England. In the last column is report<strong>ed</strong> the distance withrespect to VIRGO.Corner locationDistanceProject Latitude N (deg) Longitude E (deg) ξ (deg) ζ (deg) (km)VIRGO 43.63 10.50 71.5 341.5 —LIGO-LA 30.56 −90.77 108.0 198.0 7905.9LIGO-WA 46.45 −119.41 36.8 126.8 8158.2GEO-600 52.25 9.82 25.94 291.61 958.2TAMA-300 35.68 139.54 90.0 180.0 8864.1Figure 12.1. The overlap r<strong>ed</strong>uction function in the case of correlation between VIRGOand the other major interferometers.


The overlap r<strong>ed</strong>uction function 195Figure 12.2. The overlap r<strong>ed</strong>uction function for the correlation of VIRGO with a coalign<strong>ed</strong>interferometer whose central (corner) station is locat<strong>ed</strong> at: (A) (43.2 ◦ N, 10.9 ◦ E),d = 58.0 km (Italy); (B) (43.6 ◦ N, 4.5 ◦ E), d = 482.7 km (France); (C) the GEO-600(see table 12.2).function intervening in expression (12.45) of ρ k , there are always frequencies forwhich γ(f ) vanishes. Second, in all the consider<strong>ed</strong> cases |γ(f )| is well below oneon the whole frequency range, and practically zero for f > 100 Hz. In the caseof correlation with the two LIGOs and TAMA-300, |γ(f )| becomes its maximumvalue (≃0.2) before the first zero or imm<strong>ed</strong>iately after it, that is in the frequencyregion of 10–20 Hz, just where the sensitivity of the interferometric detectors islower because of the thermal noise. The case is different where the correlationis between VIRGO and GEO-600. In this case, in fact, as a consequence of therelatively small distance between the two detectors, |γ(f )|, eveniflow(≃ 0.1),remains constant up to ∼ 100 Hz. This implies that from the point of view ofthe frequency-averag<strong>ed</strong> value of |γ(f )|, this correlation is more efficient than theothers. Let us note that in terms of gw the minimum detectable signal by a pairof detectors is proportional to |γ(f )| −1 (see equations (12.26) and (12.34)) and,therefore, the low values shown in figure 12.1 lead to a substantial (even one orderof magnitude) r<strong>ed</strong>uction of the sensitivity of the correlation.As an example, figure 12.2 shows the overlap r<strong>ed</strong>uction function for thecorrelation of VIRGO with three coalign<strong>ed</strong> interferometers locat<strong>ed</strong> at about 50 km(probably the minimum distance sufficient to decorrelate local seismic and emnoises) (curve A), 480 km (curve B) and 960 km (curve C), respectively, fromthe VIRGO site. The frequency location of the first zero of γ(f ) turns out to be


196 Generalities on the stochastic GW backgroundTable 12.3. Site and orientation of the resonant bars nearest to VIRGO. In the last columnis report<strong>ed</strong> the distance with respect to VIRGO.LocationProject Latitude N (deg) Longitude E (deg) λ (deg) Distance (km)AURIGA 45.35 11.95 39.3 222.84NAUTILUS 41.80 12.67 39.3 270.12EXPLORER 46.25 6.25 39.3 443.14with good approximation inversely proportional to the distance between the twodetectors.12.3.2 Interferometer—barThe detector tensor of a resonant bar locat<strong>ed</strong> at a given point ⃗r turns out to be (seeequation (12.41))D(ˆr) = 1 2 {cos 2λe+ (ˆr) − sin 2λe × (ˆr) + b(ˆr) − 2 3I} (12.47)where the tensorb(ˆr) = ˆm(ˆr) ⊗ˆm(ˆr) +ˆn(ˆr) ⊗ˆn(ˆr)has been introduc<strong>ed</strong>, and we have indicat<strong>ed</strong> with λ the orientation of the bar axismeasur<strong>ed</strong> anticlockwise from the true North. The orientation and location of theresonant bars nearest to VIRGO are report<strong>ed</strong> in table 12.3.The behaviour of the overlap r<strong>ed</strong>uction function for the correlation of thes<strong>ed</strong>etectors with VIRGO is shown in figure 12.3, where, since the resonant detectorscan be easily rotat<strong>ed</strong>, we consider<strong>ed</strong> the case in which the bars are orient<strong>ed</strong>along the direction of the first arm of VIRGO. The frequency dependence ofthe overlap r<strong>ed</strong>uction function is meaningless for narrow-band detectors like bars(and spheres). Therefore, the results shown in figure 12.3 have been calculat<strong>ed</strong> byassuming to vary the resonance frequency by varying the length of the bar.12.3.3 Interferometer—sphereThe same calculation can be repeat<strong>ed</strong> by replacing bars with spheres of variablesize. The overlap r<strong>ed</strong>uction function for the correlation of VIRGO with each ofthe five modes of a sphere (see equation (12.42)) is shown in figure 12.4. Noticethat, since the normalization factors vanish (see table 12.1) for the m = 0, ±1modes, the quantity report<strong>ed</strong> in figure 12.4 for these modes is the Ɣ( f ) functiondefin<strong>ed</strong> in equation (12.35). It is also worthwhile to notice that the result for them = 2 case coincides exactly with that obtain<strong>ed</strong> with two interferometers, incomplete accordance with the quadrupole nature of the GW excitation.


Achievable sensitivities to the SGWB 197Figure 12.3. The overlap r<strong>ed</strong>uction function for the correlation of VIRGO withNAUTILUS (full curve), AURIGA (broken curve), and EXPLORER (dott<strong>ed</strong> curve) in thecase in which the angles λ, having the values report<strong>ed</strong> in table 12.3, coincide with the angleξ of VIRGO.12.4 Achievable sensitivities to the SGWBIn the following we apply the results of the prec<strong>ed</strong>ing sections to estimatethe expect<strong>ed</strong> sensitivities of the correlation among the various detectors to thestochastic background.12.4.1 Single detectorsTo better appreciate the importance of correlating two detectors, it is instructive toconsider first the sensitivity that can be obtain<strong>ed</strong> using only one detector. In thiscase a hypothetical signal would manifest itself as an excess noise, and shouldtherefore satisfy S h ( f ) ² S n ( f ). By imposing this condition to equation (12.28)and introducing the notation ˜h 2 f= S n ( f ), for the minimum detectable value ofh 2 0 gw one hash 2 0 min( ) ()gw ( f ) ≃ 1.3 × f 3 2˜h f10−2100 Hz 10 −22 Hz −1/2 . (12.48)This function in the case of VIRGO is plott<strong>ed</strong> in figure 12.5, where thefollowing analytical approximation for the noise power spectrum has been us<strong>ed</strong>


198 Generalities on the stochastic GW backgroundFigure 12.4. The overlap r<strong>ed</strong>uction function for the correlation of VIRGO with the fivemodes of a sphere locat<strong>ed</strong> at (43.2 ◦ N, 10.9 ◦ E) (d = 58.0 km).[14]:S n ( f ) = S 1(f0f) 5 ( ) [ ( ) ]f0f2+ S 2 + S 3 1 +ff 0(12.49)where S 1 = 3.46 × 10 −50 Hz −1 , S 2 = 6.60 × 10 −46 Hz −1 , S 3 = 3.24 ×10 −46 Hz −1 and f 0 = 500 Hz. Figure 12.5 shows also the minimum detectablevalue of h 2 0 gw for the NAUTILUS detector (target sensitivity ˜h f = 8.6 ×10 −23 Hz −1/2 at f = 907 Hz) [15], and for a hollow sphere, made of Al 5056,with a mass of 200 ton, an outer radius of 3 m, and an inner radius of 2.1 m (targetsensitivity ˜h f = 4.3 × 10 −24 Hz −1/2 at 273 Hz; ˜h f = 3.4 × 10 −24 Hz −1/2 at935 Hz) [16].Unfortunately, all these sensitivity levels are not interesting. As we shall findin section 2, an interesting sensitivity level for h 2 0 gw is at least of order 10 −7 –10 −6 . To reach such a level with a single interferometer we ne<strong>ed</strong>, for example,˜h f


Achievable sensitivities to the SGWB 199Figure 12.5. The minimum (SNR = 1) detectable value of h 2 0 gw by a single detector.at f = 1 kHz. These values for ˜h f are very far from the sensitivity offirst generation interferometers, and are in fact even well below the limitationdue to quantum noise. A very interesting sensitivity, possibly even of orderh 2 0 gw ∼ 10 −13 , could instead be reach<strong>ed</strong> with a single detector, with the plann<strong>ed</strong>space interferometer LISA [17], at f ∼ 10 −3 Hz.Note also that, while correlating two detectors the SNR improves withintegration time (see equation (12.12)) this is not so with a single detector. So,independently of the low sensitivity, with a single detector it is conceptuallyimpossible to tell whether an excess noise is due to a physical signal or is a noiseof the apparatus that has not been properly account<strong>ed</strong> for. This might not be agreat problem if the SNR is very large, but certainly with a single detector wecannot make a reliable detection at SNR of order one, so that the above estimates(which have been obtain<strong>ed</strong> setting SNR = 1) are really overestimates.12.4.2 Two detectorsThe statistical treatment of the problem of the extraction of information from thecorrelation between two detectors has been extensively discuss<strong>ed</strong> in [7] in theframe of the so-call<strong>ed</strong> ‘decision theory’ [18]. According to the results of [7],fix<strong>ed</strong> a false alarm rate α (1 − α is the fraction of experimental outcomes that th<strong>ed</strong>ecision rule correctly identifies the absence of a signal) and a minimum detectionrate δ (the fraction of experimental outcomes that the decision rule correctlyidentifies the presence of a signal with fix<strong>ed</strong> mean value > 0), the theoretical


200 Generalities on the stochastic GW backgroundTable 12.4. Minimum values of h 2 0 gw for cross-correlation measurements betweenVIRGO and the major interferometers for one year of observation, and different valuesof the pair (δ, α).Correlation (0.95, 0.05) (0.95, 0.10) (0.90, 0.10)VIRGO ⋆ LIGO-LA 5.3 × 10 −6 4.7 × 10 −6 4.2 × 10 −6VIRGO ⋆ LIGO-WA 6.4 × 10 −6 5.7 × 10 −6 5.0 × 10 −6VIRGO ⋆ GEO-600 7.2 × 10 −6 6.4 × 10 −6 5.6 × 10 −6VIRGO ⋆ TAMA-300 1.2 × 10 −4 1.1 × 10 −4 9.5 × 10 −5SNR verifies the following boundwhere the complementary error functionSNR 2 ≥ √ 2[erfc −1 (2α) − erfc −1 (2δ)]erfc(x) ∼ √ 2 ∫ ∞dz e −z2πhas been introduc<strong>ed</strong>. Therefore, in the case of a stochastic GW background havinga constant frequency spectrum gw ( f ) = gw , from equation (12.12) one has:x gw ≥ 4π 23H 2 0[ ∫1 ∞F √ d fT0γ 2 ( f )f 6 S (1)n ( f )S (2)n ( f )] −1/2[erfc −1 (2α) − erfc −1 (2δ)].(12.50)Two interferometersThe minimum values of h 2 0 gw for cross-correlation measurements betweendifferent interferometer pairs (F = 2/5 in equation (12.50)) in the case T = 10 7 s(4 months), α = 0.05, and δ = 0.95, are summariz<strong>ed</strong> in table II of [7]. Intable 12.4 we report the minimum values of h 2 0 gw for the correlation of VIRGOwith the major interferometers for one year of observation and different values ofthe detection and false alarm rates 4 .We now consider the sensitivity that could be obtain<strong>ed</strong> at VIRGO if theplann<strong>ed</strong> interferometer were correlat<strong>ed</strong> with a second identical interferometerlocat<strong>ed</strong> at a few tens of kilometres from the first, and with the same orientation(the case A at the end of section 12.3.1).A rough estimate of the sensitivity can be given using the same line ofreasoning develop<strong>ed</strong> in section 12.2.3. From equation (12.49) we see that we4 Let us note that α and δ ne<strong>ed</strong> not sum to one.


Achievable sensitivities to the SGWB 201can take, for our estimate, ˜h f ∼ 10 −22 Hz −1/2 over a bandwidth f ∼ 1 kHz.Moreover, as clearly shown in figure 12.2, the overlap r<strong>ed</strong>uction function isapproximately equal to one up to f ∼ 1 kHz. Therefore, from equation (12.31)and (12.27) one has( ) 1 year 1/2 (h 2 0 min gw ( f ) ∼ 1.3 × 10−7 SNR 2 fT 100 Hz) 3(˜h f10 −22 Hz −1/2 ) 2.(12.51)which shows that correlating two VIRGO interferometers for 1 year we can detecta relic spectrum with h 2 0 gw(100 Hz) ∼ 3 × 10 −7 at SNR = 1.65, or 10 −7at SNR = 1. Compar<strong>ed</strong> to the case of a single interferometer with SNR = 1,equation (12.48), we gain five orders of magnitude. As already discuss<strong>ed</strong>, toobtain a precise numerical value one must however consider equation (12.12).This involves an integral over all frequencies, (that replaces the somewhatarbitrary choice of f made above) and depends on the functional form ofh 2 0 gw( f ). For h 2 0 gw( f ) independent of the frequency, using the analyticalapproximation of equation (12.49) for S n(i) ( f )(i = 1, 2) and equation (12.44) forγ(f ), we get 5h 2 0 min gw( ) 1 year 1/2 ≃ 7 × 10−8 SNR 2 , (h 2 0T gw( f ) = constant). (12.52)It is interesting to note that the main contribution to the integral comes fromthe region f < 100 Hz. In fact, neglecting the contribution of the regionf > 100 Hz, the result for h 2 0 min gw changes only by approximately 4%. Also,the lower part of the accessible frequency range is not crucial. For instance,restricting the numerical integration to the regions 20 Hz ≤ f ≤ 200 Hzand 30 Hz ≤ f ≤ 100 Hz the sensitivity on h 2 0 gw degrades by 1% and10%, respectively. This means that the most important source of noise for themeasurement of a flat stochastic background is the thermal noise 6 . In particular,the sensitivity is limit<strong>ed</strong> by the mirror thermal noise, which dominates in theregion 40 Hz º f º 200 Hz, while the pendulum thermal noise dominates below5 The integral has been evaluat<strong>ed</strong> numerically in the frequency interval 2 Hz–10 kHz. The analyticalfit of equation (12.49) underestimates the noise power spectrum in the region f < 2 Hz and, in anycase, this frequency region gives no appreciable contribution to the integral. Above 10 kHz the overlapr<strong>ed</strong>uction function is negligible (see figure 12.2).6 Note also that it is not very meaningful to give more decimal figures in the minimum detectablevalue of h 2 0 min gw . Apart from the various uncertainties which enter the computation of the sensitivitycurve, a trivial source of uncertainty is the fact that the computation of the thermal noises areperform<strong>ed</strong> using a temperature of 300 K. A 5% variation, corresponding to an equally plausible valueof the temperature, gives a 5% difference in h 2 0 min gw . Quoting more figures is especially meaninglesswhen the minimum detectable h 2 0 gw is estimat<strong>ed</strong> using the approximate quantities h c ( f ), h n ( f ),i.e. approximating the integrand of equation (12.12) with a constant over a bandwidth f . Fora broadband detector these estimates typically give results which agree with the exact numericalintegration of equation (12.12) at best within a factor of two.


202 Generalities on the stochastic GW backgroundTable 12.5. Minimum values of h 2 0 gw for the correlation VIRGO–VIRGO for one yearof observation and different values of the pair (δ, α) (different values of SNR), for the threepossible locations of the second VIRGO consider<strong>ed</strong> in section 12.3.1.Correlation (0.95, 0.05) (0.95, 0.10) (0.90, 0.10)A 2.4 × 10 −7 2.1 × 10 −7 1.8 × 10 −7B 2.5 × 10 −7 2.2 × 10 −7 1.9 × 10 −7C 2.8 × 10 −7 2.5 × 10 −7 2.2 × 10 −7Table 12.6. Minimum values of h 2 0 gw(100 Hz) for the correlation VIRGO–VIRGO(T = 1 year, SNR = 1), for different values of the exponent β in equation (12.53). The rowsrefer to the three different locations of the second VIRGO consider<strong>ed</strong> in section 12.3.1.Correlation β = 0 β = 1 β =−1 β = 3A 7.2 × 10 −8 1.1 × 10 −7 3.0 × 10 −8 2.0 × 10 −8B 7.6 × 10 −8 1.3 × 10 −7 3.1 × 10 −8 9.7 × 10 −8C 8.5 × 10 −8 1.6 × 10 −7 3.2 × 10 −8 2.5 × 10 −7approximately 40 Hz. This calculation has been perform<strong>ed</strong> also for the othertwo geographical locations of the second VIRGO interferometer consider<strong>ed</strong> insection 12.3.1. Although the range of frequency where γ(f ) has a sensible valueis very different in the three cases (see figure 12.2), it is evident from table 12.5that, for fix<strong>ed</strong> values of (δ, α), the corresponding sensitivities to a frequencyindependentbackground are practically the same, and approximately one order ofmagnitude better than the ones achievable with the two LIGO detectors [7].Let us now consider the case of a frequency-dependent stochasticbackground, that in the VIRGO frequency band can be parametriz<strong>ed</strong> in thefollowing way( gw ( f ) = gw (100 Hz)f100 Hz) β. (12.53)The same proc<strong>ed</strong>ure appli<strong>ed</strong> to obtain equation (12.52) gives, in the cases β =±1, 3, the minimum detectable values for h 2 0 gw(100 Hz) report<strong>ed</strong> in table 12.6(T = 1 year, SNR = 1). For the sake of comparison, we also report<strong>ed</strong> theanalogous quantities in the case of a frequency-independent background (β = 0).Some comments about this table are necessary. First, to compare thesensitivities of a correlation to backgrounds with different frequency behaviour,we have to take into account that in the case β ̸= 0 the spectrum will exhibit apeak within the VIRGO band: at low frequency for β


Achievable sensitivities to the SGWB 203Table 12.7. Minimum values of h 2 0 gw for cross-correlation measurements between twobars and between one bar and VIRGO for one year of observation and SNR = 1.AURIGA ⋆ NAUTILUSAURIGA ⋆ VIRGONAUTILUS ⋆ VIRGO2.0 × 10 −41.6 × 10 −42.8 × 10 −4β>0. For example, in the case of the correlation A, according to table 12.6 andfrom equation (12.53), we find that a spectrum with β = 1 will be detect<strong>ed</strong>, atSNR=1,ifh 2 0 gw(1 kHz) = 1.1 × 10 −6 , while a spectrum with β =−1 willbe detect<strong>ed</strong> if h 2 0 gw(2 Hz) = 1.5 × 10 −6 . So, both for increasing or decreasingspectra, to be detectable in one year at SNR = 1, h 2 0 gw must have a peak value,within the VIRGO band, of the order of a few 10 −6 in the case β =±1. Underthe same conditions, a frequency-dependent spectrum can be detect<strong>ed</strong> at the level7 × 10 −8 (see the first column of table 12.6). Clearly, for detecting increasing(decreasing) spectra, the upper (lower) part of the frequency band becomes moreimportant, and this is the reason why the sensitivity degrades compar<strong>ed</strong> to flatspectra, since for increasing or decreasing spectra the maximum of the signal is atthe <strong>ed</strong>ges of the accessible frequency band, where the interferometer sensitivityis worse. Second, for each β the sensitivity on h 2 0 gw( f ) always degrades goingfrom A to C. Relatively to A, the size of this decay for B and C increases with β,and for β = 3 the correlation C show a sensitivity one order of magnitude lowerthan the sensitivity of A. This means that in order to have the same sensitivity, thein case C has to be two orders of magnitude lower than in case A(see equation (12.50)). The case β = 3 is particularly important, because this isthe frequency dependence of the spectrum pr<strong>ed</strong>ict<strong>ed</strong> by the string cosmology.product S n (1) S n(2)Resonant masses and resonant mass-interferometerResonant mass detectors includes bars like NAUTILUS, EXPLORER andAURIGA (see, e.g., [15, 21] for reviews). Spherical [16, 22] and truncat<strong>ed</strong>icosah<strong>ed</strong>ron (TIGA) [23] resonant masses are also being develop<strong>ed</strong> or studi<strong>ed</strong>.The correlation between two resonant bars and between a bar and aninterferometer has been extensively consider<strong>ed</strong> in [19, 24–26]. The resultsobtain<strong>ed</strong> in [19] for the minimum values of h 2 0 gw in the case of a backgroundhaving constant frequency spectrum, for one year of integration and SNR = 1, aresummariz<strong>ed</strong> in table 12.7.Using resonant optical techniques, it is possible to improve the sensitivityof interferometers at special values of the frequency, at the expense of theirbroadband sensitivity. Since bars have a narrow band anyway, narrow-bandingthe interferometer improves the sensitivity of a bar-interferometer correlation by


204 Generalities on the stochastic GW backgroundabout one order of magnitude [20].While resonant bars have been taking data for years, spherical detectors areat the moment still at the stage of theoretical studies (although prototypes mightbe built in the near future), but could reach extremely interesting sensitivities. Inparticular, the correlation between VIRGO and one sphere with a diameter of 3 m,made of Al 5056 (M = 38 ton), gives sensitivities that are, respectively, h 2 0 gw ∼2 × 10 −5 , if the sphere is locat<strong>ed</strong> in the AURIGA site, and h 2 0 gw ∼ 4 × 10 −5 ifthe sphere is at the NAUTILUS site [19]. Instead, the correlation of two spheres ofthis type but locat<strong>ed</strong> at the same site could reach a sensitivity h 2 0 gw ∼ 4 × 10 −7[19]. All these figures improve using a denser material or increasing the spher<strong>ed</strong>iameter, but it might be difficult to build a heavier sphere. Another verypromising possibility is given by hollow spheres [16]. The theoretical studiesof [16] suggest for the correlation of two 40-ton colocat<strong>ed</strong> hollow spheres, madeof Al 5056, a sensitivity of h 2 0 gw ∼ 6 × 10 −8 at f = 218 Hz (one year ofobservation and SNR = 1).12.4.3 More than two detectorsWhen the outputs of M > 2 detectors are available, the information aboutthe magnitude of the stochastic GW background can be extract<strong>ed</strong> in two ways:combining the measurements from each detector pair or directly correlating theoutputs of the detectors. Both these techniques have been extensively treat<strong>ed</strong>in [7], and here we shall only review the key results obtain<strong>ed</strong> in this analysis.Multiple detector pairsWe indicate withS (ij)1, S (ij)2,...,S n (ij)ij, (i, j = 1,...,M)the n ij different measurements, each of length T , of the optimally-filter<strong>ed</strong> crosscorrelationsignal S (ij) between the ith and jth detectors (see equation (12.8)).Under the following hypothesis:• T ≫ the light travel time between any pair of detectors;• the optimal filter functions (see equation (12.8)) for each detector pair arenormaliz<strong>ed</strong> in a way that〈S (ij) 〉= 1n ij∑n ijS (ij)k= β T, ∀(i, j)k=1for a stochastic background having a power-law spectrum gw ( f ) = β f β ;• the noise is large, i.e. the covariance matrix built from the cross-correlationsignals taken during the same time interval T is approximately diagonalC (ij)(kl) =〈S (ij) S (kl) 〉−〈S (ij) 〉〈S (kl) 〉≃δ ik δ kl (σ (ij) ) 2


Achievable sensitivities to the SGWB 205Table 12.8. Minimum values of h 2 0 gw for one year of observation, for(δ, α) = (0.95, 0.05), for the optimal combination of cross-correlation measurementsbetween multiple detector pairs, taken from all possible triples of interferometerscontaining VIRGO.VIRGO ⋆ LIGO-WA ⋆ LIGO-LAVIRGO ⋆ LIGO-LA ⋆ GEO-600VIRGO ⋆ LIGO-WA ⋆ GEO-600VIRGO ⋆ LIGO-LA ⋆ TAMA-300VIRGO ⋆ LIGO-WA ⋆ TAMA-300VIRGO ⋆ GEO-600 ⋆ TAMA-3002.6 × 10 −64.2 × 10 −64.8 × 10 −65.4 × 10 −66.6 × 10 −67.3 × 10 −6Table 12.9. The same as table 12.8 for the case of quadruples.VIRGO ⋆ LIGO-WA ⋆ LIGO-LA ⋆ GEO-600VIRGO ⋆ LIGO-WA ⋆ LIGO-LA ⋆ TAMA-300VIRGO ⋆ LIGO-LA ⋆ GEO-600 ⋆ TAMA-300VIRGO ⋆ LIGO-WA ⋆ GEO-600 ⋆ TAMA-3002.4 × 10 −62.6 × 10 −64.2 × 10 −64.8 × 10 −6and, therefore, all the measurements can be consider<strong>ed</strong> uncorrelat<strong>ed</strong>;the authors of [7] show that the optimal SNR turns out to be 7SNR 4 =M∑i=1M∑n ij (SNR (ij) ) 4 (12.54)j>iwhere SNR (ij) is given in equation (12.12). Thus, in the case of a frequencyindependentspectrum gw ( f ) = gw , the minimum value of gw for givendetection (δ) and false alarm (α) rates turns out to be:( )1 2= gw (δ, α)M∑i=1( ) 2 M∑ 1 (ij) . (12.55)gw (δ, α)where (ij)gw (δ, α) are the analogous quantities for the ij detector pairs (seeequation (12.50)). In the case (δ, α) = (0.95, 0.05), the minimum values ofh 2 0 gw obtain<strong>ed</strong> considering all triples and quadruples of interferometers havingin common VIRGO, are report<strong>ed</strong> in tables 12.8 and 12.9, respectively.7 Remember that we defin<strong>ed</strong> the SNR as the square root of the SNR defin<strong>ed</strong> in [7].j>i


206 Generalities on the stochastic GW backgroundMany-detector correlationThe M-detector correlation signal S is the obvious generalization ofequation (12.8) for a two-detector:∫ T/2∫ T/2S = dt 1 ... dt M s 1 (t 1 )s 2 (t 2 )...s M (t M )Q(t 1 ,...,t M ).−T/2 −T/2Let us remark that, as a consequence of the fact that each detector output signalis assum<strong>ed</strong> to be a random Gaussian variable having zero mean value, 〈S〉 isdifferent from zero only for an even number of detectors (M = 2N).Under the same hypothesis assum<strong>ed</strong> in section 12.2.1 for the treament of thetwo-detector case, the authors of [7] show that the SNR for the optimally-filter<strong>ed</strong>2N-detector correlation is given bySNR 4 ≈ ∑ {...}(SNR (12) SNR (34) ...SNR (2N,2N−1) ) 4 (12.56)where the sum run over all the possible permutations of the sequence{(ij), (kl),...,(pq)} with (i < j, k < l,...,p < q). In the case of afrequency-independent background, with fix<strong>ed</strong> detection and false alarm rates,the minimum detectable value of gw from data obtain<strong>ed</strong> via a 2N-detectorcorrelation experiment, is given by(1) 2 gw (δ, α)[= C(δ, α) ∑ ()12 ] 1/N{...} gw (12) (δ, α) gw (34) (δ,α)... (2N−1,2N)gw (δ, α)(12.57)withC(δ, α) ={ √ 2[erfc −1 (2α) − erfc −1 (2δ)]} 2(N−1)/N .In table 12.10 are report<strong>ed</strong> the minimum values of h 2 0 gw obtain<strong>ed</strong> fromequation (12.57) for the four-interferometer correlations. It is important to notethat, as clearly shown from the comparison with tables 12.8 and 12.9, theseminimum values are always greater than those for optimal combination of datafrom multiple detector pairs.By summarizing the results of this section we can say that the improvementin sensitivity obtain<strong>ed</strong> using more detectors is not large compar<strong>ed</strong> to the case ofa single pair correlation (see table 12.4). This is due to the fact that correlating2, 4,...,2N detectors or combining in a optimal way data from multiple detectorpairs, one does not change the general dependence that gw ∼ T −1/2tot


Observational bounds 207Table 12.10. Minimum values of h 2 0 gw for one year of observation, for(δ, α) = (0.95, 0.05), for optimally-filter<strong>ed</strong> four-detector correlations.VIRGO ⋆ LIGO-WA ⋆ LIGO-LA ⋆ GEO-600VIRGO ⋆ LIGO-WA ⋆ LIGO-LA ⋆ TAMA-300VIRGO ⋆ LIGO-LA ⋆ GEO-600 ⋆ TAMA-300VIRGO ⋆ LIGO-WA ⋆ GEO-600 ⋆ TAMA-3003.7 × 10 −61.4 × 10 −52.2 × 10 −52.7 × 10 −5of the minimum detectable value on the total observation time: what changes fromone case to another is only the numerical factors multiplying T −1/2tot . Although theimprovement in sensitivity is limit<strong>ed</strong> the availability of the signals from variousdetectors would be important in ruling out spurious effects.12.5 Observational boundsWe close this section discussing what is actually known from the observationalside about the stochastic background. At present, there are strict limits on thisbackground in only a couple of frequency ranges, but other than that, only onevery general constraint. In the following we will only discuss the general one, thenucleosynthesis bound, because it is the only relevant one for the frequency region1Hz< f < 1 kHz cover<strong>ed</strong> by the ground bas<strong>ed</strong> detectors. The other bounds areinferr<strong>ed</strong> from the timing irregularities in the arrival times of the pulses emitt<strong>ed</strong> bysome millisecond pulsars and the anisotropies on large angular scales of the CMB.They, respectively, constrain gw in the frequency regions f ∼ 10 −8 Hz and10 −18 Hz º f º 10 −16 Hz, which are far below any frequency band accessiblefor the present-day ground- or space-bas<strong>ed</strong> experiments. A complete discussionof these bounds can be found in [4].Nucleosynthesis successfully pr<strong>ed</strong>icts the primordial abundances ofdeuterium, 3 He, 4 He and 7 Li in terms of one cosmological parameter η, thebaryon to photon ratio. In the pr<strong>ed</strong>iction parameters of the underlying particletheory also enter, which are therefore constrain<strong>ed</strong> in order not to spoil theagreement. In particular, the pr<strong>ed</strong>iction is sensitive to the effective number ofspecies at the time of nucleosynthesis, g ∗ = g(T ≃ 1 MeV). With somesimplifications, the dependence on g ∗ can be understood as follows. A crucialparameter in the computations of nucleosynthesis is the ratio of the numberdensity of neutrons, n n , to the number density of protons, n p . As long as thermalequilibrium is maintain<strong>ed</strong> we have (for non-relativistic nucleons, as appropriateat T ∼ 1 MeV) n n /n p = exp(−Q/T ) where Q = m n − m p ≃ 1.3 MeV.Equilibrium is maintain<strong>ed</strong> by the process pe ↔ nν, with width Ɣ pe→nν , as longas this width is greater than H . When the rate drops below the Hubble constantH , the process cannot compete anymore with the expansion of the universe and,


208 Generalities on the stochastic GW backgroundapart from occasional weak processes, is dominat<strong>ed</strong> by the decay of free neutrons,the ratio n n /n p remains frozen at the value exp(−Q/T f ), where T f is the value ofthe temperature at the time of freeze-out. This number, therefore, determinesthe density of neutrons available for nucleosynthesis, and since practically allneutrons available will eventually form 4 He, the final primordial abundance ofthis nucleus is very sensitive to the freeze-out temperature T f . If we assumefor simplicity Ɣ pe→nν ≃ G 2 F T 5 (which is really appropriate only in the limitT ≫ Q), where G F is the Fermi constant, from equation (12.3) turns out that T fis determin<strong>ed</strong> by the conditionG 2 F T 5f≃(4π 3 g ∗45) 1/2T2fM Pl. (12.58)This shows that T f ∼ g∗1/6 , at least with the approximation that we us<strong>ed</strong> forƔ pe→nν . A large energy density in relic gravitons gives a large contribution to thetotal density ρ and therefore to g ∗ . This results in a larger freeze-out temperature,more available neutrons and then in overproduction of 4 He. This is the idea behindthe nucleosynthesis bound [28]. More precisely, since the density of 4 He increasesalso with the baryon to photon ratio η, we could compensate an increase in g ∗ witha decrease in η, and therefore we also ne<strong>ed</strong> a lower limit on η, which is provid<strong>ed</strong>by the comparison with the abundance of deuterium and 3 He.Rather than g ∗ , N ν is often us<strong>ed</strong> as an ‘effective number of neutrino species’and is defin<strong>ed</strong> as follows. In the standard model, at T ∼ a few MeV, the activ<strong>ed</strong>egrees of fre<strong>ed</strong>om are the photon, e ± , neutrinos and antineutrinos, and they havethe same temperature, T i = T . Then, for N ν families of light neutrinos, one hasg ∗ (N ν ) = 2 + 7 8 (4 + 2N ν), (12.59)where the factor of two comes from the two elicity states of the photon, fourfrom e ± in the two elicity states, and 2N ν counts the N ν neutrinos and the N νantineutrinos, each with their single elicity state. According to the standardmodel, N ν = 3 and therefore g ∗ = 43/4. Therefore, we can define an ‘effectivenumber of neutrino species’ N ν fromg ∗ (N ν ) ∼ 434 + ∑i=extra bosonsg i(TiT) 4+ 7 8∑i=extra fermions( ) 4 Tig i . (12.60)TOne extra species of light neutrino, at the same temperature as the photons, wouldcontribute one unit to N ν , but all species, weight<strong>ed</strong> with their energy density,contribute to N ν , which, of course, in general is not an integer. For i = gravitons,we have g i = 2 and (T i /T ) 4 = ρ gw /ρ γ , where ρ γ = 2(π 2 /30)T 4 is the photonenergy density. If gravitational <strong>waves</strong> give the only extra contribution to N ν ,compar<strong>ed</strong> to the standard model with N ν = 3, using equation (12.59), the above


equation gives imm<strong>ed</strong>iately(ρgwρ γ)NSObservational bounds 209= 7 8 (N ν − 3), (12.61)where the subscript NS reminds us that this equality holds at the time ofnucleosynthesis. If more extra species, not includ<strong>ed</strong> in the standard model,contribute to g ∗ (N ν ), then the equals sign in the above equation is replac<strong>ed</strong> byless than or equal. The same happens if there is a contribution from any otherform of energy present at the time of nucleosynthesis and not includ<strong>ed</strong> in theenergy density of radiation, like, for example, primordial black holes.To obtain a bound on the energy density at the present time, we notethat from the time of nucleosynthesis to the present time ρ gw scal<strong>ed</strong> as 1/a 4 ,while, as a consequence of the assum<strong>ed</strong> adiabatic expansion of the universe (seesection 12.1), ρ γ ∼ T 4 ∼ 1/(a 4 g 4/3 ). Therefore, one has(ρgwS( ( )ρgw gS (T 0 ) 4/3 ( ( )ρgw 3.914/3==. (12.62)ρ γ)0ρ γ)NSg S (1 MeV) ρ γ)NS10.75Therefore we get the nucleosynthesis bound at the present time,(ρgwρ γ)0≤ 0.227(N ν − 3). (12.63)Of course this bound holds only for GWs that were already produc<strong>ed</strong> at thetime of nucleosynthesis (T ∼ 1 MeV, t ∼ 1 s). It does not apply toany stochastic background produc<strong>ed</strong> later, like backgrounds of astrophysicalorigin (see section 13.5). Note that this is a bound on the total energydensity∫in gravitational <strong>waves</strong>, integrat<strong>ed</strong> over all frequencies. Writing ρ gw =d(ln f ) dρgw /dln f , multiplying both ρ gw and ρ γ in equation (12.63) by h 2 0 /ρ c,and inserting the numerical value h 2 0 ρ γ /ρ c ≃ 2.474 × 10 −5 [29], we get∫ f =∞f =0d(ln f ) h 2 0 gw( f ) ≤ 5.6 × 10 −6 (N ν − 3). (12.64)The bound on N ν from nucleosynthesis is subject to various systematic errorsin the analysis, which have to do mainly with the issues of how much of theobserv<strong>ed</strong> 4 He abundance is of primordial origin, and of the nuclear processing of3 He in stars, and as a consequence over the last five years have been quot<strong>ed</strong> limitson N ν ranging from 3.04 to around 5. The situation has been recently review<strong>ed</strong>in [30]. The conclusion of [30] is that, until current astrophysical uncertaintiesare clarifi<strong>ed</strong>, N ν < 4 is a conservative limit. Using extreme assumptions, ameaningful limit N ν < 5 still exists, showing the robustness of the argument.Correspondingly, the right-hand side of equation (12.64) is, conservatively, oforder 5 × 10 −6 and anyway cannot exce<strong>ed</strong> 10 −5 .


210 Generalities on the stochastic GW backgroundIf the integral cannot exce<strong>ed</strong> these values, also its positive definite integrandh 2 0 gw( f ) cannot exce<strong>ed</strong> it over an appreciable interval of frequencies, ln f ∼1. One might still have, in principle, a very narrow peak in h 2 0 gw( f ) at somefrequency f , with a peak value larger than say 10 −5 , while still its contributionto the integral could be small enough. However, apart from the fact thatsuch behaviour seems rather implausible, or at least is not suggest<strong>ed</strong> by anycosmological mechanism, it would also be probably of little help in the detectionat broadband detectors like VIRGO, because even if we gain in the height of thesignal we lose because of the r<strong>ed</strong>uction of the useful frequency band f , seeequation (12.12).These numbers, therefore, give a first idea of what can be consider<strong>ed</strong> aninteresting detection level for h 2 0 gw( f ), which should be at least a few times10 −6 , especially considering that the bound (12.64) refers not only to gravitational<strong>waves</strong>, but to all possible sources of energy which have not been includ<strong>ed</strong>, likeparticles beyond the standard model, primordial black holes, etc.As a final remark, we note that a very weak bound in the region f ∼1 kHz has been obtain<strong>ed</strong> from the analysis of correlation between bar detectors.Preliminary results on a NAUTILUS–EXPLORER correlation, using 12 hours ofdata, have been report<strong>ed</strong> in [27], and give h 2 0 gw( f = 920 Hz) ∼ 120.


Chapter 13Sources of SGWBHere we review the present knowl<strong>ed</strong>ge about the potential processes, both ofcosmological and astrophysical origin, from which a stochastic GW backgroundmight arise. We examine in some detail the mechanisms at work in each case andthe features of the corresponding spectrum of GW radiation.13.1 Topological defectsThe concept of the spontaneous symmetry breaking, the idea that there areunderlying symmetries of Nature that are not manifest in the structure of thevacuum, play a crucial role in the modern description of the particle interactions.Of particular interest for cosmology is the theoretical expectation that at hightemperatures, symmetries that are spontaneously broken today were restor<strong>ed</strong>[31]. In the context of the hot big bang cosmology this implies a sequence ofphase transitions in the early universe, with critical temperatures relat<strong>ed</strong> to thecorresponding symmetry breaking scales.To illustrate the phenomenon of the high-temperature symmetry restorationwe consider a complex scalar field with a ‘Mexican-hat’ potentialV (φ) = 1 2 λ(φ† φ − η 2 ) 2 (λ > 0). (13.1)The Lagrangian is invariant under the group U(1) of the global phasetransformations, φ → e iα φ. The minima of the potential are at nonzero valuesφ, and so the symmetry is spontaneously broken and φ acquires a vacuumexpectation value (VEV)〈φ〉 =ηe iθ (13.2)where the phase θ is arbitrary. We thus have a manifold Å of degenerate vacuumstates corresponding to different choices of θ, that, in this case, is the circle|φ| =η in the complex φ plane.211


212 Sources of SGWBV (Τ φ)High TLow T−η+ηφFigure 13.1. The effective potential.At finite temperature the effective potential for φ acquires additional,temperature-dependent terms. In the high-temperature limit, one has:V T (φ) = AT 2 φ † φ + V (φ) (13.3)where the dimensionless constant A depend on the self-coupling λ and is assum<strong>ed</strong>to be positive. From equations (13.1) and (13.3) we see that the effective mass ofthe field φ at temperature T is:( ) λ 1/2m 2 (T ) = 2A(T 2 − Tc 2 ) where T c = η.AFrom this expression we see that for T > T c , m 2 (T ) is positive, the minimumof V T (φ) is at φ = 0 and so the symmetry is restor<strong>ed</strong> (see figure 13.1).The temperature T c is the critical temperature of the phase transition from thesymmetric (〈φ〉 =0) to the broken-symmetry (〈φ〉 ̸=0) phase. Unless λ is verysmall, T c ∼ η. In this example the transition is second order: the symmetric phasecorresponds to a maximum for V T at T < T c and the transition occurs smoothly.More complicat<strong>ed</strong> models can lead to first-order transitions, where the symmetricphase remains a local minimum at T < T c separat<strong>ed</strong> by a barrier from the minimaat φ ̸= 0. In this case the transition occurs through the bubble nucleation.In a cosmological context, as the universe cools through the criticaltemperature the Higgs field φ develop an expectation value 〈φ〉 correspondingto some point in the manifold Å. Since all points of this manifold are equivalent,the choice depends on random fluctuations and is different in different regionsof space. Therefore, associat<strong>ed</strong> to the phase transition there is a length scale ξrepresenting the maximum distance over which the Higgs field can be correlat<strong>ed</strong>.


Topological defects 213This correlation length depends upon the details of the phase transition and istemperature dependent. In any case, since correlations cannot be establish<strong>ed</strong> atspe<strong>ed</strong> greater than the spe<strong>ed</strong> of light, ξ cannot exce<strong>ed</strong> the causal horizon d H , th<strong>ed</strong>istance travell<strong>ed</strong> by light during the lifetime of the universe. In the standardcosmology d H ∼ t and, thus, one hasξ ≤ t cwhere t c is the time at which the phase transition is complet<strong>ed</strong>. The actualmagnitude of ξ at the phase transition and afterwards is determin<strong>ed</strong> bycomplicat<strong>ed</strong> dynamical processes and can be much smaller than this causalityupper bound.Since the field must be continous, on the boundaries between differentcorrelation regions φ leaves the vacuum manifold Å and assume valuescorresponding to a high potential energy. For topological reasons, these regionsof false vacuum are stable and survive to further evolution of the universe frozenin the form of topological defects.The cosmological production mechanism describ<strong>ed</strong> above is known as theKibble mechanism [32] and is very much akin to the mechanism for productionof various defects in solid state and condens<strong>ed</strong> matter systems. Crystal defects, forexample, form when the water freezes or when a metal crystallizes. The analogiesbetween defects in particle physics and condens<strong>ed</strong> matter physics are quite deep.Defects form for the same reason: the vacuum manifold is topologically nontrivial.However, the defect dynamics is different. The motion of defects incondens<strong>ed</strong> matter are friction-dominat<strong>ed</strong>, whereas the defect in cosmology obeyrelativistic equations, second order in time derivatives, since they come from arelativistic field theory.Depending on the topology of the manifold Å the defects can occur inthe form of points, lines or surfaces. They are call<strong>ed</strong> monopoles, strings anddomain walls, respectively [32]. Hybrid defects can be form<strong>ed</strong> in a sequence ofphase transitions, for example, the first transition produces monopoles, which getconnect<strong>ed</strong> by strings at the second phase transition. The main conclusions of thestudies about these defects can be summariz<strong>ed</strong> as follows.• Domain walls and monopoles are disastrous for cosmological models andtheir presence should be avoid<strong>ed</strong>.• Strings cause no harm, but can lead to very interesting cosmologicalconsequences. In particular, they can generate density fluctuations sufficientto explain galaxy formation and can produce a number of distinctive andunique observational effects (anisotropies in the CMB temperature, doubleimages of objects behind them and a stochastic GW background).• Hybrid defects are transient and eventually decay into relativistic particles.If this happens at a sufficiently early time, the decay products thermalize andwe can see no trace of the defects, except perhaps in the form of gravitational<strong>waves</strong>.


214 Sources of SGWB13.1.1 StringsTo illustrate the (cosmic) strings let us come back to the simple model consider<strong>ed</strong>at the beginning of the prec<strong>ed</strong>ing section (see equations (13.1) and (13.2)). Since〈φ〉 is single valu<strong>ed</strong>, the total change of the phase θ around any clos<strong>ed</strong> path inspace must be an integer multiple of 2π. Let us now consider a clos<strong>ed</strong> path withθ = 2π. If no singularity is encounter<strong>ed</strong>, as the path shrunk to a point, θcannot change continuously from 2π to zero, and, thus, we must encounter atleast one point where θ is undefin<strong>ed</strong>, i.e. 〈φ〉 =0. This means that at least onetube of false vacuum should be caught inside any path with θ ̸= 0. Such tubesof false vacuum, call<strong>ed</strong> strings, must either be close or infinite in length, otherwiseit would be possible to contract the path to a point without crossing the string.The simplest strings are produc<strong>ed</strong> in the phase transition associat<strong>ed</strong> to thespontaneous breaking of a local U(1) symmetry. In this case the Lagrangiancontains a gauge field A µ and a complex Higgs field φ which carries U(1) chargeg and with self-interaction of the form (13.1):Ä = (D µ φ) † (D µ φ) − 1 4 F µν F µν − V (φ) (13.4)whereD µ = ∂ µ − igA µ , F µν = ∂ µ A ν − ∂ ν A µ .In this case the string solution has a well-defin<strong>ed</strong> core outside of which φ containsno energy density in spite of non-vanishing gradients ∇φ: the gauge field A µ canabsorb the gradient, i.e. D µ φ = 0 when ∂ µ φ ̸= 0. The radius δ of the string coreis determin<strong>ed</strong> by the Compton wavelengths of the Higgs and vector bosons. Form φ ≪ m A , which is usually the case, one has:δ ∼ m −1φ= λ−1/2 η −1and the energy of the string per unit length within this width is finite and givenby:µ ∼ λη 4 δ 2 = η 2(independent of the coupling λ). The value of µ (or equivalently η) is the onlyfree parameter of the string. For a phase transition at the grand unification (GUT)energy scale, η = 10 16 GeV andδ ∼ 10 −30 cm, µ ∼ 10 22 gcm −1 . (13.5)Strings of cosmological interest have sizes much greater than their width. Inthis case the internal structure of the string is unimportant and physical quantitiesof interest, such as the energy-momentum tensor, can be averag<strong>ed</strong> over the crosssection. For a long, thin, straight string lying along the z-axis we define˜T ν µ = δ(x)δ(y) ∫dx dyT ν µ .


Topological defects 215It can be shown [33] that, as a consequence of the invariance of the string underLorentz boost along z and the conservation law Tµ,ν ν one has:˜T ν µ= µδ(x)δ(y) diag(1, 0, 0, 1). (13.6)This expression shows a remarkable property of the strings: the pressure isnegative, i.e. is a string tension, and this tension is equal to the mass per unitlength µ. One recalls from classical mechanics that small transverse <strong>waves</strong> in astring with tension T move at spe<strong>ed</strong> (T/µ) 1/2 , so it is apparent that <strong>waves</strong> movealong a string at the velocity of light.Appli<strong>ed</strong> to strings, the Kibble mechanism implies that at the time of phasetransition a network of strings with typical length ξ(t c ) will form. According tonumerical simulations at formation about one fifth of the initial energy is in smallclos<strong>ed</strong> loops and the remaining in ‘infinite’ long strings. The evolution of thisnetwork for t > t c is complicat<strong>ed</strong>. The key processes are:(i)The intercommutation of intersecting string segments, in which the twosegments swap partners, rather than passing through one another (see figurebelow). Bas<strong>ed</strong> upon numerical simulations it appears that the probability forthis to occur is nearly unity. This process leads to the continual chopping oflong strings into smaller segments and/or loops.(ii) The decay of small loops through the emission of gravitational radiation.The strings oscillate relativistically under their own tension and, thus, a loopof characteristic radius R will radiate gravitational <strong>waves</strong> at a characteristicfrequency ω ∼ R −1 due to its time-varying quadrupole moment, Q ∼ µR 3 .For an order of magnitude estimate the power radiat<strong>ed</strong> in gravitational <strong>waves</strong>can be calculat<strong>ed</strong> using the quadrupole formula (G is Newton’s constant)P ∼ G〈 ˙¨Q ˙¨Q〉 ∼Gµ 2 . (13.7)Because this power is constant, independent of the loop size, the mass-energyof the loop decreases linearly with time. In a characteristic timeτ ∼RGµthe loop shrinks to a point and vanishes. From equation (13.5), in the case ofa GUT phase transition, one has Gµ ∼ 10 −6 .


216 Sources of SGWBThese two processes combine to create a mechanism by which the infinitestring network loses energy (and length as measur<strong>ed</strong> in comoving coordinates),preventing the network from dominating the universe. Inde<strong>ed</strong>, bas<strong>ed</strong> uponcomputer simulations and analytical arguments, there is strong evidence that thecosmological evolution of the network becomes self-similar, approaching what iscall<strong>ed</strong> a ‘scaling’ solution. In the simplest scale-invariant model, the correlationlength ξ of the network is proportional to its causality bound:ξ(t) ∼ t ∼ d H (t)and, thus, the statistical properties of the network are time independent if all th<strong>ed</strong>istances are scal<strong>ed</strong> to the causal horizon.This scaling property can be us<strong>ed</strong> to obtain qualitative relations. Forexample, the energy density of long strings is given by:ρ ∞ = A µdH 2 (t) ∼ A µ t 2where A is a dimensionless constant representing the number of long stringspresent per horizon siz<strong>ed</strong> volume. Numerical simulations suggest the valueA = 52 in the radiation-dominat<strong>ed</strong> era, and A = 31 in the matter-dominat<strong>ed</strong>one. True scale invariance implies that the size of a newly form<strong>ed</strong> loop produc<strong>ed</strong>by the network is a fix<strong>ed</strong> fraction of the horizonl(t) = αd H (t) ∼ αt.Although the loops are observ<strong>ed</strong> to form with relativistic peculiar velocities v i (theloop centre of mass is moving with respect to the rest frame of the cosmologicalfluid), these are rapidly r<strong>ed</strong>shift<strong>ed</strong> to zero by the expansion of the universe, leavinga generic loop with only a fraction f r = (1 − v 2 i )1/2 of its initial energy. Thisr<strong>ed</strong>shifting of peculiar velocities does not affect the loop production rate, but itdoes change the loop size imm<strong>ed</strong>iately after its formation to:l(t) = f r αd H (t). (13.8)Hence, only a fraction f r of the total loop energy is convert<strong>ed</strong> into GWs. Bynumerical simulations this fraction turns out to be 0.71.This scale-invariant model is implement<strong>ed</strong> by the assumption that theuniverse is describ<strong>ed</strong> by a spatially flat ( = 1) FRW cosmology. A full treatmentof this model is develop<strong>ed</strong> in [34]. In particular, the effect of the string networkon the expansion of the universe and the rate of loop formation are calculat<strong>ed</strong>.The spectrum of GWs produc<strong>ed</strong> by a network of string loops can beobtain<strong>ed</strong> implementing the scale-invariant model describ<strong>ed</strong> above with a modelof the emission of gravitational radiation by string loops. This model has beendevelop<strong>ed</strong> in [34] and is compos<strong>ed</strong> of the following three elements:


Topological defects 217(i)A loop radiates with powerP = ƔGµ 2 (13.9)where Ɣ is a dimensionless radiation efficiency that does not depend on theloop size, but only on its shape. Recent studies of realistic loop configurationindicate that the distribution of the values of this parameter has mean value〈Ɣ〉 ≈60. From equations (13.8) and (13.9), it follows that a loop form<strong>ed</strong> attime t b at a time t > t b has lengthl(t, t b ) = f r αd H (t) − ƔGµ(t − t b ).The loop disappears when this length reaches zero, at a time(t d = 1 + f )r αd H (t b )t b = βt bƔGµt b(β is not function of t b because d H (t b ) ∼ t b ).(ii) The frequency of GWs emitt<strong>ed</strong> at time t by a loop form<strong>ed</strong> at time t b is givenbyf n (t, t b ) =2n (n = 1, 2, 3,...). (13.10)l(t, t b )(iii) The fraction of the total power emitt<strong>ed</strong> in the nth oscillation mode atfrequency f n is given by the coefficient P n , whereƔ =∞∑P n . (13.11)n=1Analytic and numerical studies suggest that the radiation efficiencycoefficients behave as P n ∝ n −q .This model has several shortcomings. First, the spectral index q has not beenwell determin<strong>ed</strong>. Numerical simulations suggest q = 4/3 while from analyticcalculations q = 2 is obtain<strong>ed</strong>. The simulations have limit<strong>ed</strong> resolution of theimportant small-scale features of the long strings and loops and, therefore, theanalytic pr<strong>ed</strong>iction may be more realistic. Second, the effect of the back reactionon the motion of the strings has been ignor<strong>ed</strong>. The authors of [35] have argu<strong>ed</strong>that this effect result in an effective high-frequency cut off in the oscillation modenumber n. Thus, the loop radiates only in a finite range of frequencies andequation (13.11) is replac<strong>ed</strong> byƔ =n ∗ ∑n=1P n . (13.12)By comparing the back-reaction length scale to the loop size these authorsestimate that the cut off should be no larger than ∼(ƔGµ) −1 .


218 Sources of SGWB-5-6-7-8-9-10 0 10Figure 13.2. The spectrum of gravitational radiation produc<strong>ed</strong> by a string network for agiven set of dimensionless parameters.We are now in a position to construct the differential equation describingthe rate of change of the energy in loops present in a volume V (t) at time t.The numerical integration of this equation and the method appli<strong>ed</strong> to calculatethe power spectrum of gravitational radiation are discuss<strong>ed</strong> in detail in [34].The results of this calculation are shown in figure 13.2. In this paper analyticexpressions for the latter have also been deriv<strong>ed</strong>. Even though simplifi<strong>ed</strong> forconvenience, these analytic expressions offer the opportunity to examine thevarious dependences of the spectrum on string and cosmological parameters.The spectrum of gravitational radiation produc<strong>ed</strong> by a network of strings hastwo main features:• A nearly equal gravitational radiation energy density per logarithmicfrequency interval in the range (10 −8 , 10 10 ) Hz. This portion (‘r<strong>ed</strong> noise’)of the spectrum corresponds to GWs emitt<strong>ed</strong> during the radiation-dominat<strong>ed</strong>era and does not show a significant dependence by the spectral index q andthe cut-off n ∗ [35].


Topological defects 219• A peak near f ∼ 10 −12 Hz. The shape of this portion depends criticallyon the model for the emission by a loop. The overall height of the spectrumdepends linearly on Gµ, while the frequency at which the peak<strong>ed</strong> spectrummerges to the r<strong>ed</strong> noise portion depends inversely on α. The important resultis that for values n ∗ < 10 2 and q ≥ 4/3 the spectrum drops off as 1/f forany value.Because the frequency band is accessible to VIRGO, in the following weshall concentrate on the ‘r<strong>ed</strong> noise’ portion of the spectrum, referring to [34, 35]for details about the region of the peak.An analytic expression for the ‘r<strong>ed</strong> noise’ portion of the GW spectrum isgiven as follows: gw ( f ) = 8π 9AƔG 2 µ 2α(1 −〈v 2 〉) β3/2 − 11 + z eq. (13.13)Let us remark that the above expression for the spectrum does not accountfor the r<strong>ed</strong>uction in the number of relativistic degrees of fre<strong>ed</strong>om that occurs everytime the temperature falls through a particle mass threshold. This has the effectof modifying equation (13.13) by a factorN =(g∗ag ∗b) 1/3where g ∗a (g ∗b ) is the number of relativistic degrees of fre<strong>ed</strong>om at a temperatureabove (below) the particle mass threshold. Within the standard model SU(3) C ⊗SU(2) L ⊗ U(1) Y , one has:⎧1, f ∈ (10 −8 , 10 −10 α −1 ) Hz( )⎪⎨ 3.36 1/3= 0.68, f ∈ (10 −10 α −1 , 10 −4 α −1 ) HzN = 10.75(13.14)( ) 3.36 1/3⎪⎩= 0.32, f ∈ (10 −4 α −1 , 10 8 ) Hz106.75where we take α ∼ ƔGµ in evaluating the above frequency range. Hence, thethermal history of the cosmological fluid reflects on the r<strong>ed</strong> noise spectrum bya series of steps down in amplitude with increasing frequency. The detection ofsuch a shift would provide unique insight into the particle content of the earlyuniverse (at times much earlier than the electroweak phase transition).There is a substantial body of astronomical evidence which suggests thatthe cosmological density parameter 0 is less than one, i.e. the universe is open.The evolution of strings in an open universe will differ from that in the flat caseonly after the time at which the expansion of the universe becomes curvatur<strong>ed</strong>ominat<strong>ed</strong>:after this time the linear regime no longer exists. This is importantfor consideration of GWs creat<strong>ed</strong> with low frequencies during the matter era, but,


220 Sources of SGWBapart from a shift in the frequency corresponding to equal matter-radiation hasa negligible effect on the r<strong>ed</strong> noise spectrum, which is produc<strong>ed</strong> in the radiationera [35].The values of the dimensionless parameters appearing in equation (13.13)are not completely known. Numerical simulations provide a reasonable estimatefor A and 〈v 2 〉 in the radiation-dominat<strong>ed</strong> era: A = 52 ± 10, 〈v 2 〉=0.43 ±0.02. Comparing detail<strong>ed</strong> calculations of large angular scale CMB temperatureanisotropies induc<strong>ed</strong> by strings [36] with observations, the string mass per unitlength has been normaliz<strong>ed</strong> toGµ = 1.05 +0.35−0.20 × 10−6 .This value is below the upper bound obtain<strong>ed</strong> from the pulsar timingand nucleosynthesis constraints on the gravitational radiation spectrum [35].However, the value of α (the size of the loop at formation) is still unknowm. Thehigh-resolution numerical simulations show that α < 10 −2 . This surprisinglysmall relative size is a result of the small-scale structure on the long strings and,since this is cut off by gravitational back-reaction, we may reasonably expectthat α > ƔGµ ≈ 6 × 10 −5 [33]. This uncertanty in the value of α leads toa wide uncertainty in β that, because this parameter governs the lifetime of theloop, has a large effect on the spectrum. However, it is easy to verify that thespectral density gw ( f ) reach a minimum value when α → 0. Therefore, in theradiation-dominat<strong>ed</strong> era (d H = 2t): gw ( f ) ≥ 8π 3 Af r Gµ 1 −〈v2 〉1 + z eqN, f ∈ (10 −8 , 10 10 ) Hz. (13.15)Since 1 + z eq = 2.32 × 10 4 0 h 2 0 , in the case 0 = 1 one has:h 2 0 gw ≥ 5.0 × 10 −9 N.In the case of the standard model thermal scenario, the value of N to be insert<strong>ed</strong> inthis expression is 0.32 (see equation (13.14)). But, as anticipat<strong>ed</strong> in section 12.1,the evolution of N with the temperature is known only up to T ∼ 10 3 , i.e. upto frequencies f ∼ 10 −3 α −1 ∼ 10 Hz. If the particle physics model has mor<strong>ed</strong>egrees of fre<strong>ed</strong>om beyond this temperature ( f ² 10 Hz) there could be othersteps in the function N(T ) associat<strong>ed</strong> with other phase transitions. However,the dependence of the spectral density on the number of degrees of fre<strong>ed</strong>om isreasonably weak and, thus, we can conservatively estimatein the frequency range explor<strong>ed</strong> by VIRGO.h 2 0 gw ≥ 1.6 × 10 −9 (13.16)


Topological defects 22113.1.2 Hybrid defectsIf the symmetry breaking responsible for the formation of the strings occur as apart of a much more complicat<strong>ed</strong> breaking scheme, it is likely that hybrid systemscompos<strong>ed</strong> by topological defects of different dimensionality may form. In thecase when the hybrid system does not annihilate imm<strong>ed</strong>iately this may lead to astochastic background in a similar way to that for the strings [38]. The interestingfeature of these objects is that they evade the constraints from the CMB and pulsartiming allowing for larger values of Gµ and hence larger contributions to theSGWB in the detectable range of frequencies.Walls bound<strong>ed</strong> by stringsDomain walls are form<strong>ed</strong> when a discrete symmetry is broken. The simplestmodel of this sort is that of real scalar field with a potentialV (φ) = 1 2 λ(φ2 − η 2 w )2 .The reflection symmetry group Z 2 of the Lagrangian (invariance under φ →−φ)is spontaneously broken when φ takes on the VEV 〈φ〉 =±η, and so the manifoldÅ consists of only two points. As we go from a region with 〈φ〉 =η to a regionwith 〈φ〉 =−η, we should necessarily pass through 〈φ〉 =0 and, thus, the tworegions must be separat<strong>ed</strong> by a wall of false vacuum. Therefore, the simplestsequence of phase transitions that results in walls bound<strong>ed</strong> by strings isG → H ⊗ Z 2 → Hwhere at first transition (T ∼ η s ) strings form and at the second (T ∼ η w ) eachstring gets attach<strong>ed</strong> to a domain wall.Before the formation of walls the evolution of strings is as in the standardscenario describ<strong>ed</strong> above. After a period of overdamp<strong>ed</strong> motion, the stringsstart moving relativistically at time t s and approach a scaling regime where thecharacteristic scale of the network is comparable to the horizon. After the timet w at which the domain walls form the evolution of the network depends on theratio between the string tension µ ∼ ηs 2 and wall surface tension σ ∼ η3 w . Thewalls become dynamically important at t ∼ µ/σ , when they pull the stringstowards one another, and the network breaks into pieces of wall bound<strong>ed</strong> by string.Alternatively, if t w >µ/σthe breakup of the network occurs imm<strong>ed</strong>iately afterthe wall formation. By indicating with t ∗ = max{µ/σ, t w }, the typical size of thepieces is expect<strong>ed</strong> to be ∼αt ∗ .<strong>Gravitational</strong> <strong>waves</strong> emitt<strong>ed</strong> by oscillating loop of strings during the periodt ∈ (t s , t ∗ ) form a SGWB with a nearly flat spectrum extending over thefrequency range f ∈ [2/α(t ∗ t eq ) 1/2 , 2/α(t s t eq ) 1/2 ]. For this frequency rangeto overlap with the one cover<strong>ed</strong> by VIRGO the strings must decay before th<strong>ed</strong>ecoupling, and, thus, the only constraint on this background comes from big


222 Sources of SGWBbang nucleosynthesis. It can be shown [38] that: gw ( f ) ∼ 6 × 10 −3 Gµh −20with Gµ º 8 × 10−4ln(t ∗ /t s ) . (13.17)Although t s and t ∗ are model-dependent (t s through the assum<strong>ed</strong> cosmologicalscenario; t ∗ through µ and σ ), they do not have to be separat<strong>ed</strong> by many orders ofmagnitude. Following [38] if we assume ln(t ∗ /t s ) º 1 equation (13.17) gives:h 2 0 gw º 5 × 10 −6 . (13.18)At t > t ∗ these string loops spann<strong>ed</strong> by domain walls will collapse intogravitational radiation (and other decay products) in about a Hubble time. Thisprocess takes place over a relatively short frequency range and may lead to a sharppeak in the spectrum. However, since the exact nature of this contribution displaya strong dependence on model phenomenology, it is not possible at present topr<strong>ed</strong>ict whether or where such a peak should occur [37].Strings connect<strong>ed</strong> by monopolesMonopoles are point defects which form when the manifold Å of equivalentvacua contains unshrinkable surfaces. A simple model that illustrates themonopole solution is that of an SO(3) gauge theory spontaneously broken downto U(1) by a Higgs triplet φ a with a potentialV (φ) = 1 2 λ(φa φ a − η 2 m )2 , (a = 1, 2, 3).The magnitude of 〈φ a 〉 is fix<strong>ed</strong> by the minimization of the potential to |φ| =η m ,but its direction in group space is not and the manifold Å is a sphere in thisspace. If we choose φ(⃗r) = (0, 0,η m ) the SO(3) symmetry is broken down toU(1) because φ is invariant under rotations about the ê 3 -axis. Another, less trivial,choice is represent<strong>ed</strong> byφ i (⃗r) = η m ˆr iwhere ˆr i is a unit vector in coordinate space. For this configuration of the fieldthere must be a point in space where φ = 0 and the energy density is nonvanishing.This point of false vacuum is a monopole.Therefore, the prototypical sequence of symmetry breakings resulting inmonopoles connect<strong>ed</strong> by strings isG → H ⊗ U(1) → Hwhere monopoles form<strong>ed</strong> at first transition (T ∼ η m ) and strings get connect<strong>ed</strong>by monopole/antimonopole (M ¯M) pairs at the second (T ∼ η s ). If the monopoleformingtransition occurs after any period of inflation which can have takenplace, then the average separation of the monopoles is always smaller than the


Inflation 223Hubble radius and when the M ¯M pairs get connect<strong>ed</strong> by strings the hybrid systemcollapse in less than one Hubble time by dissipating energy into friction with thecosmological fluid.For our purposes, the most interesting scenario is when the monopoles areform<strong>ed</strong> during the inflation but are not completely inflat<strong>ed</strong> away. The strings areform<strong>ed</strong> later with a length scale ξ that is much smaller than the average monopoleseparation d. In the course of the evolution ξ grows like t and eventually becomescomparable to d, so that at some time t m we are left with M ¯M pairs connect<strong>ed</strong> bystrings. If the strings are form<strong>ed</strong> during inflation, soon after the monopoles, theydo not go through a period of relativistic evolution and no gravitational radiation isproduc<strong>ed</strong> prior to t m . In contrast, if the strings are form<strong>ed</strong> in the post-inflationaryepoch, they can have a period of relativistic evolution and a nearly flat stochasticbackground identical to that for walls bound<strong>ed</strong> by strings is produc<strong>ed</strong> before t m .At t > t m , the M ¯M pairs oscillate and gradually convert their energy ingravitational radiation. This process has been studi<strong>ed</strong> in detail in [38] for thesimplest case of a straight string connecting the monopoles. It was found that thespectral density of the radiation emitt<strong>ed</strong> by this simple configuration verify theboundh 2 0 gw º 2 × 10 −8 (13.19)in the frequency range of VIRGO.13.2 InflationIt is well known that many of the shortcomings of the standard cosmologicalmodel (such as isotropy of the CMBR, structure formation, flatness and monopoleproblems) can be successfully fac<strong>ed</strong> in the framework of the so-call<strong>ed</strong> inflationarymodels (see [39, 40] for a review).The basic idea shar<strong>ed</strong> by almost all the various models [41–44] of inflationis that in early times the universe was dominat<strong>ed</strong> by the vacuum energy ofsome scalar field, which provid<strong>ed</strong> an exponential growth of the scale factorof the universe; then, as a result of a phase transition (maybe associat<strong>ed</strong> witha spontaneous symmetry breaking), the scalar field was captur<strong>ed</strong> by the trueminimum of its potential, made some oscillations and finally settl<strong>ed</strong> down in it.The energy previously stor<strong>ed</strong> in the false vacuum was convert<strong>ed</strong> into the decayproducts of the scalar field, produc<strong>ed</strong> mainly during the oscillatory stage: thisprocess is responsible for the reheating of the universe, thus providing the usual‘hot’ initial conditions for the beginning of the radiation-dominat<strong>ed</strong> era.Despite the fact that a complete and satisfying model of inflation is not yet athand, the ‘inflationary paradigm’, i.e. the idea of a primordial stage of accelerat<strong>ed</strong>expansion, is widely accept<strong>ed</strong> and consider<strong>ed</strong> as a necessary ingr<strong>ed</strong>ient of everycosmological model. Moreover, an inflationary stage provides a very interestingmechanism of amplification of perturbations (see [47]) that, as well as beingimportant for what concerns the problem of the structures formation, generates


224 Sources of SGWBalso a stochastic background of gravitational <strong>waves</strong>. The basic concept is quitesimple: zero point quantum fluctuations of every sufficiently light field can beamplifi<strong>ed</strong> when the size of the perturbation becomes larger than the event horizon.This is the classical interpretation of the well-known fact that a time-dependentgravitational field produces particles even if the initial state contains no quanta(see [45] for details).13.2.1 Classical pictureSince the final number of quanta creat<strong>ed</strong> is typically very large, it comes out thata classical analysis of the phenomenon is appropriate and gives the right answer.So, let us show how this amplification works by studying the classical equation ofmotion for gravitational perturbations; we will return to the quantum descriptionlater.The starting point is to separate the metric into a ‘background’ 1 and a‘propagating’ part:g µν ∼ R 2 (η)(−dη 2 + d⃗x 2 ) + h µν ,where R is the scale factor of the universe, depending only on the conformal timeη (in terms of the usual cosmic time dη = dt/R(η)). The equation of motion forh µν in the FRW background is obtain<strong>ed</strong> by taking the first-order variation of theEinstein equations. One can defineh µ ν ∼ 1R(η) εµ ν ( ⃗k)ψ(η)e i ⃗k·⃗x , (13.20)where ε ν µ (⃗k) is the polarization tensor. It comes out that, with a suitable gaugechoice, ψ obeys the following equation:( )ψ ′′ (η) + k 2 − R′′ (η)ψ ∼ 0, (13.21)R(η)where ‘ ′ ’ denotes the derivative with respect to the conformal time. This is aSchrödinger-like equation with the potential given by V = R ′′ /R. Fork ≫ V (η)we have for ψ the obvious plane wave solution ψ ∼ e −ikη , so that |h µ ν |∼1/R;if, instead, k ≪ V (η), we get the two following solutions:ψ 1 ∼ R(η),∫ψ 2 ∼ R(η)dηR 2 (η) .As will be seen later, in the case of our interest ψ 1 is the dominant solution, so, inthis case, |h µ ν |∼1. This means that a solution characteriz<strong>ed</strong> by a long wavelength(k ≪ V (η)) is amplifi<strong>ed</strong> with respect to a short wavelength one by a factor R(η);1 For the sake of simplicity we consider a spatially flat universe.


Inflation 225as a first approximation we can assume that this amplification takes place as longas k < V (η). The last inequality, apart from irrelevant numerical factors due toour approximation, turns out to have the same analytical form of the condition:kR < R′R 2 ≡ H (η),that relates the physical momentum of the perturbation to the Hubble parameterH , whose inverse corresponds to the horizon size: this is just the conditionanticipat<strong>ed</strong> above, i.e. that a perturbation is amplifi<strong>ed</strong> when its wavelengthbecomes larger than the event horizon.13.2.2 Calculation of the spectrumTo calculate the spectrum of perturbations we ne<strong>ed</strong> to find particular solutionsof the Schrödinger-like equation previously introduc<strong>ed</strong>. Let us, then, specify ourmodel: the form of the potential V is determin<strong>ed</strong> by the evolution of R(η); forour purposes it is sufficient to consider a three-stage model in which, as shownin figure 13.3, a de Sitter phase (characteriz<strong>ed</strong> by a constant Hubble parameterH ds ) is follow<strong>ed</strong> by a radiation-dominat<strong>ed</strong> (RD) and, then, by a matter-dominat<strong>ed</strong>(MD) era:⎧− 1H ds η , η < η 1 < 0 de Sitter⎪⎨ η − 2η 1R(η) ∼H ds η 2 , η 1


226 Sources of SGWBR 2MDR(η) (arbitrary units)R 110 0de SitterRD1 η2 1 3 η4 2 η50η (arbitrary units)Figure 13.3. The scale factor as a function of the conformal time. The universe undergoesa transition from a de Sitter- to a radiation-dominat<strong>ed</strong> phase at the time η 1 , and from aradiation to a matter dominat<strong>ed</strong> phase at the time η 2 . The present epoch corresponds to η 0 .At first sight, the first and the third modes written in equation (13.24), seemnot to be pure plane <strong>waves</strong> solutions. This comes from the fact that we are incurv<strong>ed</strong> spacetime: it can be easily verifi<strong>ed</strong> that in the limit k →∞, i.e. when thewavelength is so short that the particle does not ‘feel’ the curvature of spacetime,all the modes in equation (13.24) r<strong>ed</strong>uce to the standard plane-wave form. Thefact that ψ r± is already in the standard form is due to the conformal invariance ofthe radiation-dominat<strong>ed</strong> spacetime.The coefficients α, β, γ and δ in equation (13.23) are calculat<strong>ed</strong> by requiringthe overall solution ψ(η) to be continuous with its first derivative at transitiontimes η 1,2 . Equation (13.21) says that this problem is similar to the wellknownquantum mechanics problem of tunnelling through a potential barrierV (η). Therefore, α, β, γ and δ can be consider<strong>ed</strong> as the transmission/reflectioncoefficients and one can show that the number of creat<strong>ed</strong> gravitons with comovingfrequency k is( )3HN k ∼|δ k | 2 3= dsR(η 1 ) 4 218R(η 2 ) k 6 . (13.25)This expression should be taken with care, because the simple modeldescrib<strong>ed</strong> by equation (13.22) does not take into account two important physicaleffects that force equation (13.25) to hold only in a limit<strong>ed</strong> range of frequencies.First of all, the perturbations whose physical wavelength is greater than thepresent Hubble length (R(η 0 )/k > 1/H 0 ) do not contribute to the energy density;so k 0 = R ′ 0 /R 0, corresponding to a physical frequency f 0 ∼ 10 −19 Hz, provides


Inflation 227the lower bound to the extension of the spectrum. To find the upper boundone must note that in our model the transition between the different regimestakes place instantaneously; in a more realistic situation such transitions havea typical duration time t, and consequently there is a cut-off physical frequencyof order 1/t. In our case the typical time of the transitions is of the order ofthe Hubble parameters at the transition points (H (η 1 ), H (η 2 )). This means thatfor k > k 2 = R2 ′ /R 2, i.e. for f ≥ 10 −16 Hz, equation (13.25) is not the correctexpression for N k ; above this frequency the radiation-matter transition does notaffect the spectrum, but the inflation-radiation one is still important. Hence,the number of gravitons creat<strong>ed</strong> is given by the corresponding equation, with δreplac<strong>ed</strong> by β( )HN k =|β k | 2 2= dsR(η 1 ) 2 22k 2 . (13.26)In turn, equation (13.26) has a high frequency cut-off given by k 1 = R1 ′ /R 1 =−1/η 1 ; η 1 , and consequently the corresponding value of the physical cut-offfrequency f 1 , is an almost free parameter of the model and depends on thereheating temperature. A typical value for f 1 is the one report<strong>ed</strong> in [46] ( f 1 ∼10 10 Hz), but other (also much lower) values are possible, depending on themodel; in general one has f 1 > 1 kHz. Beyond f 1 there is no amplificationmechanism at all and the spectrum goes rapidly to zero.Once obtain<strong>ed</strong> the result for N k , it is straightforward to obtain its contributionto ρ gw , and consequently to gwdρ gw ( f ) = 2 · 2π f · N k ( f ) · 4π f 2 d f,where the factor of two is due to polarization, and f = k/2π R 0 is the physicalfrequency. The final result for gw in terms of the physical frequency is then⎧ ( )3H3 2 ( ) 2 ( ) 6ds R1 R1 18⎪⎨ 16π R 2 R 0 f 2 , k 0 < 2π f < k 2 gw ( f ) =3π H0 2M2 H 4 ( ) 4Pl ds R1, k 2 < 2π f < k 1⎪⎩ 4 R 00, 2π f > k 1(13.27)where M Pl is the Planck mass and H 0 is the present value of the Hubble parameter.Because of the flatness of gw ( f ) the strongest constraint turns out to be theone relat<strong>ed</strong> to the observ<strong>ed</strong> anisotropy in CMBR. This constrains the spectrum asfollows( ) 2h 2 0 gw ≤ 7 × 10 −11 H0for f ∈ (10 −19 , 10 −16 ) Hz, (13.28)fwhich in turn impliesH ds ≤ 10 39 Hz ≃ 5 × 10 14 Gev ≃ 5 × 10 −5 M Pl .


228 Sources of SGWB−6−9−12−15−18−20 −15 −10 −5 0 5log (f / Hz)Figure 13.4. h 2 0 gw against the physical frequency f (logarithmic scales). Heref 1 = 100 kHz.Given the shape of the spectrum (see figure 13.4) one obtains in thefrequency region of interest for VIRGO the following bound:h 2 0 gw < 8 × 10 −14 , (13.29)many orders of magnitude below the sensitivity limit of the plann<strong>ed</strong> detectors.Finally we remark that the result report<strong>ed</strong> in equation (13.27) is valid ifthe Hubble parameter during inflation is strictly constant; the calculation can b<strong>ed</strong>one also for other, more realistic models, but the final result is not significantlydifferent from the one report<strong>ed</strong> above. For instance, in the so-call<strong>ed</strong> ‘slowroll’inflation the Hubble parameter is slightly decreasing during the inflationarystage and makes the spectrum slightly tilt<strong>ed</strong>, instead of constant, in the radiationdominat<strong>ed</strong>region. Anyway the tilt is so small that the value of gw at interestingfrequencies is different from the one of equation (13.29) only for an order ofmagnitude (in addition this correction has the ‘wrong’ sign, i.e. it lowers thebound!).


String cosmology 22913.3 String cosmologyThe basic mechanism of generation of relic gravitational <strong>waves</strong> in cosmology hasbeen discuss<strong>ed</strong> in the prec<strong>ed</strong>ing section about inflation where the main conceptshave been expos<strong>ed</strong>. The crucial point of the outcome was the flatness of thespectrum which, combin<strong>ed</strong> with the COBE bound at very low frequency, givesa very strong constraint on the spectrum even at high frequency. To satisfy theCOBE bound and still have a chance of being observable at VIRGO or LIGO thespectrum must grow with frequency. A spectrum of this kind has been found in acosmological model suggest<strong>ed</strong> by string theory [48].In string theory the fundamental objects are one-dimensional extend<strong>ed</strong>entities, i.e. strings. Their fundamental excitation of given energy and angularmomentum are particles of given mass and spin. String theory has one onlyfundamental (dimensionful) constant: the string tension T which can be trad<strong>ed</strong>for a fundamental length λ s ≡ √¯hc/T . The mass scale of the excitation of thestring is therefore √ T which, as string theory includes gravity, should be near thePlanck mass 2 . The gauge couplings are not constant (neither at classical level)but depend on the expectation value of a scalar field, the dilaton. For example,Newton’s gravitational constant G is given byG ∼ λ s28π eφ ,while all the gauge couplings are proportional to e φ/2 . In the regime where allcouplings and derivatives are small string physics can be suitably describ<strong>ed</strong> by afield theory action which describes the dynamics of the light (massless) fields ofthe string (the graviton g µν and the dilaton φ), which isS =− 1 ∫2λ 2 d 4 x √ −g[e −φ (Ê + ∂ µ φ∂ µ φ) + higher derivatives]s+ [higher order in e φ ]. (13.30)where g = det ‖g µν ‖, and Ê is the Ricci scalar. Higher derivative termsare relevant whenever spacetime derivatives become of the order of one (in λ sunits), whereas the higher orders in e φ acquire importance when e φ ∼ 1, thussignalling the beginning of a full stringy-quantum regime. The first terms withoutcorrections reproduce Einsteinian gravity for constant dilaton.To analyse a situation of cosmological interest it suffices to study spatiallyhomogeneous fields, i.e. φ = φ(t) and the metric is chosen so that the line elementds 2 can be written asds 2 = dt 2 − R 2 (t) d⃗x 2 , (13.31)2 This is not a compelling argument as models have been present<strong>ed</strong> in which √ T can be consistentlyput some orders of magnitude below the Planck mass (see, e.g., [49]), but we will not treat thesemodels.


230 Sources of SGWBwhich is a suitable ansatz for the metric describing an evolving homogeneousisotropic universe. The cosmological field equations deriv<strong>ed</strong> from the low-energypart of (13.30) have the symmetry (in four dimensions)R(t) → 1R(−t) ,φ(t) → φ(−t) − 6ln[R(−t)]which relates ordinary FRW cosmology characteriz<strong>ed</strong> by H ≡ Ṙ/R > 0, Ḣ < 0,constant φ (where a dot means a derivative with respect to t) att > 0, with aninflationary one with H > 0, Ḣ > 0 and ˙φ > 0att < 0. This symmetrymay suggest that the universe start<strong>ed</strong> its evolution from the state of perturbativevacuum, i.e. empty, cold, flat and decoupl<strong>ed</strong> with increasing Hubble parameterand eventually emerg<strong>ed</strong> in the standard cosmology at t > 0 with decreasing H and‘frozen’ dilaton. The dual cosmology is call<strong>ed</strong> pre-big bang (PBB) phase and itdoes not ne<strong>ed</strong> a beginning time as H → 0 and φ →−∞for t →−∞. However,the low-energy equations of motion do not smoothly interpolate between thesetwo phases. Instead they lead to singularities, in the past for the FRW phase(as ordinary cosmology) and in the future for the PBB phase; in both cases att = 0 where big bang should be plac<strong>ed</strong>. However, the approximations on whichthe validity of equation (13.30) relies, break down whenever H ∼ O(λ s −1)or φ ∼ 0. The corrections indicat<strong>ed</strong> in equation (13.30) may prevent reachingthe singularity, allowing a smooth transition to standard hot big bang and FRWcosmology. Actually there are indications (see, e.g., [50–54]) that a regularizationmechanism can be really provid<strong>ed</strong> by taking into account of quantum and stringyeffects on the evolution of the system. The big bang should then be identifi<strong>ed</strong> withthe epoch of maximum but not infinite curvature which should be follow<strong>ed</strong> by apost-big bang evolution, by which we mean a standard evolution with = 1, asit has been achiev<strong>ed</strong> by the long period of PBB inflation. The above describ<strong>ed</strong>scenario is display<strong>ed</strong> in figure 13.5.Nevertheless in the strong coupling regime physics is really new and theperturbative approach propos<strong>ed</strong> by equation (13.30) is no more valid. One shouldresort to a new picture, with an effective action written in terms of the newlight modes appropriate to the strong coupling regime [55]. Anyway, admittingthat a regularization is still possible makes this picture noteworthy from boththe theoretical and the phenomenological point of view. From the theoreticalside it addresses issues left open by standard inflationary models such as theinitial singularity and the theoretical origin of the field driving inflation; fromthe phenomenological side in the next subsection the production of gravitational<strong>waves</strong> (studi<strong>ed</strong> in [56, 57]) will be analys<strong>ed</strong>.13.3.1 The modelThe solution of the equations of motion deriv<strong>ed</strong> from equation (13.30) whichpresents the initial conditions typical of PBB phase are (for time-dependent only


String cosmology 231HPre-big bangPost-big bang0.0Figure 13.5. Qualitative behaviour of H in the suggest<strong>ed</strong> cosmological model. Brokencurves indicate the lowest order solution singular at t = 0. At around t = 0 a regularevolution interpolating between the two phases is display<strong>ed</strong>.0.0tfields):R(η) =− 1H s η s( η − (1 − α)ηsαη s) −α, (13.32)andφ(η) = φ s − γ ln η − (1 − α)η s,η swhere η is the conformal time (see section 13.2) ranging between −∞ and η s < 0,α = ( √ 3 − 1)/2, and γ = √ 3.At a value η = η s the curvature becomes of order of the string scale and thelowest order effective action does not give any more a good description of physics:we enter a ‘full stringy’ regime. One expects that higher order corrections to theeffective action tame the growth of the curvature, and both H and dφ/dt stayapproximatively constant. In terms of conformal time this meansR(η) =− 1H s η , φ(η) = φ s − 2β ln η η s, (13.33)where β is a free parameter which measures the growth of the dilaton in thisphase, which lasts for η s


232 Sources of SGWBdominat<strong>ed</strong> phase characteriz<strong>ed</strong> byR(η) = 1H s η12 (η − 2η 1 ), φ = φ 0 . (13.34)The time-dependent component of a metric perturbation h ν µ can be express<strong>ed</strong> asin equation (13.20), where ψ obeys the Schrödinger-like equation (13.21) withpotential given, in this case, by:V (η) = eφ/2Rd 2dη 2 (Re−φ/2 ).Taking into account equations (13.32) and (13.34) one finds⎧1 4ν 2 − 1⎪⎨ 4 [η − (1 − α)η s )] 2 , −∞


String cosmology 233Figure 13.6. h 2 0 gw( f ) as a function of f per f s = 100 Hz,H s = 0.15M Pl , f 1 = 4.3 × 10 7 kHz, µ = 1.4 compar<strong>ed</strong> with the low- andhigh-frequency limits (from [57]).by the growth of the dilaton during the string phase; H s , the constant value of theHubble parameter during the string phase, which we expect to be of the order ofλ s −1, η s and η 1 . It is convenient to trade η s and η 1 with the associat<strong>ed</strong> frequenciesf s = 1/(2π R 0 |η s |), being R 0 the present value of the scale factor of the universe,and f 1 defin<strong>ed</strong> analogously; f 1 can be estimat<strong>ed</strong> to be f 1 ∼ 10 GHz and itcorresponds to the maximum amplifi<strong>ed</strong> frequency, whereas the only constrainton f s is f s < f 1 . As shown in figure 13.6, the spectrum presents an f 3 raise forf < f s (and f s ≪ f 1 ) and a series of oscillations around the line of slope 3 − 2µfor f s < f < f 1 , while for frequency higher than f 1 the spectrum is exponentiallysuppress<strong>ed</strong>, as it corresponds to the maximum of the potential in equation (13.35).In the most favourable case (µ = 1.5 and f s less than the smaller frequency in theVIRGO frequency range), the maximum value of the spectrum is reach<strong>ed</strong> as longas f > f s [57]h 2 0 max gw ≃ 3.0 × 10−7 (Hs0.15M Pl)(t1λ s) 2,being t 1 the time of transition to the post-big bang phase.This is just an example of possible cosmological dynamics driven by stringtheory, as it stands it cannot be identifi<strong>ed</strong> as the pr<strong>ed</strong>iction of string cosmology.Nevertheless the low-frequency behaviour, i.e. the power-law raise, is a quite


234 Sources of SGWBgeneral feature which belongs to any PBB-dictat<strong>ed</strong> spectrum of gravitational<strong>waves</strong>.13.3.2 Observational bounds to the spectrumBecause of the power raise of the spectrum for low frequency the COBE boundis easily evad<strong>ed</strong> (see figure 13.7) and the same statement holds for the constraintderiv<strong>ed</strong> from pulsar timing observation as long as f s < 10 −7 Hz. The moststringent bound is then the nucleosynthesis one given in section 12.5. In the mostfavourable case (a flat spectrum gw ( f ) = maxgw for f s < f < f 1 , which meansµ = 1.5) that bound translates intoh 2 0 max gw ln f 1< 6.3 × 10 −6 .f sIn order to have experimentally interesting values, the spectrum must have alreadyreach<strong>ed</strong> the maximum value gwmax in the VIRGO frequency range. Under thisassumption, a favourable choice of f s is 100 Hz, that leads toh 2 0 max gw < 3.2 × 10−7 . (13.36)The effect of the observational bound on the parameters of the model is display<strong>ed</strong>in figure 13.8, where the independent parameters are chosen to be η s /η 1 = f 1 /f sand β (see equation (13.33)).Figure 13.7. h 2 0 gw( f ) as a function of f per f s = 10 Hz, f 1 = 4.3 × 10 7 kHz, µ = 1.5compar<strong>ed</strong> with observational bounds (from [57]).


First-order phase transitions 2358070Ω=10 −960ln(η s/η 1)5040Ω=10−830Ω=10−720100.00 0.05 0.10 0.15 0.20βFigure 13.8. The forbidden region in the parameter space is the shad<strong>ed</strong> area. Along thechain line ≡ h 2 0 gw(1 kHz) = 10 −7 , along the dott<strong>ed</strong> line = 10 −8 and along th<strong>ed</strong>ash<strong>ed</strong> line = 10 −9 (from [58]).The maximum value obtain<strong>ed</strong> for gw is still below the experimental plann<strong>ed</strong>sensibility of VIRGO and LIGO, but hopefully within the plann<strong>ed</strong> sensitivity ofthe advanc<strong>ed</strong> projects.13.4 First-order phase transitionsFirst-order phase transitions are thought to have occurr<strong>ed</strong> in the early stage of theexpansion of the universe, each time its temperature dropp<strong>ed</strong> sufficiently belowthe critical temperature T c of a transition. Candidates for such a phase transitioninclude GUT-symmetry breaking (T c = 10 15±1 GeV), electroweak-symmetrybreaking (T c ∼ 200 GeV), and phase transitions yet to be discuss<strong>ed</strong>.In a first-order phase transition, the universe starts in a metastable (highenergy-density)false-vacuum state. If the energy barrier separating this stateto the (low-energy-density) true-vacuum state is sufficiently large, significantsupercooling occurs and the transition between these two states proce<strong>ed</strong>s vianucleation [59] of bubbles of true vacuum within the false vacuum phase [60].Once nucleat<strong>ed</strong>, each bubble larger than a critical size begins to expand, beingdriven by the energy difference (latent heat) between its true-vacuum interiorand the false-vacuum exterior. This energy difference is convert<strong>ed</strong> into kineticenergy of the bubble wall, which become thinner and more energetic as thebubble expands, and rapidly approaches a velocity near the spe<strong>ed</strong> of light. As


236 Sources of SGWBa consequence of the high velocities and energy densities involv<strong>ed</strong>, when, withina Hubble expansion time, these bubbles collide a large fraction of the energy thatwas in the bubble walls is convert<strong>ed</strong> in gravitational radiation.Unlike the case of a network of cosmic strings, where gw ( f ) is flat as aconsequence of the existence of a ‘scaling’ solution, the spectrum of the GWsproduc<strong>ed</strong> by this bubble collision process is strongly peak<strong>ed</strong> at a frequencycharacteristic of the particular cosmological time at which the phase transitionand bubble collisions took place. Under the assumption that the expansion of theuniverse has been adiabatic since the phase transition, for the present value of thischaracteristic frequency one has [61]:( )( ) βf max ≈ 5.2 × 10 −8 T∗ ( g∗) 1/6(13.37)H ∗ 1 GeV 100where, assuming an exponential bubble nucleation rate, β −1 is, roughly, th<strong>ed</strong>uration of the phase transition (see section 12.1 for the meaning of the otherquantities). In general, for the temperatures of interest (say, 1–10 16 GeV), it turnsout to be [62]:( )β MPl∼ 4ln ∼ 100, (13.38)H ∗ T ∗and, because g ∗ is also of order 100 in typical GUT models, from equation (13.37)we find that the most ‘promising’ cosmological phase transition, from thepoint of view of the VIRGO sensitivity band, would be one that occurr<strong>ed</strong> ata temperature compris<strong>ed</strong> between 10 7 and 10 8 GeV (for which we have nocompelling candidate).The amplitude of the spectrum depends mostly upon the difference in the freeenergy between the inside and the outside of the bubble, driving the expansion ofthe bubble. By indicating with v the propagation velocity of the bubble walls, andwith α the ratio of vacuum energy to the thermal energy in the symmetric phase(the high-temperature phase before the transition), the amplitude at the frequencyf max is approximately [61]:h 2 0 gw( f max ) ≈ 1.1 × 10 −6 κ 2 (H∗β) 2 ( α) ( 21 + α) (100v 3 ) 1/30.24 + v 3 ,g ∗(13.39)where k is an increasing function of α quantifying the fraction of the availablevacuum energy that goes into kinetic (rather than thermal) energy of the fluid.The parameter α characterizes the strength of the phase transition, and the limitsα → 0, α → ∞ correspond to very weak and very strong first-order phasetransitions, respectively. For α ranging between these two extremes, typically,0.01 < κ < 1. For strongly first-order phase transitions it turns out to bealso v → 1, and, thus, for a transition at T ∼ 2 × 10 7 GeV, according toequations (13.37) and (13.39), one finds:h 2 0 gw( f ∼ 100 Hz) ≈ 10 −10 (13.40)


Astrophysical sources 237which is three orders of magnitude too small to be observ<strong>ed</strong> by any of the groundbas<strong>ed</strong>interferometers presently under construction.13.5 Astrophysical sourcesTo the stochastic background of gravitational radiation that should pervadeour universe, the GWs associat<strong>ed</strong> with unresolv<strong>ed</strong> astrophysical sources alsocontribute. Since the generation of these <strong>waves</strong> dates back to more recent epochs,when galaxies and stars start<strong>ed</strong> to form and evolve, their contribution to gw isnot subject to the nucleosynthesis bound. Therefore, our first concern is whetherthis component of the stochastic background can give a contribution to h 2 0 gw( f )larger than the bound (12.64), or anyway larger than the expect<strong>ed</strong> relic signal,thereby masking the background of cosmological origin.A first observation is that there is a maximum frequency at whichastrophysical sources can radiate. This comes from the fact that a source of massM, even if very compact, will be at least as large as its gravitational radius 2GM,the bound being saturat<strong>ed</strong> by black holes. Even if its surface were rotating at thespe<strong>ed</strong> of light, its rotation period would be at least 4πGM, and the source cannotemit <strong>waves</strong> with a period much shorter than that. Therefore, we have a maximumfrequency [63],1f º4πGM ∼ M 104 ⊙Hz. (13.41)MTo emit near this maximum frequency an object must presumably have a massof the order of the Chandrasekhar limit ∼1.2M ⊙ , which gives a maximumfrequency of the order of 10 kHz [63], and this limit can be saturat<strong>ed</strong> onlyby very compact objects (see [64] for a recent review of GWs emitt<strong>ed</strong> in thegravitational collapse to black holes, with typical frequencies f º 5 kHz). Thesame numbers, apart from factors of order one, can be obtain<strong>ed</strong> using the fact thatfor a self-gravitating Newtonian system with density ρ, radius R, there is a naturaldynamical frequency [65]f dyn = 1( ) 3GM 1/22π (πGρ)1/2 =16π 2 R 3 . (13.42)With R ≥ 2GM we recover the same order of magnitude estimate apart from afactor (3/8) 1/2 ≃ 0.6. This is already an encouraging result, because it showsthat the natural frequency domains of cosmological and astrophysical sources canbe very different. We have seen in section 12.1 that the natural frequency scale forPlanckian physics is the GHz, while no astrophysical objects can emit above, say,(6–10) kHz. Therefore, a stochastic background detect<strong>ed</strong> above these frequencieswould be unambiguously of cosmological origin.However, ground-bas<strong>ed</strong> interferometers have their maximum sensitivityaround 100 Hz, where astrophysical sources hopefully produce interestingradiation (since these sources were the original motivation for the construction


238 Sources of SGWBof interferometers). The radiation from a single source is not a problem, sinceit is easily distinguish<strong>ed</strong> from a stochastic background. The problem arises ifthere are many unresolv<strong>ed</strong> sources. (Of course this is a problem from the pointof view of the cosmological background, but the observation of the astrophysicalbackground would be very interesting in itself; techniques for the detection of thisbackground with a single interferometer using the fact that it is not isotropic andexploiting the sideral modulation of the signal have been discuss<strong>ed</strong> in [66].)The stochastic background from rotating neutron stars has been discuss<strong>ed</strong>in [67] and references therein. The main uncertainty comes from the estimateof the typical ellipticity ɛ of the neutron star, which measures its deviation fromsphericity. An upper bound on ɛ can be obtain<strong>ed</strong> assuming that the observ<strong>ed</strong>slowing down of the period of known pulsars is entirely due to the emission ofgravitational radiation. This is almost certainly a gross overestimate, since mostof the spin down is probably due to electromagnetic losses, at least for Crab-likepulsars. With realistic estimates for ɛ, [67] gives a value of h 2 0 gw(100 Hz) ∼10 −15 . This is very far from the sensitivity of even the advanc<strong>ed</strong> experiments. Anabsolute upper bound can be obtain<strong>ed</strong> assuming that the spin down is due only togravitational losses, and this gives h 2 0 gw(100 Hz) ∼ 10 −7 , but again this valueis probably a gross overestimate. Very recently the gravitational background hasalso been estimat<strong>ed</strong> [68] produc<strong>ed</strong> by a cosmological population of hot, youngand rapidly rotating neutron stars. Within a reasonable range of values of themain parameters which characterize the energy spectrum of a single source, theauthors of [68] find that h 2 0 gw( f ) show a long plateau extending from ≈300 Hzup to ≈1.7 kHz, with an amplitude of ≈(2.2–3.3) × 10 −8 .In [69] the stochastic background has been consider<strong>ed</strong> emitt<strong>ed</strong> by acosmological population of core-collapse supernovae for a range of progenitormasses leading to a black hole. The expect<strong>ed</strong> frequencies in this case are ofthe order of kHz or lower, depending on the r<strong>ed</strong>shift when these objects areproduc<strong>ed</strong>. Using the observational data on the star formation rate, it turnsout that the duty cycle, i.e. the ratio between the duration of a typical burstand the typical time interval between successive bursts is low, of order 0.01.Therefore, this background is not stochastic, but rather like a ‘pop noise’, andcan be distinguish<strong>ed</strong> from a really stochastic background. The value of h 2 0 gwfor the background from supernovae have been comput<strong>ed</strong> in [69] assumingaxially symmetric collapse, and assuming that all sources have the same valueof a = J/(GM 2 ), where J is the angular momentum. The results depend onthe value of a, and on h 0 . Assuming h 0 = 0.5 and typical values of a, onefinds that gw ( f ) has a maximum amplitude ranging between 10 −11 –10 −10 inthe frequency interval (1.5–2.5) kHz.These results suggest that astrophysical backgrounds might not be a problemfor the detection of a relic background at VIRGO/LIGO frequencies. Thesituation is different in the LISA frequency band [63, 65, 67, 70, 71]. LISA canreach a sensitivity of order h 2 0 gw ∼ afew×10 −13 at f ∼ 10 −3 Hz (see figure 1.3


References 239of [71]). However, for frequencies below a few mHz, one expects a stochasticbackground due to a large number of galactic white dwarf binaries. The estimateof this background depends on the rate of white dwarf mergers, which is uncertain.With rates of order 4 × 10 −3 per year (which should be a secure upper limit [67]),the background can be as large as h 2 0 gw ∼ 10 −8 at f = 10 −3 Hz. This number isactually quite uncertain, and in [71], it is us<strong>ed</strong> another plausible rate, which givesfor instance h 2 0 gw ∼ 10 −11 –10 −10 at f = 10 −3 Hz. Above a frequency of theorder of a few times 10 −2 Hz, most signals from galactic binaries can be resolv<strong>ed</strong>individually and no continuous background of galactic origin is presently knownat the level of sensitivity of LISA.It should be observ<strong>ed</strong> that, even if an astrophysical background is present,and masks a relic background, not all hopes are lost. If we understand wellenough the astrophysical background, we can subtract it, and the relic backgroundwould still be observable if it is much larger than the uncertainty that we haveon the astrophysical background. In fact, LISA should be able to subtract thebackground due to white dwarf binaries, since there is a large number of binariesclose enough to be individually resolvable [71]. This should allow us to pr<strong>ed</strong>ictwith some accuracy the space density of white dwarf binaries in other parts of theGalaxy, and therefore to compute the stochastic background that they produce.Furthermore, any background of galactic origin is likely to be concentrat<strong>ed</strong> nearthe galactic plane, and this is another handle for its identification and subtraction.The situation is more uncertain for the contribution of extragalactic binaries,which again can be relevant at LISA frequencies. The uncertainty in the mergingrate is such that it cannot be pr<strong>ed</strong>ict<strong>ed</strong> reliably, but it is believ<strong>ed</strong> to be lower thanthe galactic background [67]. In this case the only handle for the subtractionwould be the form of the spectrum. In fact, even if the strength is quite uncertain,the form of the spectrum may be quite well known [71].References[1] Kolb E and Turner M 1990 The Early Universe (New York: Addison-Wesley)[2] Kosowsky A, Kamionkowski M and Turner M 1994 Phys. Rev. D 49 2837[3] Fixsen D J et al 1996 Astrophys. J. 473 576[4] Allen B 1996 Relativistic gravitation and gravitational radiation Proc. Les HouchesSchool on Astrophysical Sources of <strong>Gravitational</strong> Waves <strong>ed</strong> J Marck and J Lasota(Cambridge: Cambridge University Press)Allen B 1996 Preprint gr-qc/9604033[5] Maggiore M 1998 High-energy physics with gravitational-wave experiments grqc/9803028[6] Birrel N and Davies P C W 1982 Quantum Fields in Curv<strong>ed</strong> Space (Cambridge:Cambridge University Press)[7] Allen B and Romano J 1999 Phys. Rev. D 59 10<strong>2001</strong>[8] Michelson P 1987 Mon. Not. R. Astron. Soc. 227 933[9] Christensen N 1992 Phys. Rev. D 46 5250[10] Flanagan E 1993 Phys. Rev. D 48 2389


240 Sources of SGWB[11] Thorne K S 1987 300 Years of Gravitation <strong>ed</strong> S Hawking and W Israel (Cambridge:Cambridge University Press) p 330[12] Misner C, Thorne K and Wheeler J 1973 Gravitation (New York: Freeman)[13] Zhou C Z and Michelson P 1995 Phys. Rev. D 51 2517[14] Cuoco E, Curci G and Beccaria M 1997 Adaptive identification of VIRGO-like noisespectrum gr-qc/9709041[15] Astone P et al 1997 Astropart. Phys. 7 231[16] Coccia E, Fafone V, Frossati G, Lobo J and Ortega J 1998 Phys. Rev. D 57 2051[17] Bender P et al 1995 LISA Pre-Phase A Report unpublish<strong>ed</strong>[18] Poor H V 1994 An Introduction to Signal Detection and Estimation (Berlin: Springer)[19] Vitale S, Cerdonio M, Coccia E and Ortolan A 1997 Phys. Rev. D 55 1741[20] Compton K and Schutz B 1997 Bar-interferometer observing <strong>Gravitational</strong> Waves;Sources and Detectors <strong>ed</strong> I Ciufolini and F Fidecaro (Singapore: World Scientific)p 173[21] Coccia E 1997 Proc. 14th Int. Conf. on General Relativity and Gravitation <strong>ed</strong>M Francaviglia et al (Singapore: World Scientific) p 103Prodi G A et al 1997 <strong>Gravitational</strong> Waves; Sources and Detectors <strong>ed</strong> I Ciufolini andF Fidecaro (Singapore: World Scientific) p 166Cerdonio M et al 1997 Class. Quantum Grav. 14 1491[22] Coccia E, Pizzella G and Ronga F 1994 Proc. 1st Edoardo Amaldi Conf. (Frascati)(Singapore: World Scientific)Bianchi M, Coccia E, Colacino C, Fafone V and Fucito F 1996 Class. Quantum Grav.13 2865Coccia E 1997 <strong>Gravitational</strong> Waves; Sources and Detectors <strong>ed</strong> I Ciufolini andF Fidecaro (Singapore: World Scientific) p 201[23] Johnson W and Merkowitz S 1993 Phys. Rev. Lett. 70 2367[24] Astone P, Lobo J and Schutz B 1994 Class. Quantum Grav. 11 2093[25] Astone P, Pallottino G and Pizzella G 1997 Class. Quantum Grav. 14 2019[26] Astone P, Frasca S, Papa M A and Ricci F 1997 Spectral detection strategy ofstochastic gravitational wave search in VIRGO VIRGO Note VIR-NOT-ROM-1390-106, unpublish<strong>ed</strong>[27] Astone P 1999 Proc. 2nd Edoardo Amaldi Conf. on <strong>Gravitational</strong> Waves <strong>ed</strong> E Cocciaet al (Singapore: World Scientific)[28] Schwartzmann V F 1969 JETP Lett. 9 184[29] Particle Data Group 1996 Review of particle properties Phys. Rev. D 54 66[30] Copi C J, Schramm D N and Turner M S 1997 Phys. Rev. D 55 3389[31] Kirzhnits D A 1972 JETP Lett. 15 745[32] Kibble T W B 1976 J. Phys. A: Math. Gen. 9 1387[33] Vilenkin A and Shellard S 1994 Cosmic Strings and other Topological Defects(Cambridge: Cambridge University Press)[34] Caldwell R R and Allen B 1992 Phys. Rev. D 45 3447[35] Caldwell R R et al 1996 Phys. Rev. D 54 7146[36] Allen B et al 1996 Phys. Rev. Lett. 77 3061[37] Battye R A et al 1997 <strong>Gravitational</strong> <strong>waves</strong> from cosmic strings Topological Defects inCosmology <strong>ed</strong> F Melchiorrei and M Signore (Singapore: World Scientific) pp 11–31 (astro-ph/9706013)[38] Martin X and Vilenkin A 1996 Phys. Rev. Lett. 77 2879Martin X and Vilenkin A 1997 Phys. Rev. D 55 6054


References 241[39] Liddle A R 2000 An introduction to cosmological inflation: high energy physics andcosmology Proc. ICTP Summer School <strong>ed</strong> A Masiero and A Smirnov (Singapore:World Scientific) astro-ph/9901124[40] Turner M S 1997 Ten things everyone should know about inflation Generation ofCosmological Large Scale Structure <strong>ed</strong> P Galeotti and D Schramm (Dordrecht:Kluwer) pp 153–92 (astro-ph/9704062)[41] Guth A H 1981 Phys. Rev. D 23 347[42] Linde A 1982 Phys. Lett. B 108 389[43] Albrecht A and Steinhardt P 1982 Phys. Rev. Lett. 48 1220[44] Linde A 1983 Phys. Lett. B 129 177[45] Birrel N D and Davies P C W 1982 Quantum Field Theory in Curv<strong>ed</strong> Space(Cambridge: Cambridge University Press)[46] Allen B 1988 Phys. Rev. D 37 2078[47] Allen B 1996 Proc. Les Houches School on Astrophysical Sources of <strong>Gravitational</strong>Waves <strong>ed</strong> J Marck and J Lasota (Cambridge: Cambridge University Press)Allen B 1996 Preprint gr-qc/9604033[48] Veneziano G 1991 Phys. Lett. B 265 287Gasperini M and Veneziano G 1993 Astropart. Phys. 1 317Gasperini M and Veneziano G 1993 Mod. Phys. Lett. A 8 3701 (an upto-datecollection of references on string cosmology can be found athttp://www.to.infn.it/∼gasperin/[49] Antoniadis I, Arkani-Ham<strong>ed</strong> N, Dimopoulos S and Dvali G 1998 Phys. Lett. B 429263[50] Gasperini M, Maggiore M and Veneziano G 1997 Nucl. Phys. B 494 315[51] Maggiore M 1998 Nucl. Phys. B 525 413[52] Brustein R and Madden R 1997 Phys. Lett. B 410 110Brustein R and Madden R 1998 Phys. Rev. D 57 712[53] Foffa S, Maggiore M and Sturani R 1999 Phys. Rev. D 59 043507[54] Foffa S, Maggiore M and Sturani R 1999 Nucl. Phys. B 552 395[55] Maggiore M and Riotto A 1999 Nucl. Phys. B 548 427[56] Brustein R, Gasperini M, Giovannini M and Veneziano G 1995 Phys. Lett. B 361 45[57] Buonanno A, Maggiore M and Ungarelli C 1997 Phys. Rev. D 55 3330[58] Maggiore M and Sturani R 1997 Phys. Lett. B 415 335[59] Coleman S 1977 Phys. Rev. D 15 2929Callan C G and Coleman S 1977 Phys. Rev. D 16 1762[60] Turner M S et al 1992 Phys. Rev. D 46 2384See also Hogan C J 1983 Phys. Lett. B 133 172[61] Kamionkowsky M et al 1994 Phys. Rev. D 49 2837[62] Kosowsky A et al 1992 Phys. Rev. Lett. 69 2026[63] Thorne K S 1995 Proc. Snowmass 1994 Summer Study on Particle and NuclearAstrophysics and Cosmology <strong>ed</strong> E Kolb and R Peccei (Singapore: WorldScientific) p 398[64] Ferrari V and Palomba C 1998 Is the gravitational collapse to black holesand interesting source for VIRGO? VIRGO Note VIR-NOT-ROM-1390-109,unpublish<strong>ed</strong>[65] Schutz B F 1997 Low-frequency sources of gravitational <strong>waves</strong>: a tutorial Proc. 1997Alpbach Summer School on Fundamental Physics in Space <strong>ed</strong> A Wilson (ESA)[66] Giazotto A, Bonazzola S and Gourgoulhon E 1997 Phys. Rev. D 55 2014


242 Sources of SGWB[67] Postnov K 1997 Astrophysical sources of stochastic gravitational radiation in theuniverse astro-ph/9706053[68] Ferrari V, Matarrese S and Schneider R 1999 Mon. Not. R. Astron. Soc. 303 258[69] Ferrari V, Matarrese S and Schneider R 1999 Mon. Not. R. Astron. Soc. 303 247[70] Bender P and Hils D 1997 Class. Quantum Grav. 14 1439[71] Bender P et al 1995 LISA Pre-Phase A Report unpublish<strong>ed</strong>


PART 4THEORETICAL DEVELOPMENTSHermann Nicolai, Alessandro Nagar, Donato Bini,Fernando De Felice, Maurizio Gasperini andLuc Blanchet


Chapter 14Infinite-dimensional symmetries in gravityHermann Nicolai 1 and Alessandro Nagar 21 Max-Planck-Institut für Gravitationsphysik(Albert-Einstein-Institut), Mühlenberg 1, D-14476 Golm,Germany2 Dipartimento di Fisica, Università degli Studi di Parma14.1 Einstein theory14.1.1 IntroductionIn these lectures we review the symmetry properties of Einstein’s theory whenit is r<strong>ed</strong>uc<strong>ed</strong> from four to two dimensions. We explain how, in this r<strong>ed</strong>uction,the theory acquires an infinite-dimensional symmetry group, the Geroch group,whose associat<strong>ed</strong> Lie algebra is the affine extension of SL(2, R). The actionof the Geroch group, which is nonlinear and non-local, can be lineariz<strong>ed</strong>, therebypermitting the explicit construction of many solutions of Einstein’s equations withtwo commuting Killing vectors ∂ 2 and ∂ 3 . A non-trivial example of this methodfor a colliding plane wave metric is given.The lectures review some well-known material at a p<strong>ed</strong>agogical level.Therefore, rather than including references in the text, we have chosen to collectsome basic references at the end, which readers are invit<strong>ed</strong> to use as a guide to thevast literature on the subject of exact solutions, on the integrability of Einstein’sequations in the r<strong>ed</strong>uction to two dimensions, and finally on the generalization ofthese structures to other theories, including supergravity.14.1.2 Mathematical conventionsOur main interest is in studying the structural properties of Einstein’s theory andits generalizations. We will first formulate it in D dimensions, with coordinates245


246 Infinite-dimensional symmetries in gravityx M = (x 0 ,...,x D−1 ). The metric can be express<strong>ed</strong> in terms of the vielbein asg MN = E A M E B N η AB (14.1)with the flat metric η AB ≡ (+, −,...,−). For the following it will be importantthat the vielbein can be view<strong>ed</strong> as an element of a coset space according toE A MThe metric must be covariantly conserv<strong>ed</strong>∈ GL(D, R)/SO(1, D − 1). (14.2)D N (Ɣ)g MP = 0 (14.3)where Ɣ is the Christoffel symbol of the metric g MN . We next introduce a spinconnection one-form, with coefficients ω MA B . The vielbein postulate, that is thecovariant constancy of the vielbein, which agrees exactly with Cartan’s structureequation for the torsion two-form, isWriting out this equation, we haveD M (ω, Ɣ)E N A = 0. (14.4)∂ [M E N] A + ω MA B E N B = Ɣ [MN] P E P A . (14.5)We assume there is no torsion, so the Christoffel symbols are symmetric inspacetime indices, henceThe coefficients of the anholonomy are∂ [M E N] A + ω [MA N] = 0. (14.6) AB C = 2E [A M E B] N ∂ M E N C . (14.7)Using the torsion-free condition for the spin connection and permuting the indicesof the coefficients of the anholonomy we obtain the following equations ABC + ω AC B − ω BC A = 0− BC A − ω BAC + ω CAB = 0 (14.8) CAB + ω CBA − ω ABC = 0.Employing then the property of the spin-connection ω ABC =−ω AC B , due tothe fact that the generators of the algebra of the D-Lorentz Group are totallyantisymmetric matrices, we have the expression of the spin connection as afunction of ABCω ABC = 1 2 ( ABC − BC A + CAB ). (14.9)


The Riemann tensor (the curvature two-form) can be defin<strong>ed</strong> byEinstein theory 247[D M (ω), D N (ω)]V A = R MNA B V B . (14.10)The explicit expression in terms of the spin connection isR MNA B = 2∂ [M ω N]A B + 2ω [MA C ω N]C B . (14.11)From the Riemann tensor the Ricci tensor and the curvature scalar are obtain<strong>ed</strong> inthe usual way: R MN := R MPN P and R := g MN R MN . The metric determinant is14.1.3 The Einstein–Hilbert actionE = det E M A = √ −g. (14.12)Now we have all the elements to define Einstein theory. The Einstein–Hilbertaction is∫S = d 4 x Ä (14.13)and the Lagrangian Ä can be express<strong>ed</strong> in function of the spin connectionÄ =− 1 4 ER =−1 4 EE A M E B N R MNAB=− 1 2 EE A M E B N ∂ M ω N AB − 1 4 Eω A AC ω BC B + 1 4 ω BACω AC B . (14.14)Substituting now the expression ω = ω ( ) , integrating by parts and droppingtotal derivatives, we arrive at− 1 4 ER = 16 1 E( ABC ABC − 2 ABC BC A − 4 C AC A D D ). (14.15)This is the expression best suit<strong>ed</strong> for dimensional r<strong>ed</strong>uction of Einstein’s theory.In the remainder, we will now set D = 4, i.e. work in four spacetime dimensions.14.1.4 Dimensional r<strong>ed</strong>uction D = 4 → D = 3‘Dimensional r<strong>ed</strong>uction’ is equivalent to searching for solutions of Einstein’sequations with one Killing vector, which we take to be ξ M ∂ M ≡ ∂ 3 . For thispurpose, we proce<strong>ed</strong> from the ‘Kaluza–Klein ansatz’ for the vierbein.(E A M =−1/2 ea m 1/2 )(B m0 1/2 , E M A =1/2 e m a −e n a B n 1/2 )0 −1/2 .(14.16)The matrix e a m is the three-bein; B m is call<strong>ed</strong> the Kaluza–Klein vector and theKaluza–Klein scalar. The ansatz fixes a part of the SO(1, 3) Lorentz symmetry.The residual symmetry group preserving the gauge condition E a 3 = 0 is thegauge group SO(1, 2).


248 Infinite-dimensional symmetries in gravityAfter some algebra, we find abc = 1 2 ((3)abc − e [a m η b]c −1 ∂ m ) (14.17) 3 ab = 3/2 e m a e n b B mn (14.18) 3 3b =− 1 2 e b m −1/2 ∂ m (14.19) c 3b = 0. (14.20)Substituting the ansatz for the vierbein in the field action and making use of theabove decomposition of the Einstein action, after some calculations we arrive atthe following result− 1 4 ER(E) =−1 4 eR(3) (e) −16 1 e2 B mn B mn + 1 8 egmn −2 ∂ m ∂ n (14.21)where B mn = ∂ m B n − ∂ n B m .Duality transformationThe very special feature of three dimensions is that the Kaluza–Klein vector fieldcan be dualiz<strong>ed</strong> to a scalar. This is achiev<strong>ed</strong> by adding to the Einstein–HilbertLagrangian the expressionÄ ′ = 1 8 e˜ɛmnp B mn ∂ p B (14.22)where B is a Lagrange multiplier and ˜ɛ mnp the Levi-Civita totally antisymmetricsymbol. The dualization makes the Lagrangian depend only on B. So, adding Ä ′to Ä and varying B n leads toe 2 B mn = ɛ mnp ∂ p B (14.23)modulo a numerical constant. Here we have set ɛ mnp = e˜ɛ mnp . Whenwe substitute this expression in the three-dimensional r<strong>ed</strong>uc<strong>ed</strong> Einstein–HilbertLagrangian, we get a new one with two scalar fieldsÄ =− 1 4 eR(3) (e) + 1 8 egmn −2 (∂ m ∂ n + ∂ m B∂ n B). (14.24)This is consistent with the equation of motion ∂ m (e 2 B mn ) = 0. In fact, the termwe add to the Lagrangian, which is now three dimensional, can be dropp<strong>ed</strong> by anintegration by parts and the use of the three-dimensional Bianchi identities for thetensor B mn .14.1.5 Dimensional r<strong>ed</strong>uction D = 3 → D = 2Then we perform a dimensional r<strong>ed</strong>uction from three to two, i.e. we have twoKilling commuting vectors (∂ 3 and ∂ 2 ) and there is no dependence on x 2 at all.x m = (x µ , x 2 ). (14.25)


Einstein theory 249Repeating the same steps as before, the two-bein now takes the form( ) (e a eµαρ Am =µ, e m eαµ−e0 ρ a =α σ A )σ0 ρ −1 . (14.26)Detail<strong>ed</strong> calculation shows thatwith− 1 4 e(3) R (3) =− 1 4 ρeR(2) − 116 ρ3 eA µν A µν (14.27)eR (2) =−2∂ µ (ee αµ αβ β ) (14.28)where e is the determinant of the two-bein. At this point we can write theequations of motion for the theory. The equation of motion for the Kaluza–Kleinvector is given by∂ µ (ρ 3 eA µν ) = 0. (14.29)In two dimensions, a Maxwell field does not propagate, as there are notransverse degrees of fre<strong>ed</strong>om. Neglecting topological effects (i.e. non-vanishingholonomies) we can, therefore, set A µ = 0.For the remaining equations of motion, we can fix the gauge, and thencalculate them in a particular gauge, call<strong>ed</strong> the conformal gauge. The termeR (2) (e) is Weyl-invariant. To see why this is so, let us consider the termAn integration by parts gives− 1 4 ρ R(2) = 1 2 ρ∂ ν(ee α ν αγ γ ). (14.30)− 1 4 ρ R(2) ˙=− 1 2 eαγ γ e α µ ∂ µ ρ. (14.31)Then, using the definition of the anholonomy, we get= − 1 2 e(e α ν e τ γ ∂ ν e γ τ − e ν γ e τ α ∂ ν e γ τ )e αµ ∂ µ ρ (14.32)= − 1 2 egµν e τ γ ∂ ν e γ τ ∂ µ ρ − 1 2 e∂ νe ν α e αµ ∂ µ ρ (14.33)where another integration by parts and the definition of the two-bein have beenus<strong>ed</strong>.Now we can set the gauge, i.e. the 2D diffeomorphisms, by a condition onthe two-bein. So we writee α αµ = λẽ µ (14.34)with det ẽ µ α = 1 and λ = λ(x); hence, we are not considering the whole groupGL(2, R) but only its restriction to unimodular matrices SL(2, R).As we said before, we can set a particular gauge, the conformal gauge, byimposing the following condition on the two-bein. It is given byẽ µ α = δ α µ . (14.35)


250 Infinite-dimensional symmetries in gravityThen, after an integration by parts, we getIn this gauge it is obviously− 1 4 ρ R(2) ˙=− ˜g µν λ −1 ∂ ν λ∂ µ ρ + 1 2 ẽα ν ∂ ν (ẽ αµ ∂ µ ρ). (14.36)˜g µν =ẽ αµẽ να = λ 2 g µν . (14.37)So we have three fields: the dilaton ρ, the λ and the unimodular two-bein ẽ µ α .We can calculate the equations of motion varying the Lagrangian with respect toall these fields. Varying it with respect to λ we get∂ µ ( ˜g µν ∂ ν ρ) ≡ £ρ = 0 (14.38)because in conformal gauge ˜g µν = η µν . The solution of this equation isρ(x) = ρ + (x + ) + ρ − (x − ) (14.39)with x ± = x 0 ± x 1 .The dilaton can be dualiz<strong>ed</strong>: in two dimensions, the dual of a scalar field isagain a scalar field. We will refer to the dual of the dilaton field as the ‘axion’;itis defin<strong>ed</strong> by∂ µ ρ + ɛ µν ∂ ν ˜ρ = 0 (14.40)where ˜ρ is just the axion. In the conformal gauge this field isThe equation obtain<strong>ed</strong> by varying ρ is˜ρ(x) = ρ + (x + ) − ρ − (x − ). (14.41)∂ µ ( ˜g µν λ −1 ∂ ν λ) = matter contribution. (14.42)Note that with matter contribution we refer to the fields and B coming out fromdimensional r<strong>ed</strong>uction. The terminology matter part will be clear in the followingsection, where we will be able to identify this fields with the fields of a bosonicnonlinear σ -model Lagrangian.Before writing the complete Lagrangian of the two-dimensional r<strong>ed</strong>uc<strong>ed</strong>gravity we must still consider the equation that is obtain<strong>ed</strong> from (14.36) byvariation with respect to the unimodular two-bein. The corresponding equationsmust be interpret<strong>ed</strong> as constraint equations (in standard conformal field theory,they would just correspond to the Virasoro constraints). We have−δẽ αµẽ αν λ −1 ∂ (µ λ∂ ν) ρ + 1 2 δẽ α µ ∂ µ (ẽ αν ∂ ν ρ) − 1 2 ∂ µẽ α µ δẽ αν ∂ ν ρ + matter = 0.(14.43)In conformal gauge, ẽ α µ = δ µ α , this expression becomes− 1 2 δ ˜gµν λ −1 ∂ µ λ∂ ν ρ + 1 4 δ ˜gµν ∂ µ ∂ ν ρ = matter (14.44)


Einstein theory 251where δ ˜g µν = 2δẽ αµẽ αν has been us<strong>ed</strong>. The metric is diagonal in this gauge, sothe equations for the conformal factors become(traceless part of){− 1 2 λ−1 ∂ µ λ∂ ν ρ + 1 4 ∂ µ∂ ν ρ}=matter. (14.45)We have not explicitly written the matter sector yet. This will be done in the nextsection.The pure gravity action written in two dimensions in the conformal gaugereads− 1 4 e(3) R (3) =− 1 2 λ−1 ∂ µ λ∂ µ ρ (14.46)where the equation £ρ = 0 has been us<strong>ed</strong> to make the second term of (14.36)vanish. The whole Lagrangian is thenÄ E =− 1 2 λ−1 ∂ µ λ∂ µ ρ + 1 8 ρ−2 (∂ µ ∂ µ + ∂ µ B∂ µ B) (14.47)where the second term with the Kaluza–Klein scalar and the dual of the Kaluza–Klein vector has been obtain<strong>ed</strong> considering only the two-dimensional part of theaction. The subscript E stands for Ehlers, who did this analysis for the first timein the 1950s.Here we have treat<strong>ed</strong> one possible way of performing dimensional r<strong>ed</strong>uction:we have seen it consists of many steps. One first r<strong>ed</strong>uces from four to three; then,dualizes the vector field and performs the r<strong>ed</strong>uction to two dimensions. However,this is not the whole story: actually, it is possible also to get the two-dimensionalLagrangian directly from the three-dimensional one, without dualization.The proc<strong>ed</strong>ure for doing the calculation is as follows: first we express theKaluza–Klein vector in the formB m = (B µ , B 2 ≡ ˜B) (14.48)and then we perform directly the dimensional r<strong>ed</strong>uction in conformal gauge byusing the previous choice of the three-bein in triangular form. Proce<strong>ed</strong>ing inthis way, we meet two electromagnetic fields in two dimensions, A µ and B µ ,which can be set to zero because they do not propagate (we have already us<strong>ed</strong> thisargument before) and there is no cosmological constant.Let us note now the various steps of the calculation. The newwriting for the Kaluza–Klein vector and the considerations on two-dimensionalelectromagnetism implyB µ = 0 → B µν = 0. (14.49)So the only non-vanishing terms of the three-dimensional Lagrangian before ther<strong>ed</strong>uction areÄ =− 1 4 e(3) R (3) (e) − 1 8 e(3) 2 B µ2 B µ2 + 1 8 e(3) g mn −2 ∂ m ∂ n . (14.50)Then we r<strong>ed</strong>uce the dimensions as before (we set A µ = 0) and choose theconformal gauge. Keeping in mind thatg µν = λ 2 η µν , g 22 =−ρ 2 (14.51)


252 Infinite-dimensional symmetries in gravitywe get a new 2D LagrangianÄ MM =− 1 2 λ−1 ∂ µ λ∂ µ ρ + 1 8 ρ−2 ∂ µ ∂ µ + 1 8 ρ−1 2 ∂ µ ˜B∂ µ ˜B (14.52)for the fields ˜B and .The subscript MM stands for Matzner–Misner, who perform<strong>ed</strong> this analysisat the end of the 1960s. It is worth noting that the link between the fields B and˜B is given by three-dimensional duality. To see why it is so, let us consider th<strong>ed</strong>uality relatione 2 B mn = ɛ mnp ∂ p B (14.53)and then r<strong>ed</strong>uce the dimensionality using the properties of the Kaluza–Kleinvector B m . The duality relation becomes thenρ −1 2 ∂ µ ˜B = ɛ µν ∂ ν B. (14.54)The deep relation between these two distinct r<strong>ed</strong>uct<strong>ed</strong> Lagrangians will beexplain<strong>ed</strong> in the next section, treating nonlinear σ -models. In the language of thenonlinear sigma models, these two r<strong>ed</strong>uct<strong>ed</strong> actions correspond to two distinctSL(2, R)/SO(2) models.14.2 Nonlinear σ -modelsIn this section, we introduce nonlinear σ -models and discover that the r<strong>ed</strong>uc<strong>ed</strong>gravity is a certain nonlinear σ -model, connect<strong>ed</strong> to a certain symmetry group.The expression of the Lagrangian of the model depends on this symmetry group.Let us start from a non-compact Lie group G and consider the maximalcompact Lie subgroup H of G. The Lie algebra decomposition iswith the following commutation rulesG = H ⊕ K (14.55)[H, H] ⊂ H, [K , K ] ⊂ H, [ H, K ] ⊂ K . (14.56)This decomposition is invariant under the symmetric space automorphismτ(H) = H, τ(K ) =−K (14.57)which can alternatively be formulat<strong>ed</strong> in terms of Lie group elements g directlythroughτ(g) = η −1 (g T ) −1 η (14.58)where the matrix η depends on the group G (e.g. η = 1 for G = SL(n, R)).


Nonlinear σ -models 253Example: G = SL(2, R), H= SO(2)The generators of the group are(Y 1 1 0=0 −1), Y 2 =( )0 1, Y 3 =1 0( )0 1. (14.59)−1 0We haveTr(Y 1 ) 2 = Tr(Y 2 ) 2 =−Tr(Y 3 ) 2 = 2 (14.60)so Y 1 and Y 2 are the non-compact generators, while Y 3 generates the SO(2)subgroup.The group can be decompos<strong>ed</strong> on its generators, asH = RY 3 , K = RY 1 ⊕ RY 2 . (14.61)Let us introduce now an element of the group v(x) ∈ G with the propertyv(x) → v ′ (x) = g −1 v(x)h(x) (14.62)where g is a rigid G transformation and h(x) a local H transformation. This typeof transformation is ne<strong>ed</strong><strong>ed</strong> for the necessity of preserving gauge choice. In fact,you can fix the gauge choosing a particular element of the group v. Then, whenyou act on v by an arbitrary g, that gauge choice will be lost. To restore the gaugeyou have to introduce the local transformation h(x) so that the rotation g can becompens<strong>ed</strong>. It follows that h does not depend only on the coordinates x, but alsoon the vector v and the rotation g.Therefore, equation (14.62) is call<strong>ed</strong> a nonlinear realization of symmetries,because h depends nonlinearly on v.This is important for the following calculation, because we can fix a gauge,call<strong>ed</strong> triangular gauge, such thatv(x) = exp ϕ(x), ϕ(x) ∈ K → v ∈ G/H. (14.63)The next step is the construction of a Lagrangian with the requir<strong>ed</strong> symmetry. Tothis aim, let us consider the Lie algebra valu<strong>ed</strong> expressionv −1 ∂ m v = Q m + P m , Q m ∈ H, P m ∈ K (14.64)or equivalentlyv −1 D m v = v −1 (∂ m v − vQ m ) = P m (14.65)which defines the H -covariant derivative D m . It is straightforward to verify thatQ m transforms like a gauge field with respect to the local group H , namelyQ ′ m = h−1 Q m h + h −1 ∂ m h and that P m ′ = h−1 P m h. The formula (14.62) impliesthe integrability relations∂ m Q m − ∂ n Q m + [Q m , Q n ] = −[P m , P n ]D m P n − D n P m = 0.


254 Infinite-dimensional symmetries in gravityThe Lagrangian is given byÄ = 1 4 egmn Tr P m P n (14.66)and then the field equations for P m readD m ( √ gg mn P n ) = 0. (14.67)In the following section we will show how it is possible to reproduce theLagrangians obtain<strong>ed</strong> by dimensional r<strong>ed</strong>uction from this general constructionof nonlinear σ -models.14.2.1 Ehlers Lagrangian as a nonlinear σ -modelTo link these arguments to the previous discussion, let us consider the groupsG = SL(2, R), H = SO(2). (14.68)The quotient space has only two degrees of fre<strong>ed</strong>om. We enforce the triangulargauge choosing for v the following expression(1/2Bv =−1/2 )0 −1/2 (14.69)and then( 12v −1 −1 ∂ m −1 )∂ m B∂ m v =0 − 1 2 −1 ∂ m = Pm 1 Y 1 + Pm 2 Y 2 + Q m Y 3 (14.70)where the coefficients of the generators of the algebra are given byPm 1 = 1 2 −1 ∂ m , Pm 2 = Q m = 1 2 −1 ∂ m B. (14.71)The evaluation of the Lagrangian is straightforward, and we getÄ = 1 4 egmn Tr P m P n = 1 8 egmn −2 (∂ m ∂ n + ∂ m B∂ n B). (14.72)This result matches exactly with the matter part of the Einstein–HilbertLagrangian found in the previous section. We have found that this expression canbe directly r<strong>ed</strong>uc<strong>ed</strong> to two dimensions, and then, coupl<strong>ed</strong> to gravity, it becomessimply the Ehlers Lagrangian Ä E seen before.The Ehlers Lagrangian after dimensional r<strong>ed</strong>uction isÄ E = gravity + 1 4 ρe(2) g µν Tr(P µ P ν )=− 1 2 λ−1 ∂ µ λ∂ µ ρ + 1 8 ρe(2) −2 g µν (∂ µ ∂ ν + ∂ µ B∂ ν B). (14.73)We saw in the previous section that, by another type of dimensional r<strong>ed</strong>uction, wegot a different r<strong>ed</strong>uc<strong>ed</strong> Lagrangian, the Matzner–Misner one.This one can be construct<strong>ed</strong> as a nonlinear σ -model too: we ne<strong>ed</strong> only adifferent gauge choice, as we will see in the next section; before this, let us lookat the equations of motion deriv<strong>ed</strong> from the Ehlers Lagrangian.


Nonlinear σ -models 25514.2.2 The Ernst equationThe equations of motion for the fields and B from the Lagrangian Ä E are∂ µ (ρ∂ µ ) = ρ(∂ µ ∂ µ − ∂ µ B∂ µ B) (14.74)∂ µ (ρ∂ µ B) = 2ρ∂ µ ∂ µ B. (14.75)Defining a complex function = + iB call<strong>ed</strong> the Ernst potential, theseequations can be combin<strong>ed</strong> into a single one, call<strong>ed</strong> the ‘Ernst equation’:∂ µ (ρ∂ µ ) = ρ∂ µ ∂ µ . (14.76)This equation figures prominently in studies of exact solutions of Einstein’sequations.Here we have got the Ernst equation from the fields equations for and B.Actually equation (14.67) is the Ernst equation, in the sense that it r<strong>ed</strong>uces to itchoosing the Ehlers triangular form for v in the conformal gauge.14.2.3 The Matzner–Misner Lagrangian as a nonlinear σ -modelRecalling the shape of the Matzner–Misner Lagrangian as written before in theconformal gaugeÄ MM =− 1 2 λ−1 ∂ µ λ∂ µ ρ + 1 8 ρ−2 ∂ µ ∂ µ + 1 8 ρ−1 2 ∂ µ B 2 ∂ µ B 2 . (14.77)This can be thought of as a nonlinear σ -model, too. We suppose our Lagrangianto be compos<strong>ed</strong> by a term of pure gravity, but r<strong>ed</strong>uc<strong>ed</strong> to two dimensions, and aterm coming from a two-dimensional nonlinear σ -model. We are in conformalgauge, namely e µ α = λδ µ α and ˜g µν = η µν . The Lagrangian isÄ =− 1 2 ˜λ −1 ∂ µ ˜λ∂ µ ρ + 1 4 ρηµν Tr ˜P µ ˜P ν . (14.78)We have to choose a proper gauge, namely an expression for v, such that the twoLagrangians match together.We refer now to the generators of SL(2, R) introduc<strong>ed</strong> at the beginning ofthis section. Let us choose for ṽ the following triangular formṽ =((ρ/)1/2B 2 (/ρ) 1/2 )0 (/ρ) 1/2 , ṽ −1 =((/ρ)1/2−B 2 (/ρ) 1/2 )0 (ρ/) 1/2 .Evaluating now the matrix product ṽ −1 ∂ µ ṽ and decomposing it on the algebragenerators. Following the standard proc<strong>ed</strong>ure seen before, the Lagrangian is builtusing only the non-compact elements of this decomposition. After calculation,we haveṽ −1 ∂ µ ṽ = ˜P µ 1 + ˜P µ 2 + ˜Q µ = 1 2 (ρ−1 ∂ µ ρ − −1 ∂ µ )+ 1 ( ) ( ) 0 1∂ µ B 2 + 1 ( ) 2 ρ 1 0 2 ρ(1 0)0 −1∂ µ B 2(0 1−1 0). (14.79)


256 Infinite-dimensional symmetries in gravityThen, the trace is14 Tr ˜P µ ˜P µ = 1 4 ( ˜P µ 1 ˜P µ1 + ˜P µ 2 ˜P µ2 ) = 1 8 ρ−1 ∂ µ ρ(ρ −1 ∂ µ ρ − 2 −1 ∂ µ )) 2∂ µ B 2 ∂ µ B 2 . (14.80)+ 1 8 −2 ∂ µ ∂ µ + 1 ( 8 ρNow, the two Lagrangians coincide if ˜λ satisfies the condition− 1 2 ˜λ −1 ∂ µ ˜λ∂ µ ρ + 1 8 ρ−2 ∂ µ ρ(ρ −1 ∂ µ ρ − 2 −1 ∂ µ ) =− 1 2 λ−1 ∂ µ λ∂ µ ρ (14.81)namely if˜λ ≡ λρ 1/4 −1/2 . (14.82)Therefore, the two-dimensional r<strong>ed</strong>uc<strong>ed</strong> gravity in conformal gauge is given by apart of pure two-dimensional gravity, characteriz<strong>ed</strong> by the conformal factor λ andthe dilaton ρ, and a matter part given by the bosonic fields and B, or ˜B: thisone has the structure of a nonlinear G/H sigma model.Following the first section of this paper, the complete Lagrangian r<strong>ed</strong>uc<strong>ed</strong> totwo dimensions in conformal gauge, for any G/H σ -model isÄ =− 1 2 λ−1 ∂ µ λ∂ µ ρ + 1 4 ρ Tr(P µ P µ ) (14.83)and we can recover, as before, the field equation for the conformal factor λ, thistime with the general σ -model matter part. It is given by the traceless part ofλ −1 ∂ µ λ∂ ν ρ = 1 2 Tr(P µ P ν ) + 1 2 ∂ µ∂ ν ρ. (14.84)This will be useful in the foregoing sections when recovering the colliding planewave solutions of Einstein’s theory.The Kramer–Neugebauer transformationNote now that the two models, that of Ehlers and that of Matzner–Misner, arerelat<strong>ed</strong> by the Kramer–Neugebauer transformation, defin<strong>ed</strong> by ↔ ρ , B ↔ B 2It is worth remembering that the fields B and B 2 are relat<strong>ed</strong> by duality too, namelyɛ µν ∂ ν B = 2ρ ∂ µ B 2 . (14.85)To sum up: in this section we have seen that the dimensional r<strong>ed</strong>uction of Einsteintheory from D = 4toD = 2 can be done in two ways, leading to two differentSL(2, R)/SO(2) σ-models.We discover two different isometry groups, that of Ehlers and that of Matzer–MisnerSL(2, R) E , SL(2, R) MM . (14.86)Combining these two groups, one gets the (infinite-dimensional) Geroch group.


14.3 Symmetries of nonlinear σ -modelsSymmetries of nonlinear σ -models 257We have seen that for preserving the gauge choice, in particular the triangulargauge, the symmetryv → v ′ = g −1 v, g ∈ G (14.87)must be realiz<strong>ed</strong> in a nonlinear way, namelyv → v ′ (x) = g −1 v(x)h(x), g ∈ G, h ∈ H. (14.88)Now consider the infinitesimal form of (14.88). The infinitesimal variation of v isδv(x) =−δg −1 v(x) + v(x)δh(x) (14.89)applying now this lineariz<strong>ed</strong> transformation to the two σ -models seen before.Considering in particular the Chevalley–Serre generators for the SL(2, R)Lie algebrae ≡ T + =( )0 1, f ≡ T − =0 0endow<strong>ed</strong> with the following commutation rules( )(0 0, h ≡ T 3 1 0=1 00 −1)(14.90)[h, e] = 2e, [h, f ] =−2 f, [e, f ] = h (14.91)one can check that this nonlinear transformation has been introduc<strong>ed</strong> to preservethe gauge. Let us now analyse the action of the Ehlers and Matzner–Misnergroups in turn.14.3.1 Nonlinear realization of SL(2, R) EWe use the Chevalley–Serre generators for the algebra. Considering the triangulargauge(1/2Bv =−1/2 )0 −1/2 (14.92)we now linearize the transformation (14.87). The variation of v is only due to thealgebra element a:δv = v ′ − v =−av. (14.93)Then, given the triangular form of v, it follows also( 12 −1/2 δ − 1δv =2 −3/2 Bδ + −1/2 )δB0 − 1 . (14.94)2 −3/2 δIn the following we will refer to the variation δ by, for example, the generator ewith the compact notation e() or e(B) for B. Now, we realize the transformationusing the Chevalley–Serre algebra generators, e, h and f .Fore we have( )(0 1 1/2Be 1 : −−1/2 ) ( )0 −−1/20 0 0 −1/2 =(14.95)0 0


258 Infinite-dimensional symmetries in gravitywhere the subscript 1 refers to the Ehlers group. From (14.94), one d<strong>ed</strong>ucese 1 () = 0, e 1 (B) =−1 (14.96)and the triangular gauge is preserv<strong>ed</strong>. The calculation is analogous for thegenerator h( )(1 0 1/2Bh 1 : −−1/2 ) (−1/2−B0 −1 0 −1/2 =−1/2 )0 −1/2 (14.97)with h 1 () =−2 and h 1 (B) =−2B. The triangular gauge is still preserv<strong>ed</strong>.This is not so for the third generator, f . Repeating the above steps we find( )(0 0 1/2Bf 1 : −−1/2 ) ( )0 01 0 0 −1/2 =− 1/2 −B −1/2 (14.98)namely the triangular gauge is not preserv<strong>ed</strong>. Therefore, we have to introduce acompensating term, i.e. we ne<strong>ed</strong> the transformation rule (14.89). We introducea local H transformation parametriz<strong>ed</strong> by a function ω, which is determin<strong>ed</strong> insuch a way as to preserve the gauge. Remember that the H generator is Y 3 :( ) (f 1 : − f 1 v + v(−ωY 3 0 0B−1/2−) =− 1/2 −B −1/2 + ω1/2 ) −1/2 .0(14.99)The triangular gauge is defin<strong>ed</strong> by the condition− √ +and so the transformation readsf 1 : δv =Hence the variations of the fields and B areω √= 0 → ω = (14.100)(B1/2− 3/2 )0 −B −1/2 . (14.101)f 1 () = 2B, f 1 (B) = B 2 − 2 (14.102)clearly not linear in the fields.Note that the SL(2, R) transformations leave the fields ρ and λ unchang<strong>ed</strong>,i.e.δλ = 0, δρ = 0.14.3.2 Nonlinear realization of SL(2, R) MMOn the other side, identical calculations can be done to evaluate the action ofSL(2, R) MM on the fields (, B 2 ). Also in this case the symmetry is realiz<strong>ed</strong> in


The Geroch group 259a nonlinear way. We have (with the suffix 0 for Matzner–Misner)e 0 () = 0, e 0 (B 2 ) =−1 (14.103)h 0 () = 2, h 0 (B 2 ) =−2B 2 (14.104)( ρ 2f 0 () =−2B 2 , f 0 (B 2 ) = B2 ) 2 − . (14.105)Again, the generator f 0 acts nonlinearly.14.4 The Geroch groupThe aim of this section is to combine the two groups, SL(2, R) E with fields(, B) and SL(2, R) MM , with (, B 2 ), into a unifi<strong>ed</strong> group, the infinit<strong>ed</strong>imensionalGeroch group. The associat<strong>ed</strong> Lie algebra is an affine Kac–Moodyalgebra.We return first to duality relationρ −1 2 ∂ µ B 2 = ɛ µν ∂ ν B (14.106)which is invariant under the Kramer–Neugebauer transformation. We ne<strong>ed</strong> thisequation because we now have to evaluate the action of SL(2) E on B 2 and ofSL(2) MM on B.14.4.1 Action of SL(2, R) E on ˜λ, B 2Keeping in mind that δρ = 0, we haveB → B + δB ⇒ ɛ µν ∂ ν (δB) = δ( 2 ρ −1 ∂ µ B 2 ) (14.107)after the functional differentiation and the usage of duality∂ µ (δB 2 ) = ρɛ µν (∂ ν δB − 2 −3 δ). (14.108)Consequently, from the change of B calculat<strong>ed</strong> before, we have the variation ofB 2 due to the SL(2) E generators.e 1 :0= ∂ µ (δB 2 ) ⇒ e 1 (B 2 ) = c 1 (= constant) (14.109)h 1 : ∂ µ (δB 2 ) = 2ρ −2 ɛ µν ∂ ν B ⇒ h 1 (B 2 ) = 2B 2 (14.110)f 1 : ɛ µν ∂ ν (δB 2 ) = 2ρ( −2 B∂ ν B + −1 ∂ ν ) ⇒ f 1 (B 2 ) = 2φ 1 . (14.111)Here a dual potential φ 1 has been introduc<strong>ed</strong>, which is defin<strong>ed</strong> such thatρ −1 ɛ µν ∂ ν φ 1 = −2 (B∂ µ B + ∂ µ ). (14.112)Careful inspection of these relations now shows the following. The contributionsdue to e 1 and h 1 are linear in the fields and local; the difference is in the


260 Infinite-dimensional symmetries in gravitytransformation generat<strong>ed</strong> by f 1 , which is clearly nonlinear and non-local, becauseone has to perform an integration to calculate explicitly the dual potential.We then evaluate also the action on ˜λ. From the definition of (14.82), andobserving that δλ = 0, it follows that˜λ −1 δ ˜λ =− 1 2 −1 δ (14.113)14.4.2 Action of SL(2, R) MM on λ, BExactly the same analysis has to be done for the other group, with generators(e 0 , h 0 , f 0 )e 0 :⇒ e 0 (B) = c 0 (14.114)h 0 :⇒ h 0 (B) = 2B (14.115)f 0 :⇒ f 0 (B) = 2φ 0 (14.116)withρɛ µν ∂ ν φ 0 =− 2 B 2 ∂ µ B 2 + ρ∂ µ( ρ). (14.117)14.4.3 The affine Kac–Moody SL(2, R) algebraThe transformations we have just deriv<strong>ed</strong> are to be identifi<strong>ed</strong> with an affineSL(2, R) Kac–Moody algebra. The latter is characteriz<strong>ed</strong> by the Cartan matrix( )2 −2A ij =(14.118)−2 2and the standard Chevalley–Serre presentation defining the algebra which can beread off from the Cartan matrix:[h i , h j ] = 0[h i , e j ] = A ij e j[h i , f j ] =−A ij f j[e i , f j ] = δ ij h j[e i [e i [e i , e j ]]] = 0[ f i [ f i [ f i , f j ]]] = 0.Here i = j = 0, 1; note that there is no summation on repeat<strong>ed</strong> indices andthat the first relation defines the Cartan subalgebra. To see the relation with theSL(2, R) transformations dealt with before, we make the identificationse 1 = T0 + , f 1 = T0 − , h 1 = T0 3 (14.119)e 0 = T − 1 , f 0 = T − + , h 0 = c − T0 3 . (14.120)


The linear system 261Using the known commutation rules of the Chevalley–Serre generators ofSL(2, R) it is possible to directly check the algebra.For example, it is straightforward to see that[h 1 , e 0 ] = [T 3 0 , T −1 ] =−2T − 1 =−2e 0 (14.121)or that[e 1 , e 0 ] = [T0 + , T 1 − ] = 2T 13 (14.122)[e 1 , [e 1 , e 0 ]] =−4T1 + ⇒ [e 1[e 1 [e 1 , e 0 ]]] = 0and so on for the other commutators.The full current algebra is now built by taking multiple commutators in allpossible ways. The Lie algebra element c = h 0 + h 1 is the central charge. It hasa trivial action on the fields , B, B 2 , ˜λ, λ.14.5 The linear systemThe aim of this section is to linearize and localize the action of the Geroch groupseen in the previous section.Let us start from the Lagrangian for arbitrary G/H in three dimensions andthen r<strong>ed</strong>uce to twoÄ =− 1 4 ρeR(e) + 1 4 ρegmn Tr P m P n . (14.123)We pick now the conformal gauge for the three-bein, as before( )em a λδ = αµ 00 ρ(14.124)where we have dropp<strong>ed</strong> the two-dimensional Kaluza–Klein vector because itcarries no physical degrees of fre<strong>ed</strong>om any more. It is well known that the choiceeµ α = λeα µ is preserv<strong>ed</strong> under conformal diffeomorphismsδx + = ξ − (x + ), δx − = ξ + (x − ) (14.125)with the light cone coordinates x ± ≡ 1 √2(x 0 ± x 1 ). This residual coordinatefre<strong>ed</strong>om can be gaug<strong>ed</strong> away, for example, by employing the dilaton and theaxion fields. One can fix the residual conformal diffeomorphisms by identifyingthe field ρ or ˜ρ with one of the coordinates.14.5.1 Solving Einstein’s equationsLet us now focus our attention on the way of solving Einstein’s equation. Firstnote that by substituting the gauge (14.124) into the scalar equation (14.67), wearrive atρ −1 D µ (ρ P µ ) = 0. (14.126)


262 Infinite-dimensional symmetries in gravityThe dependence of this equation on ρ is all that remains of three-dimensionalgravity. Equation (14.126) r<strong>ed</strong>uces to the Ernst equation for G = SL(2, R), butwe will return to this later. For the moment, note only that this equation works forthe σ -model degrees of fre<strong>ed</strong>om, namely on .The remaining equations, which follow from higher dimensions, are theequations for the dilaton ρ and for the conformal factor λ. ρ is a free field intwo dimensions which can be solv<strong>ed</strong> for in terms of two arbitrary functions (leftmoversand right-movers)£ρ = 0 ⇒ ρ(x) = ρ + (x + ) + ρ − (x − ). (14.127)The equations of motion for the conformal factor in light-cone coordinatesρ −1 ∂ ± ρλ −1 ∂ ± λ = 1 2 Tr(P ± P ± ) + 1 2 ρ−1 ∂ 2 ± ρ (14.128)can be written as∂ ± ρ∂ ± ˆσ = 1 2 ρ Tr P ± P ± (14.129)where the second term on the right-hand side of (14.128) has been reabsorb<strong>ed</strong> intothe Liouville scalar ˆσ = λ(∂ + ρ) − 1 2 (∂ − ρ) − 1 2 . Note that this equation determinesλ only up to a constant factor. Observe also that this equation has no analoguein flat space theories, and this, together with the presence of ρ, makes a greatdifference. For instance, we cannot simply put ρ = constant, for this wouldimply the vanishing of the right-hand side of (14.129), which by the positivity ofthe Killing metric on the subalgebra K would imply P ± = 0 and leave us onlywith the trivial solution v = constant (modulo H gauge transformations).Now specializing to general relativity, i.e. G/H = SL(2, R)/SO(2) cosetspace. As anticipat<strong>ed</strong> before, we start from the equation of motion (14.126),employ the triangular gauge in the Ehlers form, so to have explicit expressionsfor P µ and Q ν . After a little algebra we have again the Ernst equation∂ µ (ρ∂ µ ) = ρ∂ µ ∂ µ (14.130)in terms of the complex potential = + iB. Solving Einstein’s equationsis now simply a matter of choosing the appropriate ρ(x), finding a solution ofthe nonlinear partial differential equation (14.130) and finally determining theconformal factor λ by integration of (14.129). For the colliding plane <strong>waves</strong>olutions, that will be recover<strong>ed</strong> in the next sections, one distinguishes <strong>waves</strong> withcollinear polarization, where B = 0 and <strong>waves</strong> with non-collinear polarization.For collinearly polariz<strong>ed</strong> <strong>waves</strong>, the nonlinear Ernst equation can be r<strong>ed</strong>uc<strong>ed</strong> to alinear partial differential equation through the replacement = exp ψ.So, for collinearly polariz<strong>ed</strong> <strong>waves</strong>, with B = 0, the four-bein is⎛λ −1/2 ⎞0 0 0E A ⎜ 0 λM =−1/2 0 0 ⎟⎝0 0 ρ −1/2 ⎠ (14.131)00 0 0 1/2


The linear system 263and then the four dimensional line element isds 2 = 2 −1 λ 2 dx + dx − − −1 ρ 2 (dx 2 ) 2 − (dx 3 ) 2 (14.132)where , ρ and λ depend only on x + and x − .14.5.2 The linear systemThe integrability of the nonlinear equation of motion (14.126) is reflect<strong>ed</strong> in theexistence of a linear system. This means that there is a set of linear differentialequations, whose compatibility conditions yield just the nonlinear equations thatone tries to solve.To formulate the linear system one must introduce a so-call<strong>ed</strong> spectralparameter t as an extra variable and replace v(x) by a matrix ˆv(x) which alsodepends on t.v(x 0 , x 1 ) →ˆv(x 0 , x 1 ; t). (14.133)We postulateˆv −1 ∂ µ ˆv = Q µ + 1 + t21 − t 2 P µ + 2t1 − t 2 ɛ µν P ν . (14.134)This is a generalization of v −1 ∂ µ v = Q µ + P µ , which is obtain<strong>ed</strong> from (14.134)in the case t = 0. (14.134) is equivalent toˆv∂ ± ˆv = Q ± + 1 ∓ t1 ± t P ±. (14.135)Here we have an integrability condition, written as∂ + ( ˆv −1 ∂ − ˆv) − ∂ − ( ˆv −1 ∂ + ˆv) + [ ˆv −1 ∂ + ˆv, ˆv −1 ∂ − ˆv] = 0 (14.136)which using (14.135) can be directly check<strong>ed</strong> by calculation, making use of theintegrability condition seen before and of the equation of motion for ρ. W<strong>ed</strong>efineexplicitlyEmploying (14.135) these relations become = ∂ + ( ˆv −1 ∂ − ˆv) − ∂ − ( ˆv −1 ∂ + ˆv) (14.137) = [ ˆv −1 ∂ + ˆv, ˆv −1 ∂ − ˆv]. (14.138) = ∂ + Q − − ∂ − Q + + 1 + t1 − t ∂ + P − − 1 − t1 + t ∂ − P +( ) ( )1 + t1 − t+ ∂ + P − − ∂ − P + (14.139)1 − t1 + t = [Q + , Q − ] + [P + , P − ] + 1 + t1 − t [Q +, P − ] − 1 − t1 + t [Q −, P + ]. (14.140)


264 Infinite-dimensional symmetries in gravityThe sum now reads + = 1 + t1 − t D + P − − 1 − t1 + t D − P ++ 2t(1 − t) 2 t−1 ∂ + tP − + 2t(1 + t) 2 t−1 ∂ − tP + (14.141)where the integrability relation seen in section 14.2 has been us<strong>ed</strong>. Now let uspostulatet −1 ∂ ± t = 1 ∓ t1 ± t ρ−1 ∂ ± ρ (14.142)so it follows + = 1 + t21 − t 2 (D + P − − D − P + ) + 2t1 − t 2 (D + P − + D − P + )+ 2t1 − t 2 (ρ−1 ∂ + ρ P − + ρ −1 ∂ − ρ P + ). (14.143)Now the first term is null for the integrability relation D + P − = D − P + andthe second for the equation (14.126). Therefore, the integrability condition ischeck<strong>ed</strong>.Let us now focus on equation (14.142): it is integrable once one has asolution of £ρ = 0. This can be explicitly verifi<strong>ed</strong>, as it follows. First, let usmultiply (14.142) by (1 − t 2 ); after a little algebra this equation r<strong>ed</strong>uces to(∂ ±[ρ t + 1 ) ]− 2 ˜ρ = 0 (14.144)twhere the axion ˜ρ has been introduc<strong>ed</strong>. So one must have12 ρ (t + 1 t)−˜ρ = w (14.145)where w is an integration constant. When we substitute in this relation the explicitexpression of the dilaton and the axion as functions of incoming and outgoingfields, we get√w + ρ+ (xt(x; w) =+ ) − √ w − ρ − (x − )√w + ρ+ (x + ) + √ w − ρ − (x − ) . (14.146)For fix<strong>ed</strong> x, the function t(x; w) lives on a two-sheet<strong>ed</strong> Riemann surface over thecomplex w-plane, with an x-dependent cut extending from ρ − (x − ) to ρ + (x + ).The integration constant w can be regard<strong>ed</strong> as an alternative spectral parameter.The inverse of the spectral parameter is also important⎧⎪2ty ≡ 1 w = 2t ⎨ρ(1 + t 2 ) − 2t ˜ρ = ρ +···, t ∼ 0⎪2(14.147)⎩ρt +···, t ∼∞


The linear system 265where we consider the expansion around zero and infinity. What is thesignificance of the replacement (14.133)? A spectral parameter is requir<strong>ed</strong> if onewants to enlarge the finite Lie group to its affine extension, and the appearance oft in (14.133) fits nicely with this expectation. There is now an infinite hierarchyof fields, as one can see by expanding ˆv in t. For convenience let us pick ageneraliz<strong>ed</strong> triangular gauge, defin<strong>ed</strong> by the requirement that ˆv should be regularat t = 0, or∞∑ˆv(x; t) = exp t n ϕ n (x). (14.148)Another important feature of the linear system is the invariance under ageneralization of the symmetric space automorphism. Let us define it for η = 1n=0τ ∞ ˆv(t) = ( ˆv T ) −1 ( 1t). (14.149)In terms of the Lie algebra, the action of τ ∞ readsIt is straightforward to verify thatQ µ → Q µ , P µ →−P µ . (14.150)τ ∞ ( ˆv −1 ∂ µ ˆv) =ˆv −1 ∂ µ ˆv. (14.151)We can say that it is ˆv∂ µ ˆv ∈ H ∞ , which is the subalgebra of the Geroch groupG ∞ t which is t ∞ -invariant, as happens for finite-dimensional symmetric spaces.It is worth noticing that this property does not hold for v −1 ∂ µ v if we replace τ ∞with the transformation τ defin<strong>ed</strong> in section two.14.5.3 Derivation of the colliding plane metric by factorizationAt this point we can convince ourselves that the results obtain<strong>ed</strong> so far can beus<strong>ed</strong> to construct exact solutions of Einstein’s equations. Of central importancefor this task is the monodromy matrix, which is defin<strong>ed</strong> as follows(Å =ˆv(x; t) ˆv T x; 1 ). (14.152)tA short calculation reveals that∂ µ Å =ˆv(ˆv −1 ∂ µ ˆv − τ ∞ ( ˆv −1 ∂ µ ˆv))τ ∞ ˆv −1 = 0 (14.153)where the relation (14.151) was us<strong>ed</strong>. Consequently, Å can only depend on w.The solutions generating proc<strong>ed</strong>ure now consists in choosing a matrix Å(w) andfinding a factorization as in (14.152).


266 Infinite-dimensional symmetries in gravityThe simplest non-trivial example, that will be consider<strong>ed</strong> here, permits us torecover the Ferrari–Ibanez colliding plane wave metric. Let us consider for thisaim the monodromy matrix)Å(w) =( w0 −ww 0 +w00w 0 +ww 0 −w∈ SL(2, C) (14.154)and usew − w 0 =− ρ ( 1(t − t 0 ) 0)4t 0 t − t (14.155)with the special value w 0 = 1 2 and t 0 ≡ t(x; w 0 ).We use light cone coordinates, with the following notation to facilitate thecomparison with the standard literatureu ≡ x + , v ≡ x − (14.156)then the remaining conformal invariance is entirely fix<strong>ed</strong> by choosing thecoordinates in such a way thatρ + (u) = 1 2 (1−2u2 ), ρ − (v) = 1 2 (1−2v2 ) ⇒ ρ(u,v) = 1−u 2 −v 2 (14.157)where ρ(u,v) > 0 because the interaction region, where the <strong>waves</strong> collide, isu 2 + v 2 < 1. Substituting (14.155) into (14.154) and defining two particularsolutions (14.146) in our gauge ast 1 (u,v) ≡ tt 1 (u,v) ≡ t(u,v; w = 1 )=2(u,v; w =− 1 )2√1 − u 2 − v√1 − u 2 + v > 0 (14.158)√1 − v=−2 + u√1 − v 2 − u < 0 (14.159)where the inequalities hold in the interaction region, we obtain in astraightforward way the desir<strong>ed</strong> factorization form for the monodromy matrix.Then it follows that( √ − t 2 t−t 1)t1 t−tˆv(u,v; t) =20√ . (14.160)0 − t 1 t−t 2t2 t−t 1Putting t = 0 we recover v(u,v) in the triangular gauge, and then read directlythe result for by virtue of (14.69). We get =− t 1= 1 − ξt 2 1 + ξ , B = 0 (14.161)where the oblate spherical coordinates have been introduc<strong>ed</strong>√ √ξ ≡ u 1 − v 2 + v 1 − u 2 (14.162)√ √η ≡ u 1 − v 2 − v 1 − v 2 . (14.163)


Acknowl<strong>ed</strong>gments 267From (14.161) we have P± 2 = Q ± = 0, with P± 1 = 1 2 −1 ∂ ± . Puttingequation (14.161) into this second relation, we gain12 −1 ∂ ± = 1 2 t−1 1 ∂ ±t 1 − 1 2 t−1 2 ∂ ±t 2 (14.164)which using the formulaet −1 ∂ + t = ρ −1 ∂ + ρ 1 − t1 + t , t−1 ∂ − t = ρ −1 ∂ − ρ 1 + t1 − t(14.165)becomesP± 1 = 1 ( 1 ∓ t12 ρ−1 ∂ ± ρ − 1 ∓ t )2.1 ± t 1 1 ± t 2(14.166)Now we use the expression given here for P+ 1 to integrate the equation for theconformal factor. Some further calculations show thatλ 2 (1 − t 1 t 2 ) 2= 8uv(1 − t1 2)(1 − t2 2 ) (14.167)where the undetermin<strong>ed</strong> overall factor has been chosen for convenience. Then,this result yields the four-dimensional metric( )dξds 2 = (1+ξ) 2 21 − ξ 2 − dη21 − η 2 −ρ 2 1 − ξ1 + ξ (dx 2 ) 2 − 1 + ξ1 − ξ (dx 3 ) 2 . (14.168)This is (a special case of) the so-call<strong>ed</strong> Ferrari–Ibanez colliding plane <strong>waves</strong>olution.Acknowl<strong>ed</strong>gmentsA Nagar wishes to thank Professor Pietro Fré and Professor Vittorio Gorini fortheir encouragement, and two students of Como University, Valentina Riva andMassimo Bosetti, for their hospitality and the help given to him in arranging theselectures.Further readingKramer D et al 1980 Exact Solutions of Einstein Field Equations (Cambridge: CambridgeUniversity Press)Hoensealers C and Dietz W (<strong>ed</strong>) 1984 Solutions of Einstein’s Equations: Techniques andResults (Berlin: Springer)Maison D 1978 Phys. Rev. Lett. 41 521Belinskii V A and Zakharov V E 1978 Zh. Eksp. Teor. Fiz. 48 985Breitenlohner P and Maison D 1987 Ann. Inst. Poincaré 46 215Julia B and Nicolai H 1996 Nucl. Phys. B 482 431


Chapter 15Gyroscopes and gravitational <strong>waves</strong>Donato Bini 1 and Fernando de Felice 21 Istituto per Applicazioni della Matematica, CNR, I-80131Napoli, Italy and International Center for RelativisticAstrophysics, University of Rome, I-00185 Roma, Italy2 Dipartimento di Fisica ‘G. Galilei’, Università degli Studi diPadova, Via Marzolo, 8, I-35131, Padova, Italy and INFN,Sezione di Padova, ItalyThe behaviour of a gyroscope in geodesic motion is studi<strong>ed</strong> in the field ofa plane gravitational wave. We find that, with respect to a special set offrames, the compass of inertia undergoes a precession which, to first order in th<strong>ed</strong>imensionless amplitude h of the wave, is dominat<strong>ed</strong> by the cross-polarizationalone. This suggests that a gyro might act as a filter of the polarization state ofthe wave.15.1 IntroductionThe (direct) detection of gravitational <strong>waves</strong> is still an open question, althoughindirect evidence for their existence has been obtain<strong>ed</strong> from the observationof the binary pulsar system PSR 1913+16 [1]. Besides the well-known barantennae, there is a growing interest in laser interferometry detectors, like LIGOand VIRGO, which are sensitive to the low frequency (∼10 Hz) gravitational<strong>waves</strong> which are emitt<strong>ed</strong> by sources like coalescing binaries.The purpose of this paper is to study the behaviour of a test gyroscopewhich is act<strong>ed</strong> upon by a plane gravitational wave with the purpose to seewhether this interaction leads to observable effects. It is well known that inthe absence of significant coupling between the background curvature and themultipole moments of the energy–momentum tensor of an extend<strong>ed</strong> body, thespin vector is Fermi–Walker transport<strong>ed</strong> along the body’s own trajectory (see [2]268


Splitting formalism and test particle motion: a short review 269and references therein). The effects of a gravitational wave on a frame which isnot Fermi–Walker transport<strong>ed</strong>, are best appreciat<strong>ed</strong> by studying the precession ofa gyro at rest in that frame. Clearly it is essential to identify a class of frameswhich optimize the corresponding precession effect.In section 15.2 we give a short review of the observer dependent spacetimesplitting which enables one to describe in physical terms the motion of a testparticle as well as that of a test gyroscope. In section 15.3 we discuss thespacetime of a plane gravitational wave and confine our attention to the family ofstatic observers; we give an example of a tetrad frame adapt<strong>ed</strong> to these observerswith respect to which the precession of a gyroscope is induc<strong>ed</strong> by the crosspolarizationonly. In section 15.4 we discuss the non-trivial problem of how tofix, in an operational and non-ambiguous way, a frame of reference which is notFermi–Walker transport<strong>ed</strong> in the spacetime of a plane gravitational wave. Finally,in section 15.5 we calculate the precession of a gyroscope, in a general geodesicmotion, with respect to the above frame.In what follows, Greek indices run from 0–3, latin indices from 1–3.15.2 Splitting formalism and test particle motion: a shortreviewA given family of test observers, namely a congruence of timelike lines with unittangent vector field u (i.e. u · u =−1) induces a splitting of the spacetime intospace plus time through the orthogonal decomposition of the tangent space at eachpoint into the local time direction along u and the local rest space LRS u .Projection of spacetime tensor fields onto LRS u is accomplish<strong>ed</strong> using theprojection operatorP (u) = I + u ⊗ u (15.1)and yields a family of spatial tensors (belonging to LRS u ⊗ ··· ⊗ LRS u , i.e.for which any contraction with u vanishes). The collection of all the spatiallyproject<strong>ed</strong> tensor fields, associat<strong>ed</strong> to a given spacetime tensor field, will bereferr<strong>ed</strong> to as the ‘measure’ of the spacetime tensor itself. For instance, themeasure of the unit volume four-form η αβµν , gives only one non-trivial spatialfield: ɛ (u) αβγ= uδ η δαβγ which can be us<strong>ed</strong> in turn to define the spatial crossproduct × u in LRS u .One can also spatially project the various derivative operators so that theresult of the derivative of any tensor field is itself a spatial tensor; examplesare: the spatial Lie derivative, £ (u) X = P (u)£ X for any vector field X, thespatial covariant derivative ∇ (u) α = P β(u) P (u) α ∇ β , the Lie temporal derivative,∇ (lie,u) = P (u) £ u , the Fermi–Walker temporal derivative, ∇ (fw,u) = P (u) ∇ u andseveral other natural derivatives for which a detail<strong>ed</strong> discussion can be foundin [2].The measure of the covariant derivative of the four-velocity of the observers,gives rise to the kinematical coefficients of the observer congruence, namely the


270 Gyroscopes and gravitational <strong>waves</strong>acceleration, vorticity expansionand the spatial dual of the vorticity fielda(u) α =∇ (fw,u)u α ,γ δω (u)αβ = P (u) α P (u) β ∇ [γ u δ] ,γ δθ (u)αβ = P (u) α P (u) β ∇ (γ u δ) , (15.2)ω (u) α = 1 2 ɛ (u) αβγ ω (u) βγ . (15.3)When dealing with different families of test observers, say u and U, themix<strong>ed</strong> projection map P (U,u) = P (u) P (U) from LRS U to LRS u (and the analogouscompositions of two or more projectors) will be useful. Let l U be the world lineof a nonzero rest mass test particle with U as its unit timelike tangent vector. Theorthogonal decomposition of U relative to the family of test observers u, identifiesits relative velocity ν (U,u) = ν ˆν (U,u) where ν =‖ν (U,u) ‖=‖ν (u,U) ‖ and ˆν (U,u) isthe unit spatial vector, so thatU = γ [u + ν ˆν (U,u) ]. (15.4)Here γ = (1 − ν 2 ) −1/2 is the local relative Lorentz factor. If the four accelerationof the particlea (U) =∇ U U =D Udτ Uis non-vanishing, then its projection onto LRS u , leads to the acceleration-equalsforceequation:P (U,u) a (U) ≡ γ F (U,u)where F (U,u) is the spatial force acting on the particle as seen by the observer u.In a similar way, one defines a spatial gravitoinertial forceF (G)(fw,U,u) =−γ −1 DuP (u)dτ UDu=−P (u)dτ (U,u)= γ [g (u) + ν( 1 2 ˆν (U,u) × u H (u) − θ (u) Lˆν (U,u) )], (15.5)where τ U is a proper time parametrization for U and τ (U,u) = ∫ l Uγ dτ U is thecorresponding Cattaneo relative standard time parametrization; g (u) =−a (u) andH (u) = 2ω (u) are, respectively, the electric- and magnetic-like components ofthe gravitoinertial force. This terminology is justifi<strong>ed</strong> by the Lorentz form of thegravitoinertial force which appears in the last of equations (15.5).If we define p (U,u) = γν (U,u) , E (U,u) = γ andD (fw,U,u)dτ (U,u)= γ −1 P (u)Ddτ U=∇ (fw,u) + ν α (U,u) ∇(u) α (15.6)


Splitting formalism and test particle motion: a short review 271the latter being the measure of the (rescal<strong>ed</strong>) absolute derivative along U, the(3+1) version of the equation of motion of the particle and of the energy theorem,acquires the Newtonian formD (fw,U,u)dτ (U,u)p (U,u) = F (U,u) + F (G)(fw,U,u) ,dE (U,u)dτ (U,u)= [F (U,u) + F (G)(fw,U,u) ] ·u ν (U,u) . (15.7)Let us now consider the motion of a test gyroscope. As it is well known,the spin vector S (U) of a gyroscope carri<strong>ed</strong> by an observer U, is Fermi–Walkertransport<strong>ed</strong> along his worldline (i.e. S (U) does not precess with respect to spatialaxes which are Fermi–Walker dragg<strong>ed</strong> along U), namely:D (fw,U)dτ US (U) = P (U)Ddτ US (U) = 0. (15.8)Suppose that we have chosen a spatial triad ē(U)â which is adapt<strong>ed</strong> to the observerU and is not a Fermi–Walker frame. The observer U will then see the spin S(U)of the gyroscope to precess with respect to these axes according to the law:[ ]dS â(U)− ɛâ dτ ˆbĉ ζ ˆb(fw,U,ē(U)â) Sĉ(U) ē(U)â = 0 (15.9)Uwhereζ â(fw,U,ē(U)â) ≡ ɛâ ˆbĉē(U)ˆb ·∇ (fw,U)ē(U)ĉ (15.10)is the precession rate vector.However, we may want the gyroscope to be analys<strong>ed</strong> by a different observer,u say, who is not comoving with the gyro’s centre of mass. In this case we ne<strong>ed</strong>a smooth family of these observers, each one intersecting the gyro’s worldline atany of its spacetime points where he measures the instantaneous precession ofthe spin vector relative to a suitably defin<strong>ed</strong> frame, adapt<strong>ed</strong> to u. Of course, werequire that the observer’s u are synchroniz<strong>ed</strong> so that their measurements can becompar<strong>ed</strong>. The results of these measurements are describ<strong>ed</strong> by a smooth and atleast once differentiable function of the proper-time of u.Let {e(u)â} be a field of spatial triads adapt<strong>ed</strong> to u; then the restriction of{e(u)â} to the worldline l U of the gyroscope, allows one to define on l U a fieldof tetrad frames, adapt<strong>ed</strong> to U, given by {(U, ē(U)â}, where:ē(U)â = B (lrs,u,U) e(u)â, (15.11)B (lrs,u,U) = P (U) B (u,U) P (u) : LRS u → LRS U being the boost map between therest spaces of the observers U and u; this map has been studi<strong>ed</strong> extensively in[2–4]. Since the boost is an isometry, the precession of S (U) with respect to the


272 Gyroscopes and gravitational <strong>waves</strong>axes ē(U)â is the same as the precession, with respect to the axes e(u)â, of theboost<strong>ed</strong> spinvector s (u) , which reads:[s (u) = B (lrs,U,u) S (U) = P (u) −γ]γ + 1 ν (U,u) ⊗ ν (U,u) P (U,u) S (U) . (15.12)Hence, from (15.9) and (15.11) and the acceleration-equals-force equationfor U; wefind:[ ]ds â(u)− γɛâ dτ ˆbĉ [ζ (fw,U,u) + ζ (sc,fw,U,u) ] ˆb sĉ(u) e(u)â = 0 (15.13)Uζ (fw,U,u) = 1γ + 1 ν (U,u) × u [F (G)(fw,U,u) − γ F (U,u)]where 1ζ (sc,fw,U,u) = 1 2 δâ ˆb D (fw,U,u)dτ (U,u)e(u)â × u e(u) ˆb(15.14)so thatζ (fw,U,ē(U)â) = γ B (lrs,u,U) [ζ (fw,U,u) + ζ (sc,fw,U,u) ]. (15.15)Finally, by rescaling equation (15.13) to the proper-time of u, one hasds â (u)− ɛâ dτ ˆbĉ [ζ (fw,U,u) + ζ (sc,fw,U,u) ] ˆb s (u)ĉ = 0 (15.16)(U,u)whereζ (fw,U,u) + ζ (sc,fw,U,u) ≡ ˜ζ (fw,u,e(u)â) = γ −1 B (lrs,U,u) ζ (fw,U,ē(U)â) (15.17)is the angular velocity precession of the gyroscope as measur<strong>ed</strong> by the observer uwith respect to the axes e(u)â.It is worth mentioning here that while the observer U, who is comovingwith the gyro’s centre of mass, measures the precession (15.10) along his ownworldline, the observer’s u can only compare the instantaneous measurementsof ˜ζ (fw,u,e(u)â) in (15.17), made by each of them along the gyro’s worldline.Evidently either type of measurements requires the tetrad frames to beoperationally well defin<strong>ed</strong>. This will be discuss<strong>ed</strong> in the following section.15.3 The spacetime metricThe metric of a plane monochromatic gravitational wave, elliptically polariz<strong>ed</strong>and propagating along a direction which we fix as the coordinate x direction, canbe written in the ‘TT’ gauge as [5]:ds 2 =−dt 2 + dx 2 + (1 − h 22 ) dy 2 + (1 + h 22 ) dz 2 − 2h 23 dy dz (15.18)1 This notation for the Fermi–Walker relative angular velocity ζ (fw,U,u) and the Fermi–Walker spacecurvaturerelative angular velocity ζ (sc,fw,U,u) has been introduc<strong>ed</strong> in [2].


The spacetime metric 273with h AB = h AB (t − x), (A, B = 2, 3). Let u denote the tangent vector field ofa family of geodesic, non-rotating and expanding observers defin<strong>ed</strong> byu ♭ =−dt, u = ∂ t . (15.19)One can adapt to these observers an infinite number of spatial frames, by rotatingany given one arbitrarily. For example, consider the following u-frame {e ˆα }={eˆ0 = u, e â = e(u)â} with its dual {ω ˆα }={ωˆ0 =−u ♭ ,ωâ = ω(u)â}u = ∂ te(u)ˆ1 = ∂ xe(u)ˆ2 = (1 − h 22) −1/2 ∂ y ≃ (1 + 1 2 h 22)∂ ye(u)ˆ3 = (1 − h2 22 − h2 23 )−1/2 [(1 − h 22 ) −1/2 h 23 ∂ y + (1 − h 22 ) 1/2 ∂ z ]≃ h 23 ∂ y + (1 − 1 2 h 22)∂ z (15.20)−u ♭ = dtω(u)ˆ1 = dxω(u)ˆ2 = (1 − h 22 ) 1/2 [dy − h 23ω(u)ˆ3 =(1 − h222− h 2 231 − h 22) 1/2dz ≃]dz ≃1 − h 22(1 − 1 )2 h 22 dy − h 23 dz(1 + 1 )2 h 22 dz,where ≃ denotes the corresponding weak-field limit. Any other spatial frame{ẽ(u)â}, adapt<strong>ed</strong> to the observers (15.19), can be obtain<strong>ed</strong> from this one by aspatial rotation Rẽ(u)â = e(u) ˆb R ˆbâ. (15.21)Among all the possible frames, there exists only one with respect to which thelocal compass of inertia experiences no precession. This frame is Fermi–Walkertransport<strong>ed</strong> along u, namely it satisfies the conditionP (u)Ddτ uẽ(u)â ≡∇ (fw,u) ẽ(u)â = 0. (15.22)A Fermi–Walker frame is the most natural of the u-frames; its spatialdirections, in fact, are fix<strong>ed</strong> by three mutually orthogonal axes of small sizecomoving gyroscopes. However, if a metric perturbation causes a dragging ofthe local compass of inertia, the only way to detect and measure it, is to select aframe which is not Fermi–Walker transport<strong>ed</strong>. In this frame, in fact, a gyroscopewould be seen to precess and inde<strong>ed</strong> its precession contains all the informationsabout gravitational dragging. Nonetheless, it is quite non-trivial to identify, inan operational way, a frame which is not Fermi–Walker transport<strong>ed</strong> along theobserver’s worldline when it is act<strong>ed</strong> upon by a gravitational wave.


274 Gyroscopes and gravitational <strong>waves</strong>Frame (15.20) is clearly Fermi–Walker transport<strong>ed</strong> in the absence ofgravitational <strong>waves</strong> (h 22 and h 23 being time independent), but it is not so whenthey are present. The Fermi rotation of the frame, in this case, is describ<strong>ed</strong> by the(antisymmetric) angular velocity spatial tensor [6]:C (u) ˆbâ = e(u) ˆb ·u ∇ (fw,u) e(u)â, (15.23)hence, a gyroscope carri<strong>ed</strong> by the observer u will precess with respect to frame(15.20) with an angular velocity tensor which has only one independent nonzerocomponent, namely:C =−[h 23,t(1 − h 22 ) + h 22,t h 23 ](u)ˆ3ˆ2√≃− 12(1 − h 22 ) 1 − h 2 22 − h2 2 h 23,t. (15.24)23However, frame e(u)â cannot be operationally defin<strong>ed</strong>, so result (15.24) is of littlephysical significance although it shows the existence of frames which respond toone state of polarization only, at least to first order in h AB . We are, therefore,motivat<strong>ed</strong> to search for ‘frames’ that can be fix<strong>ed</strong> from a viable experimental setup.15.4 Searching for an operational frameLet us consider the timelike geodesics of the metric (15.18). These are well known[7]; the four-velocity of a general such geodesic can be written asU g = 12E [(1 + f + E 2 )∂ t + (1 + f − E 2 )∂ x ]1+1 − h 2 22 − {[α(1 + h 22 ) + βh 23 ]∂ y + [β(1 − h 22 ) + αh 23 ]∂ z },h2 23(15.25)where α, β and E are Killing constants and f = g AB U A U B is equal to1f =1 − h 2 22 − [α 2 (1 + h 22 ) + β 2 (1 − h 22 ) + 2αβh 23 ]h2 23≃ α 2 (1 + h 22 ) + β 2 (1 − h 22 ) + 2αβh 23 . (15.26)If u = ∂ t is the family of observers who make the mesurements and {e(u) â}is an adapt<strong>ed</strong> spatial frame, then the relative velocity ν (Ug ,u)â of U g with respectto u is defin<strong>ed</strong> by the relationU g = γ (Ug ,u)[u + ν (Ug ,u)âe(u)â], (15.27)


Searching for an operational frame 275whereγ (Ug ,u) = 1 + f + E22E≃ 12E [1 + E 2 + α 2 (1 + h 22 ) + β 2 (1 − h 22 ) + 2αβh 23 ] (15.28)and the relative velocity components can be obtain<strong>ed</strong> by comparison of (15.28)and (15.25).Hereafter, we restrict ourselves to the weak field approximation (first order inh AB ) and assume, without loss of generality that, in the absence of a gravitationalwave, the spatial velocity ν (Ug ,u), was only in the y coordinate direction. Thiscorresponds to the requirement thatIn this case we findν (U,u) ˆ2 =β = 0, E =√1 + α 2 . (15.29)α 2ν ˆ1 (U,u) =2(1 + α 2 ) h 22,[]αh 22√ 1 +, (15.30)1 + α 22(1 + α 2 )ν ˆ3 (U,u) =α√1 + α 2 h 23.We now require that the four-velocity of a test gyroscope is U = U g as givenby (15.25) with conditions (15.29). The assumption of geodesicity is justifi<strong>ed</strong> ifwe consider the limit of zero size gyroscopes. In this case, in fact, not only can weneglect the multipole moments of the gyro’s stress tensor higher than the dipole,but also ignore the tidal term which enters the Papapetrou–Dixon equations andarises from the coupling of the gyro’s spin with the background curvature. Thisterm is of the order of the ratio of the average size of the gyroscope and thegravitational-wave wavelength.Let us consider the restriction of the vector field u = ∂ t of stationaryobservers, to the gyroscope’s worldline and require that these observers monitorthe behaviour of the spin of the moving gyro, measuring the (instantaneous)precession vector ˜ζ given by equation (15.17). As already mention<strong>ed</strong>, the spatialframe e(u)â given by (15.20) is not operationally well defin<strong>ed</strong>, so we have to findone which is so.To find a spatial frame which is suitable for actual experiments, note that theobserver’s u can unambiguously determine in their rest-frame, a spatial directiongiven by that of the relative velocity ν of the gyroscope. Suppose they fix, byguessing if necessary, a direction of propagation of the gravitational wave andterm this as the x-axis with unit vector e(u) ˆ1. Then from these two directions,


276 Gyroscopes and gravitational <strong>waves</strong>namely that of the relative velocity of the gyro and of the wave propagation, it ispossible to construct a spatial triad as followsλ(u)ˆ1 = e(u)ˆ1 = ∂ xλ(u)ˆ2 =ˆν (U,u) × u e(u)ˆ1λ(u)ˆ3 = λ(u)ˆ1 × uλ(u)ˆ2 . (15.31)This frame can be operationally construct<strong>ed</strong> apart from guessing th<strong>ed</strong>irection of propagation of the wave. Inde<strong>ed</strong> such a guess is also requir<strong>ed</strong> to fitdata from bar antenna detectors, for example. Obviously this frame is not unique:any other spatial triad obtain<strong>ed</strong> from it after a rotation which depends at most onthe (known) modulus of the relative velocity (or is constant) is equally useful.The spatial triads e(u)â in (15.20) and λ(u)â in (15.31), differ by a rotationλ(u)â = Ê ˆbâe(u)ˆb .In the weak field limit, the only non-trivial components of Ê ˆbâ, areʈ1 =−ʈ2 ˆ1 ˆ3 = ʈ3 ˆ2 = 1, ʈ2 ˆ2 = ʈ3 ˆ3 = h 23so thatλ(u)ˆ1 = ∂ xλ(u)ˆ2 ≃ h 23 e(u)ˆ2 − e(u)ˆ3 ≃−(1 − 1 2 h 22)∂ zλ(u)ˆ3 ≃ e(u)ˆ2 + h 23e(u)ˆ3 ≃ (1 + 1 2 h 22)∂ y + h 23 ∂ z . (15.32)15.5 Precession of a gyroscope in geodesic motionThe precession of the gyro which is measur<strong>ed</strong> by the observer’s u all along itsworldline, is the image of the precession measur<strong>ed</strong> by the comoving observer U,under the boost B(U, u) as shown in (15.17). In order to study the spin precessionseen by the observer comoving with the gyro, we must first decide with respectto what axes (non-Fermi–Walker transport<strong>ed</strong> but operationally well defin<strong>ed</strong>) theprecession will be measur<strong>ed</strong>, as explain<strong>ed</strong> in section 15.2.Since the observer’s u intersect the worldline of the observer U carrying thegyro, the u-frames {λ(u)â} in (15.31) form a smooth field of frames on it, so theobserver U can identify spatial directions in his rest space simply by boosting th<strong>ed</strong>irections λ(u)â.At each event along his worldline in fact he will see the axes λ(u) â defin<strong>ed</strong>in (15.31) to be in relative motion, therefore the boost of these axes, namely¯λ(U)â = B (lrs,u,U) λ(u)â = λ(u)â +γγ + 1 [ν (U,u) · λ(u)â](u + U), (15.33)


Precession of a gyroscope in geodesic motion 277from (15.11), will be seen by U as the corresponding axes with the sameorientation which are ‘momentarily at rest’. The orientation of the spin vectorS with respect to the axes ¯λ(U)â is also the orientation of S with respect to themoving axes λ(u)â.The velocity of spin precession then corresponds to the spatial dual ofthe Fermi–Walker structure functions of ¯λ(U)â, namely C (fw) (U, ¯λ(U)â) ˆbâ ,according to relation (15.23).Confining our attention to the weak-field approximation, the components ofthe precession velocity with respect to the triad ¯λ(U)â areζ (fw,U,¯λ(U)â) ˆ1 ≃ − 1 2 h 23,tζ (fw,U,¯λ(U)â) ˆ2 ≃ α/2h 22,tζ (fw,U,¯λ(U)â) ˆ3 ≃ α/2h 23,t . (15.34)We observe that in the limit of small linear momentum, α ≪ 1, th<strong>ed</strong>ominant precession is in the direction of wave propagation e(u) ˆ1(to zerothorder, ¯λ(U)ˆ1 ≃ e(u)ˆ1 ) and is induc<strong>ed</strong> by the cross-polarization only. (Note thatthe precession in the direction of propagation of the wave does not depend onα.) In this case we can conclude that the gyro can act as a polarization filter forgravitational <strong>waves</strong>.In the opposite limit of large linear momentum, α ≫ 1, the precession vectorlies mainly in the plane orthogonal to the propagation direction and is contribut<strong>ed</strong>likewise by both polarizations. Inde<strong>ed</strong> the measurement of the precession induc<strong>ed</strong>by a plane gravitational wave, of a gyroscope set in relativistic motion, wouldenable one to identify the local direction of propagation of the wave. A similarsituation will be encounter<strong>ed</strong> in the rest frame of u, where from (15.17), (15.28)and (15.29) we have:˜ζ ˆ1 = γ −1(fw,u,λ(u)â) (U,u) B (lrs,U,u)ζ ˆ1 (fw,U,¯λ(U)â) ≃− 1 2˜ζ ˆ2 = γ −1(fw,u,λ(u)â) (U,u) B (lrs,U,u)ζ ˆ2 (fw,U,¯λ(U)â) ≃ 1 2˜ζ ˆ3 = γ −1(fw,u,λ(u)â) (U,u) B (lrs,U,u)ζ ˆ3 (fw,U,¯λ(U)â) ≃ 1 21√ h 23,t1 + α 2α√ h 22,t1 + α 2α√ h 23,t. (15.35)1 + α 2Finally, let us note that results (15.35) are only slightly modifi<strong>ed</strong> afterrotating the frame (15.31) by a constant angle φ around the propagation directionof the wave. In fact, in this case, the new spatial u-frame becomesf (u)ˆ1 = λ(u)ˆ1f (u)ˆ2 = cos φλ(u)ˆ2 + sin φλ(u)ˆ3f (u)ˆ3 =−sin φλ(u)ˆ2 + cos φλ(u)ˆ3 ; (15.36)


278 Gyroscopes and gravitational <strong>waves</strong>when φ = 0 it r<strong>ed</strong>uces to (15.31). Once the boost<strong>ed</strong> frame ¯ f (U)â =B (lrs,U,u) f (u)â in LRS U is obtain<strong>ed</strong>, the components of the precession velocityturn out to beζ (fw,U, ¯ f (U)â) ˆ1 ≃− 1 2 h 23,tζ (fw,U, f ¯ˆ2 (U)â) ≃ α/2[cos φh 22,t − sin φh 23,t ]ζ (fw,U, f ¯ˆ3 (U)â) ≃ α/2[− sin φh 22,t + cos φh 23,t ], (15.37)again showing that, in the limit of small linear momentum α, the precessionmainly occurs about the direction of propagation of the wave.15.6 ConclusionsWe have operationally defin<strong>ed</strong> a tetrad frame adapt<strong>ed</strong> to a family of staticobservers in the background of a plane gravitational wave. Then we have us<strong>ed</strong>this family to study the precession angular velocity of a gyroscope moving alonga spacetime geodesic. The results show that, to first order in h and in thecase of non-relativistic motion (α ≪ 1), the observ<strong>ed</strong> gyroscopic precession ismainly induc<strong>ed</strong> by the cross-polarization only so a gyro appears to behave as apolarization filter.Assuming the form h 23 = h × sin(λ 2πcGW(t − x/c) + ψ) for the crosspolarization,with an obvious meaning for the symbols, the precession frequency(in conventional units) would be (gyro) (t) ≃− πch ×λ GW( )2πccos (t − x/c) + ψ . (15.38)λ GWThe values of the precession frequency are very small, as expect<strong>ed</strong>. With a typicalamplitude of 10 −21 at the Earth and a frequency of 10 3 Hz, we could hope for amaximum precession of the order of 10 −18 s −1 .Clearly the precession effect is larger with high-frequency gravitational<strong>waves</strong> or when the gyroscope is close enough to the wave source to allow fora higher value of h × . This situation is encounter<strong>ed</strong> by a spinning neutron star,say, in a compact binary system.The type of analysis we have consider<strong>ed</strong>, is most suitable to describe theinteraction of a moving gyroscope with a continuous flow of plane gravitational<strong>waves</strong> with metric form as in (15.18). These are expect<strong>ed</strong> to be emitt<strong>ed</strong> bysources like compact binaries. If we have an impulsive source, like a supernova,then one expects a burst of gravitational radiation which is better describ<strong>ed</strong> by agravitational sandwich. This latter case is now under investigation.


References 279References[1] Taylor J H and Weisberg J M 1989 Astrophys. J. 345 434[2] Jantzen R T, Carini P and Bini D 1992 Ann. Phys. 215 1[3] Bini D, Carini P and Jantzen R T 1997a Int. J. Mod. Phys. D 6 1[4] Bini D, Carini P and Jantzen R T 1997b Int. J. Mod. Phys. D 6 143[5] Misner C W, Thorne K S and Wheeler J A 1973 Gravitation (New York: Freeman)[6] de Felice F and Clarke C J S 1990 Relativity on Curv<strong>ed</strong> Manifolds (Cambridge:Cambridge University Press)[7] de Felice F 1978 J. Phys. A: Math. Gen. 12 1223


Chapter 16Elementary introduction to pre-big bangcosmology and to the relic gravitonbackgroundMaurizio GasperiniDipartimento di Fisica, Università di Bari, Via G. Amendola 173,70126 Bari, Italy and Istituto Nazionale di Fisica Nucleare,Sezione di Bari, Bari, ItalyThis is a contract<strong>ed</strong> version of a series of lectures for graduate and undergraduatestudents given at the ‘VI Seminario Nazionale di Fisica Teorica’ (Parma,September 1997), at the Second International Conference ‘Around VIRGO’ (Pisa,September 1998), and at the Second SIGRAV School on ‘<strong>Gravitational</strong> Wavesin Astrophysics, Cosmology and String Theory’ (Center ‘A. Volta’, Como, April1999). The aim is to provide an elementary, self-contain<strong>ed</strong> introduction tostring cosmology and, in particular, to the background of relic cosmic gravitonspr<strong>ed</strong>ict<strong>ed</strong> in the context of the so-call<strong>ed</strong> ‘pre-big bang’ scenario. No specialpreparation is requir<strong>ed</strong> besides a basic knowl<strong>ed</strong>ge of general relativity and ofstandard (inflationary) cosmology. All the essential computations are report<strong>ed</strong>in full details either in the main text or in the appendices. For a deeper and morecomplete approach to the pre-big bang scenario the interest<strong>ed</strong> reader is referr<strong>ed</strong> tothe updat<strong>ed</strong> collection of papers available at http://www.to.infn.it/∼gasperin/.16.1 IntroductionThe purpose of these lectures is to provide an introduction to the background ofrelic gravitational <strong>waves</strong> expect<strong>ed</strong> in a string cosmology context, and to discussits main properties. To this purpose, it seems to be appropriate to include a shortpresentation of string cosmology, in order to explain the basic ideas underlying280


Introduction 281the so-call<strong>ed</strong> pre-big bang scenario, which is one of the most promising scenariosfor the production of a detectable graviton background of cosmological origin.After a short, qualitative presentation of the pre-big bang models, we willconcentrate on the details of the cosmic graviton spectrum: we will discuss thetheoretical pr<strong>ed</strong>ictions for different models, and will compare the pr<strong>ed</strong>ictionswith existing phenomenological constraints, and with the expect<strong>ed</strong> sensitivitiesof the present gravity-wave detectors. A consistent part of these lectures willthus be devot<strong>ed</strong> to introducing the basic notions of cosmological perturbationtheory, which are requir<strong>ed</strong> to compute the graviton spectrum and to understandwhy the amplification of tensor metric perturbations, at high frequency, is moreefficient in string cosmology than in the standard inflationary context. Let me startby noting that a qualitative, but effective representation of the main differencebetween string cosmology and standard, inflationary cosmology can be obtain<strong>ed</strong>by plotting the curvature scale of the Universe versus time, as illustrat<strong>ed</strong> infigure 16.1.According to the cosmological solutions of the so-call<strong>ed</strong> ‘standard’ scenario[1], the spacetime curvature decreases in time. As we go back in time thecurvature grows monotonically, and blows up at the initial ‘big bang’ singularity,as illustrat<strong>ed</strong> in the top part of figure 16.1 (a similar plot, in the standard scenario,also describes the behaviour of the temperature and of the energy density of thegravitational sources).According to the standard inflationary scenario [2], in contrast, the Universein the past is expect<strong>ed</strong> to enter a de Sitter, or ‘almost’ de Sitter phase, during whichthe curvature tends to stay frozen at a nearly constant value. From a classicalpoint of view, however, this scenario has a problem, since a phase of expansionat constant curvature cannot be extend<strong>ed</strong> back in time for ever [3], for reasons ofgeodesic completeness. This point was clearly stress<strong>ed</strong> also in Alan Guth’s recentsurvey of inflationary cosmology [4]:. . . Nevertheless, since inflation appears to be eternal only into thefuture, but not to the past, an important question remains open. Howdid it all start? Although eternal inflation pushes this question far intothe past, and well beyond the range of observational tests, the questiondoes not disappear.A possible anwer to this question, in a quantum cosmology context, isthat the universe emerges in a de Sitter state ‘from nothing’ [5] (or from someunspecifi<strong>ed</strong> ‘vacuum’), through a process of quantum tunnelling. We willnot discuss the quantum approach in these lectures, but let us note that thecomputation of the transition probability requires an appropriate choice of theboundary conditions [6], which in the context of standard inflation are impos<strong>ed</strong> adhoc when the universe is in an unknown state, deeply inside the non-perturbative,quantum gravity regime. In a string cosmology context, in contrast, the initialconditions are referrr<strong>ed</strong> asymptotically to a low-energy, classical state which isknown, and well controll<strong>ed</strong> by the low-energy string effective action [7].


282 Elementary introduction to pre-big bang cosmologyCurvature scale versus timeBIG BANGsingularitypresent timeSTANDARD COSMOLOGYtim<strong>ed</strong>e SitterSTANDARD INFLATIONtimeM SString pert.vacuumPRE-BIG BANGPOST-BIG BANGSTRING COSMOLOGYtimeFigure 16.1. Time evolution of the curvature scale in the standard cosmological scenario,in the conventional inflationary scenario, and in the string cosmology scenario.From a classical point of view, however, the answer to the above question—what happens to the universe before the phase of constant curvature, which cannotlast for ever—is very simple, as we are left with only two possibilities. Either thecurvature starts growing again, at some point in the past (but in this case thesingularity problems remain, it is simply shift<strong>ed</strong> back in time), or the curvaturestarts decreasing.In this second case we are just l<strong>ed</strong> to the string cosmology scenario,illustrat<strong>ed</strong> in the bottom part of figure 16.1. String theory suggests inde<strong>ed</strong> forthe curvature a specular behaviour (or better a ‘dual’ behaviour, as we shall seein a moment) around the time axis. As we go back in time the curvature grows,reaches a maximum controll<strong>ed</strong> by the string scale, and then decreases towardsa state which is asymptotically flat and with negligible interactions (vanishing


Motivations: duality symmetry 283coupling constants), the so-call<strong>ed</strong> ‘string perturbative vacuum’. In this scenariothe phase of high, but finite (nearly Planckian) curvature is what replaces the bigbang singularity of the standard scenario. It thus comes naturally, in a stringcosmology context, to call ‘pre-big bang’ [8] the initial phase with growingcurvature, in contrast to the subsequent, standard, ‘post-big bang’ phase withdecreasing curvature.At this point, a number of questions may arise naturally. In particular:• Motivations: why such a cosmological scenario, characteriz<strong>ed</strong> by a ‘belllike’shape of the curvature, seems to emerge in a string cosmology contextand not, for instance, in the context of standard cosmology bas<strong>ed</strong> on theEinstein equations?• Kinematics: in spite of the differences, is the kinematic of the pre-big bangphase still appropriate to solve the well-known problems (horizon, flatness...) of the standard scenario? After all, we do not want to lose the mainachievements of the conventional inflationary models.• Phenomenology: are there phenomenological consequences that candiscriminate between string cosmology models and other inflationarymodels? Are such effects observable, at least in principle?In the following sections we will present a quick discussion of the threepoints list<strong>ed</strong> above.16.2 Motivations: duality symmetryThere are various motivations, in the context of string theory, suggesting acosmological scenario like that illustrat<strong>ed</strong> in figure 16.1. All the motivations arehowever relat<strong>ed</strong>, more or less directly, to an important property of string theory,the duality symmetry of the effective action.To illustrate this point, let us start by recalling that in general relativity thesolutions of the standard Einstein action,S =− 1 ∫2λ d−1 d d+1 x √ |g|R (16.1)p(d is the number of spatial dimensions, and λ p = Mp−1 is the Planck lengthscale), are invariant under ‘time-reversal’ transformations. Consider, for instance,a homogeneous and isotropic solution of the cosmological equations, represent<strong>ed</strong>by a scale factor a(t):ds 2 = dt 2 − a 2 (t) dxi 2 . (16.2)If a(t) is a solution, then also a(−t) is a solution. On the other hand, when t goesinto −t, the Hubble parameter H =ȧ/a changes sign,a(t) → a(−t), H =ȧ/a →−H. (16.3)


284 Elementary introduction to pre-big bang cosmologyTo any standard cosmological solution H (t), describing decelerat<strong>ed</strong> expansionand decreasing curvature (H > 0, Ḣ < 0), is thus associat<strong>ed</strong> with a ‘reflect<strong>ed</strong>’solution, H (−t), describing a contracting universe because H is negative.This is the situation in general relativity. In string theory the action, inaddition to the metric, contains at least another fundamental field, the scalardilaton φ. At the tree-level, namely to lowest order in the string couplingand in the higher-derivatives (α ′ ) string corrections, the effective action whichguarantees the absence of conformal anomalies for the motion of strings in curv<strong>ed</strong>backgrounds (see apppendix A) can be written as:S =− 1 ∫2λ d−1 d d+1 x √ |g|e −φ [R + (∂ µ φ) 2 ] (16.4)s(λ s = Ms−1 is the fundamental string length scale; see appendix B for notationsand sign conventions). In addition to the invariance under time-reversal, the aboveaction is also invariant under the ‘dual’ inversion of the scale factor, accompani<strong>ed</strong>by an appropriate transformation of the dilaton (see [9] and the first paper of [8]).More precisely, if a(t) is a solution for the cosmological background (16.2) , thena −1 (t) is also a solution, provid<strong>ed</strong> the dilaton transforms as:a →ã = a −1 , φ → ˜φ = φ − 2d ln a (16.5)(this transformation implements a particular case of T -duality symmetry, usuallycall<strong>ed</strong> ‘scale factor duality’, see appendix B).When a goes into a −1 , the Hubble parameter H again goes into −H sothat, to each one of the two solutions relat<strong>ed</strong> by time reversal, H (t) and H (−t),is associat<strong>ed</strong> a dual solution, ˜H(t) and ˜H(−t), respectively (see figure 16.2).The space of solutions is thus reach<strong>ed</strong> in a string cosmology context. Inde<strong>ed</strong>,because of the combin<strong>ed</strong> invariance under the transformations (16.3) and (16.5),a cosmological solution has in general four branches: two branches describeexpansion (positive H ), two branches describe contraction (negative H ). Also,as illustrat<strong>ed</strong> in figure 16.2, for two branches the curvature scale (∼ H 2 ) growsin time, with a typical ‘pre-big bang’ behaviour, while for the other two branchesthe curvature scale decreases, with a typical ‘post-big bang’ behaviour.It follows, in this context, that to any given decelerat<strong>ed</strong> expandingsolution, H (t) > 0, with decreasing curvature, Ḣ(t) < 0 (typical of thestandard cosmological scenario), is always associat<strong>ed</strong> a ‘dual partner’ describingaccelerat<strong>ed</strong> expansion, ˜H(−t) >0, and growing curvature,˙˜H(−t) >0. Thisdoubling of solutions has no analogue in the context of Einstein cosmology, wherethere is no dilaton, and the duality symmetry cannot be implement<strong>ed</strong>.It should be stress<strong>ed</strong>, before proce<strong>ed</strong>ing further, that the duality symmetry isnot restrict<strong>ed</strong> to the case of homogeneous and isotropic backgrounds like (16.2),but is expect<strong>ed</strong> to be a general property of the solutions of the string effectiveaction (possibly valid at all orders [10], with the appropriate generalizations).The inversion of the scale factor, in particular, and the associat<strong>ed</strong> transformation


Motivations: duality symmetry 285Expandingpre-big bangHExpandingpost-big bang~H (-t)H (t)tH (-t)~H (t)Contractingpre-big bangContractingpost-big bangFigure 16.2. The four branches of a low-energy string cosmology background.(16.5), is only a special case of a more general, O(d, d) symmetry of the stringeffective action, which is manifest already at the lowest order. In fact, the treelevelaction in general contains, besides the metric and the dilaton, also a secondrank antisymmetric tensor B µν , the so-call<strong>ed</strong> Kalb–Ramond ‘universal’ axion:S =− 1 ∫2λ d−1 d d+1 x √ [|g|e −φ R + (∂ µ φ) 2 − 1 ]s12 (∂ [µB να] ) 2 . (16.6)Given a background, even anisotropic, but with d Abelian isometries, this actionis invariant under a global, pseudo-orthogonal group of O(d, d) transformationswhich mix in a non-trivial way the components of the metric and of theantisymmetric tensor, leaving invariant the so-call<strong>ed</strong> ‘shift<strong>ed</strong>’ dilaton φ:φ = φ − ln √ | det g ij |. (16.7)In the particular, ‘cosmological’ case in which we are interest<strong>ed</strong>, the disometries correspond to spatial translations (namely, we are in the case of ahomogeneous, Bianchi I type metric background). For this background, the action(16.6) can be rewritten in terms of the (2d × 2d) symmetric matrix M,defin<strong>ed</strong> bythe spatial components of the metric, g ij , and of the axion, B ij , as:(G−1−GB ≡ B ij , G ≡ g ij , M =−1 )BBG −1 G − BG −1 . (16.8)BIn the cosmic time gauge, the action takes the form (see appendix B)S =− λ s2∫dt e[( −φ ˙φ) 2 + 1 ]8 Tr Ṁ(M−1 )˙ , (16.9)


286 Elementary introduction to pre-big bang cosmologyand is manifestly invariant under the set of global transformations [11]:( )φ → φ, M → T M , T 0 Iη = η, η = , (16.10)I 0where I is the d-dimensional unit matrix, and η is the O(d, d) metric in offdiagonalform. The transformation (16.5), representing scale factor duality, isnow reproduc<strong>ed</strong> as a particular case of (16.10) with the trivial O(d, d) matrix = η, and for an isotropic background with B µν = 0.This O(d, d) symmetry holds even in the presence of matter sources,provid<strong>ed</strong> they transform according to the string equations of motion in the givenbackground [12]. In the perfect fluid approximation, for instance, the inversionof the scale factor corresponds to a reflection of the equation of state, whichpreserves however the ‘shift<strong>ed</strong>’ energy ρ = ρ| det g ij | 1/2 :a →ã = a −1 , φ → φ, p/ρ →−p/ρ, ρ → ρ. (16.11)A detail<strong>ed</strong> discussion of the duality symmetry is outside the purpose ofthese lectures. What is important, in our context, is the simultaneous presenceof duality and time-reversal symmetry: by combining these two symmetries, infact, it is possible in principle to obtain cosmological solutions of the ‘self-dual’type, characteriz<strong>ed</strong> by the conditionsa(t) = a −1 (−t), φ(t) = φ(−t). (16.12)They are important, as they connect in a smooth way the phase of growingand decreasing curvature, and also describe a smooth evolution from the stringperturbative vacuum (i.e. the asymptotic no-interaction state in which φ →−∞and the string coupling is vanishing, g s = exp(φ/2) → 0), to the presentcosmological phase in which the dilaton is frozen, with an expectation value [13]〈g s 〉=M s /M p ∼ 0.3–0.03 (see figure 16.3).H g S = exp (ttFigure 16.3. Time evolution of the curvature scale H and of the string couplingg s = exp(φ/2) ≃ M s /M p , for a typical self-dual solution of the string cosmologyequations.


Motivations: duality symmetry 287The explicit occurrence of self-dual solutions and, more generally, ofsolutions describing a complete and smooth transition between the phase of preandpost-big bang evolution, seems to require in general the presence of higherorder (higher loop and/or higher derivative) corrections to the string effectiveaction [14] (see, however, [15, 16]). So, in order to give only a simple exampleof combin<strong>ed</strong> {duality ⊕ time-reversal} transformation, let us consider here thelow-energy, asymptotic regimes, which are well describ<strong>ed</strong> by the lowest ordereffective action.By adding matter sources, in the perfect fluid form, to the action (16.4), thestring cosmology equations for a d = 3, homogeneous, isotropic and conformallyflat background can be written as (see appendix C, equations (16.200), (16.202),(16.199), respectively):˙φ 2 − 6H ˙φ + 6H 2 = e φ ρ,Ḣ − H ˙φ + 3H 2 = 1 2 eφ p,2 ¨φ + 6H ˙φ − ˙φ 2 − 6Ḣ − 12H 2 = 0. (16.13)For p = ρ/3, in particular, they are exactly solv<strong>ed</strong> by the standard solution withconstant dilaton (see equations (16.216) and (16.217)),a ∼ t 1/2 , ρ = 3p ∼ a −4 , φ = constant, t →+∞, (16.14)describing decelerat<strong>ed</strong> expansion and decreasing curvature scale:ȧ > 0, ä < 0, Ḣ < 0. (16.15)This is exactly the radiation-dominat<strong>ed</strong> solution of the standard cosmologicalscenario, bas<strong>ed</strong> on the Einstein equations. In string cosmology, however, to thissolution is associat<strong>ed</strong> a ‘dual complement’, i.e. an additional solution which canbe obtain<strong>ed</strong> by applying on the background (16.14) a time-reversal transformationt →−t, and the duality transformation (16.11):a ∼ (−t) −1/2 , φ ∼−3ln(−t), ρ =−3p ∼ a −2 , t →−∞. (16.16)This is still an exact solution of equations (16.13) (see appendix C), describinghowever accelerat<strong>ed</strong> (i.e. inflationary) expansion, with growing dilaton andgrowing curvature scale:ȧ > 0, ä > 0, Ḣ > 0. (16.17)We note, for future reference, that accelerat<strong>ed</strong> expansion with growing curvatureis usually call<strong>ed</strong> ‘superinflation’ [17], or ‘pole-inflation’, to distinguish it from themore conventional power-inflation, with decreasing curvature.The two solutions (16.14) and (16.16) provide a particular, explicitrepresentation of the scenario illustrat<strong>ed</strong> in figure 16.3, in the two asympotic


288 Elementary introduction to pre-big bang cosmologyTable 16.1. Analogy between supersymmetry and duality.SupersymmetryDuality + Time-reversalPair of partners {bosons, fermions} {growing curvature, decreasing curvature}Known states photons, gravitons, . . . decelerat<strong>ed</strong>, standard post-big bangPr<strong>ed</strong>ict<strong>ed</strong> photinos, gravitinos, . . . accelerat<strong>ed</strong>, inflationary pre-big bangregimes of t large and positive, and t large and negative, respectively. The dualitysymmetry seems thus to provide an important motivation for the pre-big bangscenario, as it leads naturally to introduce a phase of growing curvature, and is acrucial ingr<strong>ed</strong>ient for the ‘bell-like’ scenario of figure 16.3.It should be not<strong>ed</strong> that pure scale factor duality, by itself, is not enough toconvert a phase of decreasing into growing curvature (see for instance figure 16.2,where it is clearly shown that H and ˜H, in the same temporal range, lead tothe same evolution of the curvature scale, H 2 ∼ ˜H 2 ). Time reflection is thusnecessarily requir<strong>ed</strong>, if we want to invert the curvature behaviour. From this pointof view, time-reversal symmetry is more important than duality.In a thermodynamic context, however, duality by itself is able to suggestthe existence of a primordial cosmological phase with ‘specular’ properties withrespect to the present, standard cosmological phase [18]. It must be stress<strong>ed</strong>, inaddition, that it is typically in the cosmology of extend<strong>ed</strong> objects that the phaseof growing curvature may describe accelerat<strong>ed</strong> expansion instead of contraction,and that the growth of the curvature may be regulariz<strong>ed</strong>, instead of blowingup to a singularity. For instance, it is with the string dilaton [8], or with anetwork of strings self-consistently coupl<strong>ed</strong> to the background [19], that we arenaturally lead to superinflation. Also, in quantum theories of extend<strong>ed</strong> objects,it is the minimal, fundamental length scale of the theory that is expect<strong>ed</strong> tobound the curvature, and to drive superinflation to a phase of constant, limitingcurvature [20] asymptotically approaching de Sitter, as explicitly check<strong>ed</strong> in astring theory context [21]. Duality symmetries, on the other hand, are typical ofextend<strong>ed</strong> objects (and of strings, in particular), so that it is certainly justifi<strong>ed</strong> tothink of duality as of a fundamental motivation and ingr<strong>ed</strong>ient of the pre-big bangscenario.Duality is an important symmetry of modern theoretical physics, and toconclude this section we would like to present an analogy with another veryimportant symmetry, namely supersymmetry (see table 16.1).According to supersymmetry, to any bosonic state is associat<strong>ed</strong> a fermionicpartner, and vice versa. From the existence of bosons that we know to bepresent in nature, if we believe in supersymmetry, we can pr<strong>ed</strong>ict the existenceof fermions not yet observ<strong>ed</strong>, like the photino, the gravitino, and so on.In the same way, according to duality and time-reversal, to any geometrical


Kinematics: shrinking horizons 289state with decreasing curvature is associat<strong>ed</strong> a dual partner with growingcurvature. On the other hand, our universe, at present, is in the standard postbigbang phase, with decreasing curvature. If we believe that duality has to beimplement<strong>ed</strong>, even approximately, in the course of the cosmological evolution,we can then pr<strong>ed</strong>ict the existence of a phase, in the past, characteriz<strong>ed</strong> by growingcurvature and by a typical pre-big bang evolution.16.3 Kinematics: shrinking horizonsIf we accept, at least as a working hypothesis, the possibility that our universe hadin the past a ‘dual’ complement, with growing curvature, we are l<strong>ed</strong> to the secondof the three questions list<strong>ed</strong> in section 16.1: is the kinematics of the pre-big bangphase still appropriate to solve the problems of the standard inflationary scenario?The answer is positive, but in a non-trivial way.Consider, for instance, the present cosmological phase. Today the dilaton isexpect<strong>ed</strong> to be constant, and the universe should be appropriately describ<strong>ed</strong> byEinstein equations. The gravitational part of such equations contains two types ofterms: terms controlling the geometric curvature of a space-like section, evolvingin time like a −2 , and terms controlling the gravitational kinetic energy, i.e. thespacetime curvature scale, evolving like H 2 . According to present observationsthe spatial curvature term is non-dominant, i.e.r = a−2H 2 ∼ spatial curvatureº 1. (16.18)spacetime curvatureAccording to the standard cosmological solutions, on the other hand, theabove ratio must grow in time. In fact, by putting a ∼ t β ,r ∼ȧ −2 ∼ t 2(1−β) , (16.19)so that r keeps growing both in the matter-dominat<strong>ed</strong> (β = 2/3) and in theradiation-dominat<strong>ed</strong> (β = 1/2) era. Thus, as we go back in time, r becomessmaller and smaller, and when we set initial conditions (for instance, at the Planckscale) we have to impose an enormous fine tuning of the spatial curvature term,with respect to the other terms of the cosmological equations. This is the so-call<strong>ed</strong>flatness problem.The problem can be solv<strong>ed</strong> if we introduce in the past a phase (usually call<strong>ed</strong>inflation), during which the value of r was decreasing, for a time long enough tocompensate the subsequent growth during the phase of standard evolution. It isimportant to stress that this requirement, in general, can be implement<strong>ed</strong> by twophysically different classes of backgrounds.Consider for simplicity a power-law evolution of the scale factor in cosmictime, with a power β, so that the time-dependence of r is the one given inequation (16.19). The two possible classes of backgrounds corresponding to adecreasing r are then the following:


290 Elementary introduction to pre-big bang cosmology• Class I: a ∼ t β , β>1, t →+∞. This class of backgrounds correspondsto what is conventionally call<strong>ed</strong> ‘power inflation’, describing a phase ofaccelerat<strong>ed</strong> expansion and decreasing curvature scale, ȧ > 0, ä > 0,Ḣ < 0. This class contains, as a limiting case, the standard de Sitterinflation, β →∞, a ∼ e Ht , Ḣ = 0, i.e. accelerat<strong>ed</strong> exponential expansionat constant curvature.• Class II: a ∼ (−t) β , β < 1, t → 0 − . This is the class of backgroundscorresponding to the string cosmology scenario. There are two possiblesubclasses:IIa: β < 0, describing superinflation, i.e. accelerat<strong>ed</strong> expansion withgrowing curvature scale, ȧ > 0, ä > 0, Ḣ > 0;IIb: 0


and the S-frame action (16.20) becomes∫S(g,φ)=−Kinematics: shrinking horizons 291d d+1 x ad e −φN [2dFH − 2d Ḣ − d(d + 1)H 2 + ˙φ 2 ]. (16.24)Modulo a total derivative, we can eliminate the first two terms, and the actiontakes the quadratic form∫S(g,φ)=− d d+1 x ad e −φN [ ˙φ 2 − 2dH ˙φ + d(d − 1)H 2 ]. (16.25)where, as expect<strong>ed</strong>, N plays the role of a Lagrange multiplier (no kinetic term inthe action).In the E-frame the variables are Ñ, ã, ˜φ, and the action (16.21), afterintegration by parts, takes the canonical form∫ [S( ˜g, ˜φ) =− d d+1 x ãd − 1 ]˙˜φ 2 + d(d − 1)H 2 . (16.26)Ñ 2A quick comparison with equation (16.25) finally leads to the field r<strong>ed</strong>efinition(no coordinate transformation!) connecting the Einstein and string frame:√2ã = ae −φ/(d−1) , Ñ = Ne −φ/(d−1) , ˜φ = φd − 1 . (16.27)In fact, the above transformation gives˜H = H −˙φ(16.28)d − 1and, when insert<strong>ed</strong> into equation (16.26), exactly reproduces the S-frame action(16.25).Consider now a superinflationary, pre-big bang solution obtain<strong>ed</strong> in the S-frame, for instance the isotropic, d-dimensional vacuum solutiona = (−t) −1/√d , e φ = (−t) −(√ d+1) , t < 0, t → 0 − (16.29)(see appendix B, equations (16.160) and (16.161)), and look for the correspondingE-frame solution. The above solution is valid in the synchronous gauge, N = 1,and if we choose, for instance, the synchronous gauge also in the E-frame, we canfix Ñ by the condition:which defines the E-frame cosmic time, ˜t, as:Ñ dt ≡ Ne −φ/(d−1) dt = d˜t, (16.30)d˜t = e −φ/(d−1) dt. (16.31)


292 Elementary introduction to pre-big bang cosmologyAfter integrationand the transform<strong>ed</strong> solution takes the form:t ∼ ˜t d−1d+ √ d , (16.32)ã = (−˜t) 1/d ,e ˜φ = (−˜t) − √ 2(d−1)d, ˜t < 0, ˜t → 0 − . (16.33)One can easily check that this solution describes accelerat<strong>ed</strong> contraction withgrowing dilaton and growing curvature scale:dãd˜t< 0,d 2 ãd˜t 2 < 0,d ˜Hd˜t< 0,d ˜φd˜t> 0. (16.34)The same result applies if we transform other isotropic solutions from the stringto the Einstein frame, for instance the perfect fluid solution of appendix C,equation (16.216). We leave this simple exercise to the interest<strong>ed</strong> reader.Having discuss<strong>ed</strong> the ‘dynamical’ equivalence (in spite of the kinematicaldifferences) of the two classes of string cosmology metrics, IIa and IIb, it seemsappropriate at this point to stress the main dynamical difference between standardinflation, class I metrics, and pre-big bang inflation, class II metrics. Such adifference can be conveniently illustrat<strong>ed</strong> in terms of the proper size of the eventhorizon, relative to a given comoving observer.Consider in fact the proper distance, d e (t), between the surface of the eventhorizon and a comoving observer, at rest at the origin of an isotropic, conformallyflat background [22]:∫ tMd e (t) = a(t) dt ′ a −1 (t ′ ). (16.35)tHere t M is the maximal allow<strong>ed</strong> extension, towards the future, of the cosmic timecoordinate for the given background manifold. The above integral converges forall the above classes of accelerat<strong>ed</strong> (expanding or contracting) scale factors. Inthe case of class I metrics we have, in particular,∫ ∞d e (t) = t β dt ′ t ′−β =tβ − 1 ∼ H −1 (t) (16.36)tfor power-law inflation (β >1, t > 0), and∫ ∞d e (t) = e Ht dt ′ e −Ht′ = H −1 (16.37)for de Sitter inflation. For class II metrics (β


Kinematics: shrinking horizons 293standardevolutionade SitterinflationCONSTANT HORIZONFigure 16.4. Qualitative evolution of the Hubble horizon (broken curve) and of the scalefactor (full curve) in the standard inflationary scenario.In all cases the proper size d e (t) evolves in time like the so-call<strong>ed</strong> Hubble horizon(i.e. the inverse of the Hubble parameter), and then like the inverse of the curvaturescale. The size of the horizon is thus constant or growing in standard inflation(class I), decreasing in pre-big bang inflation (class II), both in the S-frame andin the E-frame.Such an important difference is clearly illustrat<strong>ed</strong> in figures 16.4 and 16.5,where the broken curves represent the evolution of the horizon, the full curves theevolution of the scale factor. The shad<strong>ed</strong> area at time t 0 represents the portion ofuniverse inside our present Hubble radius. As we go back in time, according tothe standard scenario, the horizon shrinks linearly, (H −1 ∼ t), but the decreaseof the scale factor is slower so that, at the beginning of the phase of standardevolution (t = t 1 ), we end up with a causal horizon much smaller than the portionof universe that we presently observe. This is the well-known ‘horizon problem’of the standard scenario.In figure 16.4 the phase of standard evolution is prec<strong>ed</strong><strong>ed</strong> in time by a phaseof standard de Sitter inflation. Going back in time, for t < t 1 , the scale factorkeeps shrinking, and our portion of universe ‘re-enters’ inside the Hubble radiusduring a phase of constant (or slightly growing in time) horizon.In figure 16.5 the standard evolution is prec<strong>ed</strong><strong>ed</strong> in time by a phase of prebigbang inflation, with growing curvature. The universe ‘re-enters’ the Hubbleradius during a phase of shrinking horizon. To emphasize the difference, we haveplott<strong>ed</strong> the evolution of the scale factor both in the expanding S-frame, a(t), andin the contracting E-frame, ã(t). Unlike in standard inflation, the proper size of


294 Elementary introduction to pre-big bang cosmologystandardevolutionapre-big banginflationaãSHRINKING HORIZONFigure 16.5. Qualitative evolution of the Hubble horizon (broken curve) and of the scalefactor (full curve) in the pre-big bang inflationary scenario, in the S-frame, a(t), and in theE-frame, ã(t).the initial portion of the universe may be very large in strings (or Planck) units,but not larger than the initial horizon itself [23], as emphasiz<strong>ed</strong> in the picture.The initial horizon is large because the initial curvature scale is small, in stringunits, H i ≪ 1/λ s .This is a basic consequence of the choice of the initial state which, in thepre-big bang scenario, approaches the flat, cold and empty string perturbativevacuum [8], and which is to be contrast<strong>ed</strong> to the extremely curv<strong>ed</strong>, hot anddense initial state of the standard scenario, characterizing a universe which startsinflating at the Planck scale, H i ∼ 1/λ p .16.4 Open problems and phenomenological consequencesIn order to give an honest presentation of the pre-big bang scenario, it is fairto say that the string cosmology models are not free from various (more or lessimportant) difficulties, and that many aspects of the scenario are still unclear. Adetail<strong>ed</strong> discussion of such aspects is outside the purpose of this paper, but wewould like to mention here at least three important open problems. Present<strong>ed</strong>in ‘time-order<strong>ed</strong>’ form (from the beginning to the end of inflation) they are thefollowing.• The first concerns the initial conditions, and in particular the decay of thestring perturbative vacuum. The question is whether or not the ‘switching on’


Open problems and phenomenological consequences 295of a long inflationary phase requires fine tuning. Originally rais<strong>ed</strong> in [24],this problem was recently repropos<strong>ed</strong> as a fundamental difficulty of the prebigbang scenario [25] (see [23, 26–28]).• The second concerns the transition from the pre- to the post-big bang phase,which is expect<strong>ed</strong> to occur in the high curvature and strong coupling regime.There is a quantum cosmology approach, bas<strong>ed</strong> on the scattering of theWheeler–De Witt wavefunction in minisuperspace [7], but the problemseems to require, in general, the introduction of higher derivative (α ′ ) andquantum loop corrections [21,29] in the string effective action (see [15,16]).• The third problem concerns the final matching to the standard Fri<strong>ed</strong>man–Robertson–Walker phase, with a transition from the dilaton-dominat<strong>ed</strong> tothe radiation-dominat<strong>ed</strong> regime, and all the associat<strong>ed</strong> problems of dilatonoscillations, reheating, preheating, particle production, entropy production[30], and so on.All these problems are under active investigation, and further work iscertainly ne<strong>ed</strong><strong>ed</strong> for a final answer. However, even assuming that all the problemswill be solv<strong>ed</strong> in a satisfactory way, we are left eventually with a further question,the third one list<strong>ed</strong> in the introduction, which is the basic question (in ouropinion). Are there phenomenological consequences that can discriminate stringcosmology from the other inflationary scenarios? and, in particular, are suchconsequences observable (at least in principle)?The answer is positive. There are many phenomenological differences, evenif all the differences seem to have the same ‘common denominator’, i.e. the factthat the quantum fluctuations of the background fields are amplifi<strong>ed</strong> in differentmodels with different spectra. The spectrum, in particular, tends to follow thebehaviour of the curvature scale during the phase of inflation. In the standardscenario the curvature is constant or decreasing, so that the spectrum tends to beflat, or decreasing with frequency. In string cosmology the curvature is growing,and the spectrum tends to grow with frequency.In the following sections we will discuss in detail this effect for thecase of tensor metric perturbations. Here we would like to note that thephenomenological consequences of the pre-big bang scenario can be classifi<strong>ed</strong>into three different types, depending on the possibility of their observation: typeI effects, referring to observations to be perform<strong>ed</strong> in a not so far future (20–30 years?); type II effects, referring to observations to be perform<strong>ed</strong> in a nearfuture (a few years); type III effects, referring to observations already (in part)perform<strong>ed</strong>. To conclude this very quick presentation of the pre-big bang scenario,let me give one example for each type of phenomenological effect.• Type I: the production of a relic graviton background that, in the frequencyrange of conventional detectors (∼10 2 –10 3 Hz), is much higher (by 8–9orders of magnitude) than the background expect<strong>ed</strong> in conventional inflation[8, 31–33]. The sensitivity of the presently operating gravitational antennaeis not enough to detect it, however, and we have to wait for the advanc<strong>ed</strong>,


296 Elementary introduction to pre-big bang cosmologysecond generation interferometric detectors (LIGO [34], VIRGO [35]), orfor interferometers in space (LISA [36]).• Type II: the large-scale CMB anisotropy ‘se<strong>ed</strong><strong>ed</strong>’ by the inhomogeneousfluctuations of a massless [37] or massive [38] axion background. Metricfluctuations are inde<strong>ed</strong> too small, on the horizon scale, to be responsible forthe temperature anisotropies detect<strong>ed</strong> by COBE [39]; the axion spectrum,in contrast, can be sufficiently flat [40] for that purpose. Such a differentorigin of the anisotropy may lead to non-Gaussianity, or to differences (withrespect to the standard inflationary scenario) in the height and position ofthe first Doppler peak of the spectrum [41]. Such differences could be soonconfirm<strong>ed</strong>, or disprov<strong>ed</strong>, by the plann<strong>ed</strong> satellite observations (MAP [42],PLANCK [43], ...).• Type III: the production of primordial magnetic fields strong enough to‘se<strong>ed</strong>’ the galactic dynamo, and to explain the origin of the cosmic magneticfields observ<strong>ed</strong> on a large (galactic, intergalactic) scale [44]. In the standardinflationary scenario, in fact, the amplification of the vacuum fluctuationsof the electromagnetic field is not efficient enough [45], because of theconformal invariance of the Maxwell equations. In string cosmology, incontrast, the electromagnetic field is also coupl<strong>ed</strong> to the dilaton, and thefluctuations are amplifi<strong>ed</strong> by the accelerat<strong>ed</strong> growth of the dilaton during thephase of pre-big bang evolution.Finally, we wish to mention a further important phenomenogical effect,typical of string cosmology (and that we do not know how to classify withinthe tree types defin<strong>ed</strong> above, however): dilaton production, i.e. the amplificationof the dilatonic fluctuations of the vacuum, and the formation of a cosmicbackground of relic dilatons [46].The possibility of detecting such a background is strongly dependent onthe value of the dilaton mass, that we do not know, at present. If dilatons aremassless [47], then the amplitude and the spectrum of the relic background shouldbe very similar to those of the graviton background, and the relic dilatons couldbe possibly detect<strong>ed</strong>, in the future, by gravitational antennae able to respond toscalar modes, unless their coupling to bulk matter is too small [47], of course.If dilatons are massive, the mass has to be large enough to be compatible withexisting tests of the equivalence principle and of macroscopic gravitational forces.In addition, there is a rich phenomenology of cosmological bounds, which leavesopen only two possible mass windows [46]. Interestingly enough, however, in theallow<strong>ed</strong> light mass sector the dilaton lifetime is longer than the present age of theuniverse, and the dilaton fraction of critical energy density ranges from 0.01–1: inthis context, the dilaton becomes a new, interesting dark matter candidate (see [48]for a detail<strong>ed</strong> discussion of the allow<strong>ed</strong> mass windows, and of the possibilitythat light but non-relativistic dilatons could represent today a significant fractionof dark matter on a cosmological scale). We have no idea, however, of how todetect directly such a massive dilaton background, because the mass is light, but


Cosmological perturbation theory 297it is heavy enough (²10 −4 eV) to be far outside the sensitivity range of resonantgravitational detectors.The rest of this lecture will be devot<strong>ed</strong> to discussing various theoreticaland phenomenological aspects of graviton production, in a general cosmologicalcontext and, in particular, in the context of the pre-big bang scenario. Let us startby recalling some basic notions of cosmological perturbation theory, which arerequir<strong>ed</strong> for the computation of the graviton spectrum.16.5 Cosmological perturbation theoryThe standard approach to cosmological perturbation theory is to start with a setof non-perturb<strong>ed</strong> equations, for instance the Einstein or the string cosmologyequations,G µν = T µν , (16.39)to expand the metric and the matter fields around a given background solution,g µν → g (0)µν + δ(1) g µν , T µν → T (0)µν + δ(1) T µν , G (0)µν = T (0)µν , (16.40)and to obtain, to first order, a lineariz<strong>ed</strong> set of equations describing the classicalevolution of perturbations,δ (1) G µν = δ (1) T µν . (16.41)In principle, the proc<strong>ed</strong>ure is simple and straightforward. In practice, however,we have to go through a series of formal steps, that we list here in ‘chronological’order:• choice of the ‘frame’;• choice of the ‘gauge’;• normalization of the amplitude;• computation of the spectrum.16.5.1 Choice of the frameThe choice of the frame is that of the basic set of fields (metric includ<strong>ed</strong>) us<strong>ed</strong>to parametrize the action. The action, in general, can be express<strong>ed</strong> in terms ofdifferent fields. In string cosmology, for instance, there is a preferr<strong>ed</strong> frame, theS-frame, in which the lowest order gravidilaton action takes the form of (16.20).It is preferr<strong>ed</strong> because the metric appearing in the action is the same as the σ -model metric to which test strings are minimally coupl<strong>ed</strong> (see appendix A): withrespect to this metric, the motion of free strings is then geodesics. It is alwayspossible, however, through the field r<strong>ed</strong>efinition√2˜g µν = g µν e −2φ/(d−1) , ˜φ = φ, (16.42)d − 1


298 Elementary introduction to pre-big bang cosmologyto introduce the more conventional E-frame (16.21) in which the dilaton isminimally coupl<strong>ed</strong> to the metric, with a canonical kinetic term.In the two frames the field equations are different, and the perturbationequations are also different. This seems to raise a potential problem: which frameis to be us<strong>ed</strong> to evaluate the physical effects of the cosmological perturbations?The problem is only apparent, however, because physical observables (likethe perturbation spectrum) are the same in both frames. The reason is that thereis a compensation between the different perturbation equations and the differentbackground solution around which we expand. A general proof of this result canbe given by using the notion of canonical variable (see subsection 16.5.3). HereI will give only an explicit example for tensor perturbations in a d = 3, isotropicand spatially flat background.Let us start in the E-frame, with the background equations:R µν = 1 2 ∂ µφ∂ ν φ, (16.43)referring to the ‘tild<strong>ed</strong>’ variables of (16.42) (we will omit the ‘tilde’, forsimplicity, and will explicitly reinsert it at the end of the computation).Considering the transverse-traceless part of metric perturbations:δ (1) φ = 0, δ (1) g µν = h µν , δ (1) g µν =−h µν , ∇ ν h µ ν = 0 = h µ ν (16.44)(∇ µ denotes covariant differentiation with respect to the unperturb<strong>ed</strong> metric g,and the indices of h are also rais<strong>ed</strong> and lower<strong>ed</strong> with g). The perturbation of thebackground equations gives:δ (1) R ν µ = 0. (16.45)We can then work in the synchronous gauge, whereTo first order in h we getg 00 = 1, g 0i = 0, g ij =−a 2 δ ij ,h 00 = 0, h 0i = 0, g ij h ij = 0, ∂ j h j i = 0. (16.46)δ (1) Ɣ j 0i = 1 j 2ḣi , δ (1) Ɣ 0 ij =− 1 2 ḣij,δ (1) Ɣ k ij = 1 2 (∂ i h k j + ∂ j h k i − ∂ k h ij ). (16.47)The (0, 0) component of equation (16.45) is trivially satisfi<strong>ed</strong> (as well as theperturbation of the scalar field equation); the (i, j) components, by using theidentities (see for instance [54])g jk ḣ ik = ḣ j i + 2Hh j i ,g jk ḧ ik = ḧ j i + 2Ḣh j i + 4H ḣ j i + 4H 2 h j i , (16.48)giveδ (1) R j i =− 1 ()ḧ j i + 3H ḣ j i − ∇22a 2 h i j ≡− 1 2 £h i j = 0. (16.49)


Cosmological perturbation theory 299In terms of the conformal time coordinate, dη = dt/a, this wave equation can befinally rewritten, for each polarization mode, as˜h ′′ + 2 ã′ã ˜h ′ −∇ 2 ˜h = 0, (16.50)(where we have explicitly reinsert<strong>ed</strong> the tilde, and where a prime denotesdifferentiation with respect to the conformal time, which is the same in theEinstein and in the string frame, according to equations (16.27) and (16.31)).Let us now repeat the computation in the S-frame, where the backgroundequations for the metric (equation (16.196) with no contribution fromH µνα , T µν , V and σ ) can be written explicitly asPerturbing to first order,R µ ν + g να (∂ µ ∂ α φ − Ɣ µα ρ ∂ ρ φ) = 0. (16.51)δ (1) R µ ν − (δ (1) g να Ɣ µα 0 + g να δ (1) Ɣ µα 0 ) ˙φ = 0. (16.52)The (0, 0) component, as well as the perturbation of the dilaton equation, aretrivially satisfi<strong>ed</strong>. The (i, j) components, using again the identities (16.48), leadto [31]:£h j i − ˙φḣ j i = 0. (16.53)In conformal time, and for each polarization component,)h ′′ +(2 a′a − φ′ h ′ −∇ 2 h = 0. (16.54)This last equation seems to be different from the E-frame equation (16.50).Recalling, however, the relation (16.27) between a and ã, wehave2 ã′ã = 2 a′a − φ′ , (16.55)so that we have the same equation for h and ˜h, the same solution, and the samespectrum when the solution is expand<strong>ed</strong> in Fourier modes. The perturbationanalysis is thus frame-independent, and we can safely choose the more convenientframe to compute the spectrum.16.5.2 Choice of the gaugeThe second step is the choice of the gauge, i.e. the choice of the coordinate systemwithin a given frame. The perturbation spectrum is, of course, gauge-independent,but the the perturbative analysis is not, in general. It is possible, in fact, that thevalidity of the linear approximation is broken in a given gauge, but still valid in adifferent, more appropriate gauge.


300 Elementary introduction to pre-big bang cosmologySince this effect is particularly important, let me give, in short, an explicitexample for the scalar perturbations of the metric tensor in a d = 3, isotropicand conformally flat background, in the E-frame (we will omit the tilde, forsimplicity). The perturb<strong>ed</strong> metric, in the so-call<strong>ed</strong> longitudinal gauge, dependson the two Bardeen potentials and as [49]:ds 2 = a 2 [(1 + 2) dη 2 − (1 − 2)dxi 2 ]. (16.56)By perturbing the Einstein equations (16.43), the dilaton equation, and combiningthe results for the various components, one obtains to first order that = , andthat the metric fluctuations satisfy the equation: ′′ + 6 a′a ′ −∇ 2 = 0. (16.57)We now consider the particular, exact solution of the vacuum string cosmologyequations in the E-frame,a(η) =|η| 1/2 , φ(η) =− √ 3ln|η|, η → 0 − , (16.58)corresponding to a phase of accelerat<strong>ed</strong> contraction and growing dilaton (i.e. thepre-big bang solution (16.33), written in conformal time, for d = 3). For thisbackground, the perturbation equation (16.57) becomes a Bessel equation for theFourier modes k , ′′k + 3 η ′ k + k2 k = 0, ∇ 2 k =−k 2 k , (16.59)and the asymptotic solution, for modes well outside the horizon (|kη| ≪1),ϕ k = A k ln |kη|+B k |kη| −2 (16.60)contains a growing part which blows up (∼ η −2 ) as the background approachesthe high curvature regime (η → 0 − ). In this limit the linear approximationbreaks down, so that the longitudinal gauge is not, in general, consistent withthe perturbative expansion around a homogeneous, inflationary pre-big bangbackground, as scalar inhomogeneities may become too large.In the same background (16.58) the problem is absent, however, for tensorperturbations, since their growth outside the horizon is only logarithmic. Fromequation (16.50) we have in fact the asymptotic solutionh k = A k + B k ln |kη|, |kη| ≪1. (16.61)This may suggest that the breakdown of the linear approximation, for scalarperturbations, is an artefact of the longitudinal gauge. This is inde<strong>ed</strong> confirm<strong>ed</strong> bythe fact that, in a more appropriate off-diagonal (also call<strong>ed</strong> ‘uniform curvature’[50]) gauge,ds 2 = a 2 [(1 + 2ϕ)dη 2 − dx 2 i − 2∂ i B dx i dη], (16.62)


Cosmological perturbation theory 301Table 16.2. The four classes of accelerat<strong>ed</strong> backgrounds.α 0, ä > 0, Ḣ < 0α =−1 de Sitter ȧ > 0, ä > 0, Ḣ = 0−1 0, Ḣ > 0α>0 Accelerat<strong>ed</strong> contraction ȧ < 0, ä < 0, Ḣ < 0the growing mode is ‘gaug<strong>ed</strong> down’, i.e. it is suppress<strong>ed</strong> enough to restore thevalidity of the linear approximation [51] (the off-diagonal part of the metricfluctuations remains growing, but the growth is suppress<strong>ed</strong> in such a way thatthe amplitude, normaliz<strong>ed</strong> to the vacuum fluctuations, keeps smaller than one forall scales k, provid<strong>ed</strong> the curvature is smaller than one in string units). This resultis also confirm<strong>ed</strong> by a covariant and gauge invariant computation of the spectrum,according to the formalism develop<strong>ed</strong> by Bruni and Ellis [52].It should be stress<strong>ed</strong>, however, that the presence of a growing mode, and thene<strong>ed</strong> for choosing an appropriate gauge, is a problem typical of the pre-big bangscenario. In fact, let us come back to tensor perturbations, in the E-frame: for ageneric accelerat<strong>ed</strong> background the scale factor can be parametriz<strong>ed</strong> in conformaltime with a power α, as follows:a = (−η) α , η → 0 − , (16.63)and the perturbation equation (16.50) gives, for each Fourier mode, the Besselequationh ′′k + 2α η h′ k + k2 h k = 0, (16.64)with asymptotic solution, outside the horizon (|kη| ≪1):∫ ηdη ′h k = A k + B ka 2 (η ′ ) = A k + B k |η| 1−2α . (16.65)The solution tends to be constant for α1/2.It is now an easy exercise to re-express the scale factor (16.63) in cosmic time,dt = adη, a(t) ∼|t| α/(1+α) , (16.66)and to check that, by varying α, we can parametrize all types of accelerat<strong>ed</strong>backgrounds introduc<strong>ed</strong> in section 16.3: accelerat<strong>ed</strong> expansion (with decreasing,constant and growing curvature), and accelerat<strong>ed</strong> contraction, with growingcurvature (see table 16.2).In the standard, inflationary scenario the metric is expanding, α


302 Elementary introduction to pre-big bang cosmologycontraction is fast enough, i.e. α > 1/2 (in fact, the growing mode problemwas first point<strong>ed</strong> out in the context of Kaluza–Klein inflation and dynamicaldimensional r<strong>ed</strong>uction [53], where the internal dimensions are contracting). Forthe low-energy string cosmology background (16.58) we have α = 1/2, thegrowth is simply logarithmic (see (16.61)), and the linear approximation canbe appli<strong>ed</strong> consistently, provid<strong>ed</strong> the curvature remains bound<strong>ed</strong> by the stringscale [51]. However, for α > 1/2 the growth of the amplitude may require adifferent gauge for a consistent lineariz<strong>ed</strong> description.16.5.3 Normalization of the amplitudeThe lineariz<strong>ed</strong> equations describing the classical evolution of perturbations canbe obtain<strong>ed</strong> in two ways:• by perturbing directly the background equations of motion;• by perturbing the metric and the matter fields to first order, by expanding theaction up to terms quadratic in the first order fluctuations,g → g + δ (1) g, δ (2) S ≡ S[(δ (1) g) 2 ], (16.67)and then by varying the action with respect to the fluctuations.The advantage of the second method is to define the so-call<strong>ed</strong> ‘normal modes’for the oscillation of the system {gravity + matter sources}, namely the variableswhich diagonalize the kinetic terms in the perturb<strong>ed</strong> action, and satisfy canonicalcommutation relations when the fluctuations are quantiz<strong>ed</strong>. Such canonicalvariables are requir<strong>ed</strong>, in particular, to normalize perturbations to a spectrumof quantum, zero-point fluctuations, and to study their amplification from thevacuum state up to the present state of the universe.Let us apply such a proc<strong>ed</strong>ure to tensor perturbations, in the S-frame, for ad = 3 isotropic background. In the syncronous gauge, the transverse-traceless,first-order metric perturbations h µν = δ (1) g µν satisfy equation (16.46). Weexpand all terms of the low-energy gravidilaton action (16.20) up to order h 2 :δ (1) g µν =−h µν , δ (2) g µν = h µα h ν α ,δ (1)√ −g = 0, δ (2)√ −g =− 1 √4 −ghµν h µν , (16.68)and so on for δ (1) R µν , δ (2) R µν (see, for instance, [54]). By using the backgroundequations, and integrating by part, we finally arrive at the quadratic actionδ (2) S = 1 ∫ ()d 4 xa 3 e −φ ḣ ji4ḣi j + h j ∇ia 2 hi j . (16.69)By separating the two physical polarization modes, i.e. the standard ‘cross’ and‘plus’ gravity wave components,h ji hi j = 2(h2 + + h2 × ), (16.70)


Cosmological perturbation theory 303we get, for each mode (now generically denot<strong>ed</strong> with h), the effective scalar actionδ (2) S = 1 ∫ (d 4 xa 3 e −φ ḣ 2 + h ∇ )2a 2 h , (16.71)which can be rewritten, using conformal time, as∫δ (2) S = 1 2d 3 x dη a 2 e −φ (h ′2 + h∇h). (16.72)The variation with respect to h gives finally equation (16.54), i.e. the sameequation obtain<strong>ed</strong> by perturbing directly the background equations in the S-frame.The above action describes a scalar field h, non-minimally coupl<strong>ed</strong> to a tim<strong>ed</strong>ependentexternal field, a 2 e −φ (also call<strong>ed</strong> ‘pump field’). In order to impose thecorrect quantum normalization to vacuum fluctuations, we introduce now the socall<strong>ed</strong>‘canonical variable’ ψ, defin<strong>ed</strong> in terms of the pump field asψ = zh, z = ae −φ/2 . (16.73)With such a definition the kinetic term for ψ appears in the standard canonicalform: for each mode k, in fact, we get the actionδ (2) S k = 1 ∫ ()dη ψ k′22− k2 ψk 2 + z′′z ψ2 k , (16.74)and the corresponding canonical evolution equation:ψ ′′k + [k2 − V (η)]ψ k = 0,V (η) = z′′z , (16.75)which has the form of a Schrodinger-like equation, with an effective potentialdepending on the external pump field. This form of the canonical equation, bythe way, is the same for all types of perturbations (with different potentials, ofcourse). What is important, in our context, is that for an accelerat<strong>ed</strong> inflationarybackground V (z) → 0asη → −∞. This means that, asymptotically, thecanonical variable satisfies the free-field oscillating equationη →−∞, ψ ′′k + k2 ψ k = 0, (16.76)and can be normaliz<strong>ed</strong> to an initial vacuum fluctuation spectrum,η →−∞, ψ k = 1 √2ke −ikη , (16.77)in such a way as to satisfy the free field canonical commutation relations,[ψ k ,ψ ∗′j] = iδ kj . The normalization of ψ k then fixes the normalization of themetric variable h k = ψ k /z.


304 Elementary introduction to pre-big bang cosmologyIt is important to stress that there is no ne<strong>ed</strong> to introduce the canonicalvariable to study the classical evolution of perturbations, but that such variableis ne<strong>ed</strong><strong>ed</strong> for the initial normalization to a vacuum fluctuation spectrum. Wecan also normalize perturbation in a different way, of course but in that case weare studying the amplification not of the vacuum fluctuations, but of a differentspectrum [55].At this point, two remarks are in order. The first concerns the frameindependenceof the spectrum. The above proc<strong>ed</strong>ure can also be appli<strong>ed</strong> in theE-frame, to define a canonical variable ˜ψ: one then obatins for ˜ψ k the canonicalequation (16.75), with a pump field that depends only on the metric, ˜z = ã.However, by using the conformal transformation connecting the two frames, itturns out that the two pump fields are the same, ˜z = ã = ae −φ/2 = z, so thatfor ψ and ˜ψ we have the same potential, the same evolution equation, the samesolution, and thus the same spectrum.The second remark is that the canonical proc<strong>ed</strong>ure can be appli<strong>ed</strong> to anyaction, and in particular to the string effective action including higher curvaturecorrections of order α ′ , which can be written as [21]:∫S = d 4 x √ {}−ge −φ −R − ∂ µ φ∂ µ φ + α′4 [R2 GB − (∂ µφ∂ µ φ) 2 ] (16.78)where RGB2 ≡ R2 µναβ − 4R2 µν + R2 is the Gauss–Bonnet invariant (we havechosen a convenient field r<strong>ed</strong>efinition that removes terms with higher-than-secondderivatives from the equations of motion, see appendix A). From the quadraticperturb<strong>ed</strong> action we obtain α ′ corrections to the pump fields. The canonicalequation turns out to be the same as before, but with a k-dependent effectivepotential [54], and such an equation can be us<strong>ed</strong> to estimate the effects of thehigher curvature corrections on the amplification of tensor perturbations. Anumerical integration [54], in which the metric fluctuations are expand<strong>ed</strong> aroundthe high-curvature background solution of [21], leads in particular to the resultsillustrat<strong>ed</strong> in figure 16.6.The qualitative behaviour is similar, both with and without α ′ correctionsin the perturb<strong>ed</strong> equations: the fluctuations are oscillating inside the horizon andfrozen outside the horizon, as usual. However, the final amplitude is enhanc<strong>ed</strong>when α ′ corrections are includ<strong>ed</strong>, and this suggests that the energy spectrum ofthe gravitational radiation, comput<strong>ed</strong> with the low-energy perturbation equation,may represent a sort of lower bound on the total amount of produc<strong>ed</strong> gravitons.16.5.4 Computation of the spectrumThe final, amplifi<strong>ed</strong> perturbation spectrum is to be obtain<strong>ed</strong> from the solutionsof the canonical equation (16.75). In order to solve such an equation wene<strong>ed</strong> explicitly the effective potential V [z(η)] which, in general, vanishesasymptotically at large positive and negative values of the conformal time.Consider, for instance, the tensor perturbation equation in the E-frame, so that


Cosmological perturbation theory 305|h(k,t)|43with α’without α’21-40 -20 20 40 60 tFigure 16.6. Amplification of tensor fluctuations, with and without the higher-curvaturecorrections includ<strong>ed</strong> in the canonical perturbation equation.the pump field is simply the scale factor. The typical cosmological background inwhich we are interest<strong>ed</strong> in should describe a transition fom an initial accelerat<strong>ed</strong>,inflationary evolution,η →−∞, a ∼|η| α , V ∼ η −2 , (16.79)to a final standard, radiation-dominat<strong>ed</strong> phase,η →+∞, a ∼ η, V = 0. (16.80)In this context, the evolution of fluctuations, initially normaliz<strong>ed</strong> as inequation (16.77), can be describ<strong>ed</strong> as a scattering of the canonical variableby an effective potential, according to the Schrödinger-like perturbationequation (16.75) (see figure 16.7).However, the differential variable in equation (16.75) is (conformal) time,not space. As a consequence, the eigenfrequencies represent (comoving) energies,not momenta. Thus, even normalizing the initial state to a positive frequencymode, as in equation (16.77), the final state is, in general, a mixture of positiveand negative frequency modes, i.e. of positive and negative energy states,η →+∞, ψ out ∼ c + e −ikη + c − e +ikη . (16.81)In a quantum field theory context, such a mixing represents a process of pairproduction from the vacuum. The coefficients c ± are the so-call<strong>ed</strong> Bogoliubovcoefficients, parametrizing a unitary transformation between |in〉 and |out〉 states.In matrix form, they connect the set of |in〉 annihilation and creation operators,{ψ in , b k , b † k } to the out ones {ψ out, a k , a † k}, as follows:a k = c + b k + c ∗ − b† −k , a† −k = c −b k + c ∗ + b† −k . (16.82)


306 Elementary introduction to pre-big bang cosmologyFigure 16.7. Scattering and amplification of the canonical variable.Thus, even starting from the vacuum,n in =〈0|b † b|0〉 =0, (16.83)we end up with a final number of produc<strong>ed</strong> pairs which is nonzero, in general,and is controll<strong>ed</strong> by the Bogoliubov coefficient c − asn out =〈0|a † a|0〉 =|c − | 2 ̸= 0. (16.84)In a second quantization approach, the amplification of perturbations canthus be seen as a process of pair production from the vacuum (or from anyotherwise specifi<strong>ed</strong> initial state), under the action of a time-dependent externalfield (the gravidilaton background, in the string cosmology case). Equivalently,the process can be describ<strong>ed</strong> as a ‘squeezing’ of the initial state [56] (thisdescription is useful to evaluate the associat<strong>ed</strong> entropy production [57]), or, in asemiclassical language, as a ‘parametric amplification’ [58] of the wavefunctionψ k , which is scatter<strong>ed</strong> by an effective potential barrier through an ‘antitunnelling’process [59]. Quite independently of the adopt<strong>ed</strong> language, the differential energydensity of the produc<strong>ed</strong> radiation, for each mode k, depends on the number ofproduc<strong>ed</strong> pairs, and can be written asd 3 kdρ k = 2kn k(2π) 3 , n k =|c − (k)| 2 . (16.85)The computation of the so-call<strong>ed</strong> energy spectrum, defin<strong>ed</strong> as the spectral energydensity per logarithmic interval of frequency,dρ kdlnk ≡ k dρ kdk = k4π 2 |c −(k)| 2 , (16.86)thus requires the computation of c − (k), and then the knowl<strong>ed</strong>ge of the asymptoticsolution of the canonical pertubation equation at large positive times.


Cosmological perturbation theory 307To give an explicit example we shall consider here a very simple modelconsisting of two cosmological phases, an initial accelerat<strong>ed</strong> evolution up to thetime η 1 , and a subsequent radiation-dominat<strong>ed</strong> evolution for η>η 1 :a ∼ (−η) α , η < η 1 ,a ∼ η, η > η 1 . (16.87)The effective potential for tensor perturbations in the E-frame, |a ′′ /a|, starts fromzero at −∞, grows like η −2 , reaches a maximum ∼ η1 −2 , and vanishes in theradiation phase. We must solve the canonical perturbation equation for ηη 1 . In the first phase the equation r<strong>ed</strong>uces to a Bessel equation:[]ψ k ′′ + k 2 α(α − 1)−η 2 ψ k = 0, (16.88)with general solution [60]ψ k =|η| 1/2 [AH (2)ν (|kη|) + BH (1)ν (|kη|)], ν =|α − 1/2|, (16.89)where H ν(1,2) are the first- and second-kind Hankel functions, of argument kη andindex ν =|α − 1/2| determin<strong>ed</strong> by the kinematics of the background. By usingthe large argument limit of the Hankel functions for η →−∞,H (2)ν (kη) ∼ 1 √ kηe −ikη , H (1)ν (kη) ∼ 1 √ kηe +ikη , (16.90)we choose initially a positive frequency mode, normalizing the solution to avacuum fluctuation spectrum,A = 1/2, B = 0. (16.91)In the second phase V = 0, and we have the free oscillating solution:ψ k = 1 √k(c + e −ikη + c − e +ikη ). (16.92)The matching of ψ and ψ ′ at η = η 1 gives now the coefficients c ± (moreprecisely, the matching would require the continuity of the perturb<strong>ed</strong> metricproject<strong>ed</strong> on a spacelike hypersurface containing η 1 , and the continuity of theextrinsic curvature of that hypersurface [61]; but in many cases these conditionsare equivalent to the continuity of the canonical variable ψ, and of its first tim<strong>ed</strong>erivative).For an approximate determination of the spectrum, which is often sufficientfor practical purposes, it is convenient to distinguish two regimes, in whichthe comoving frequency k is much higher or much lower than the frequencyassociat<strong>ed</strong> to the top of the effective potential barrier, |V (η 1 )| 1/2 ≃ η −11. In the


308 Elementary introduction to pre-big bang cosmologyfirst case, k ≫ 1/|η 1 |≡k 1 , we can approximate the Hankel functions with theirlarge argument limit, and we find that there is no particle production,|c + |≃1, |c − |≃0. (16.93)In practice, c − is not exactly zero, but is exponentially suppress<strong>ed</strong> as a functionof the frequency, just like the quantum reflection probability for a wave with afrequency well above the top of a potential step. We will neglect such an effecthere, as we are mainly interest<strong>ed</strong> in a qualitative estimate of the perturbationspectrum.In the second case, k ≪ 1/|η 1 |≡k 1 , we can use the small argument limit ofthe Hankel functions,H (2)ν ∼ a(kη 1 ) ν − ib(kη 1 ) −ν , H (1)ν ∼ a(kη 1 ) ν + ib(kη 1 ) −ν , (16.94)and we find|c + |≃|c − |≃|kη 1 | −ν−1/2 , (16.95)corresponding to a power-law spectrum:dρ kdlnk = k4π 2 |c −(k)| 2 ≃ k4 1π 2 ( kk 1) 3−2ν, k < k 1 , (16.96)with a cut-off frequency k 1 = η1 −1 controll<strong>ed</strong> by the height of the effectivepotential.For a comparison with present observations, it is finally convenient toexpress the spectrum in terms of the proper frequency, ω(t) = k/a(t), and inunits of critical energy density, ρ c (t) = 3Mp 2 H 2 (t)/8π. We then obtain th<strong>ed</strong>imensionless spectral distribution,(ω,t) =ω dρ(ω)ρ c (t) dω ≃ 83πω14 ( ) ω 3−2νMp 2 H 2 ω 1≃ g 2 1 γ (t)( ωω 1) 3−2ν, ω < ω 1 , (16.97)whereω 1 = k 1a = 1 ≃ H 1a 1aη 1 ais the maximal amplifi<strong>ed</strong> proper frequency, g 1 = H 1 /M p , and(16.98) γ (t) = ρ γρ c=(H1H) 2 (a1 ) 4(16.99)is the energy density (in critical units) of the radiation that becomes dominant att = t 1 , rescal<strong>ed</strong> down at a generic time t (today, γ (t 0 ) ∼ 10 −4 ).a


The relic graviton background 309Table 16.3. Slope of the graviton spectrum.Scale factor Bessel index Spectrumde Sitter, constant curvature α =−1 3− 2ν = 0 flatPower-inflation, decreasing curvature α0 increasingIt is important to stress that the amplitude of the spectrum is controll<strong>ed</strong>by g 1 = H 1 /M p , i.e. by the curvature scale in Planck units at the time of thetransition t 1 (a fundamental parameter of the given inflationary model). The slopeof the spectrum, 3−2ν, is instead controll<strong>ed</strong> by the kinematics of the background.In fact, it depends on the Bessel index ν which, in its turn, depends on α, the powerof the scale factor (equation (16.89). The behaviour in frequency of the gravitonspectrum, in particular, tends to follow the behaviour of the curvature scale duringthe epoch of accelerat<strong>ed</strong> evolution, see table 16.3.The standard inflationary scenario is thus characteriz<strong>ed</strong> by a flat ordecreasing graviton spectrum; in string cosmology, instead, we must expect agrowing spectrum. This has important phenomenological implications, that willbe discuss<strong>ed</strong> in the following section.16.6 The relic graviton backgroundAs discuss<strong>ed</strong> in the previous section, one of the most firm pr<strong>ed</strong>ictions of allinflationary models is the amplification of the traceless-transverse part of thequantum fluctuations of the metric tensor, and the formation of a primordial,stochastic background of relic gravitational <strong>waves</strong>, distribut<strong>ed</strong> over a quite largerange of frequencies (see [62] for a discussion of the stochastic properties of sucha background, and [63] for a possible detection of the associat<strong>ed</strong> ‘squeezing’[64]).In a string cosmology context, the expect<strong>ed</strong> graviton background has beenalready discuss<strong>ed</strong> in a number of detail<strong>ed</strong> review papers [59,65,66]. Here we willsummarize the main properties of the background pr<strong>ed</strong>ict<strong>ed</strong> in the context of thepre-big bang scenario.For a phenomelogical discussion of the spetrum, it is convenient to considerthe plane { G ,ω}. In this plane there are three main phenomenologicalconstraints:• A first constraint comes from the large scale isotropy of the CMB radiation.The degree of anisotropy measur<strong>ed</strong> by COBE imposes a bound on the energydensity of the graviton background at the scale of the present Hubble radius[67], G (ω 0 ) º 10 −10 , ω 0 ∼ 10 −18 Hz. (16.100)


310 Elementary introduction to pre-big bang cosmology• A second constraint comes from the absence of distortion of the pulsartiming-data [68], and gives the bound G (ω p ) º 10 −8 , ω p ∼ 10 −8 Hz. (16.101)• A third constraint comes from nucleosynthesis [69], which implies that thetotal graviton energy density, integrat<strong>ed</strong> over all modes, cannot exce<strong>ed</strong> theenergy density of one massless degree of fre<strong>ed</strong>om in thermal equilibrium,evaluat<strong>ed</strong> at the nucleosynthesis epoch. This gives a bound for the peakvalue of the spectrum [33],∫h 100 dlnω G (ω, t 0 ) º 0.5 × 10 −5 ,h 100 = H 0 /(100 km s −1 Mpc −1 ), (16.102)which applies to all scales.A further bound can be obtain<strong>ed</strong> by considering the production of primordialblack holes [70]. The production of gravitons, in fact, could be associat<strong>ed</strong>with the formation of black holes, whose possible evaporation, at the presentepoch, is constrain<strong>ed</strong> by a number of astrophysical observations. The absenceof evaporation imposes an indirect upper limit on the graviton background. Ina string cosmology context, however, and in the frequency range of interest forobservations, this upper limit is roughly of the same order as the nucleosynthesisbound [70].For flat or decreasing spectra it is now evident that the more constrainingbound is the low-frequency one, obtain<strong>ed</strong> from the COBE data. In the standardinflationary scenario, characteriz<strong>ed</strong> by flat or decreasing spectra (see table 16.3),the maximal allow<strong>ed</strong> graviton background can thus be plott<strong>ed</strong> as in figure 16.8.The flat spectrum corresponds to de Sitter inflation, the decreasing spectra topower inflation. The breakdown in the spectrum, around ω e q ∼ 10 −16 Hz, isdue to the transition from the radiation-dominat<strong>ed</strong> to the matter-dominat<strong>ed</strong> phase,which only affects the low-frequency part of the spectrum, namely those modesre-entering the horizon in the matter-dominat<strong>ed</strong> era. For such modes there is anadditional potential barrier in the canonical perturbation equation, which inducesan additional amplification ∼(ω e q/ω) 2 > 1, with respect to the flat de Sitterspectrum.However, this break of the spectrum is not important for our purposes.What is important is the fact that the observ<strong>ed</strong> anisotropy constrains the maximalamplitude of the spectrum. But the amplitude depends on the inflation scale, asstress<strong>ed</strong> in the previous section. From the COBE bound (16.100), impos<strong>ed</strong> on themodifi<strong>ed</strong> de Sitter spectrum at the Hubble scale, G (ω 0 , t 0 ) = g1 2 γ (t)(ω e q/ω) 2 º 10 −10 , (16.103)we thus obtain a direct constraint on the inflation scale:H 1 /M p º 10 −5 (16.104)


The relic graviton background 311Figure 16.8. Graviton spectra in the standard inflationary scenario.Figure 16.9. A minimal model of the pre-big bang background.(for power-inflation the bound is even stronger [31]).This bound applies to all models characteriz<strong>ed</strong> by a flat or decreasingspectrum. The bound can be evad<strong>ed</strong>, however, if the spectrum is growing, like inthe string cosmology context. To illustrate this point, let us consider the simplestclass of the so-call<strong>ed</strong> ‘minimal’ pre-big bang models, characteriz<strong>ed</strong> by threemain kinematic phases [32, 44]: an initial low-energy, dilaton-driven phase, aninterm<strong>ed</strong>iate high-energy ‘string’ phase, in which α ′ and loop corrections becomeimportant, and a final standard, radiation-dominat<strong>ed</strong> phase (see figure 16.9). Thetimescale η s marks the transition to the high-curvature phase, and the timescaleη 1 , characteriz<strong>ed</strong> by a final curvature of order one in string units, marks thetransition to the radiation-dominat<strong>ed</strong> cosmology.


312 Elementary introduction to pre-big bang cosmologylog 10 (Ω G )de Sitter-18 -13 -8 -3 2 7 12log 10 ( Hz)Figure 16.10. Graviton spectra in minimal pre-big bang models.By computing graviton production in this background [32] we find that thespectrum is characteriz<strong>ed</strong> by two branches: a high-frequency branch, for modescrossing the horizon (or ‘hitting’ the barrier) in the string phase, ω > ω s =(aη s ) −1 ; and a low-frequency branch, for modes crossing the horizon in the initialdilaton phase, ω


The relic graviton background 313Figure 16.11. Peak and end point of the spectrum in minimal and non-minimal models.In the minimal models the position of the peak is fix<strong>ed</strong>. Actually, whatis really fix<strong>ed</strong>, in string cosmology, is the maximal height of the peak, but notnecessarily the position in frequency, and it is not impossible, in more complicat<strong>ed</strong>non-minimal models, to shift the peak at lower frequencies.In minimal models, in fact, the beginning of the radiation phase coincideswith the end of the string phase. We may also consider models, however, in whichthe dilaton coupling gs2 = exp(φ) is still small at the end of the high curvaturephase, and the radiation era begins much later, when g s 2 ∼ 1, after an interm<strong>ed</strong>iat<strong>ed</strong>ilaton-dominat<strong>ed</strong> regime. The main difference between the two cases is that inthe second case the effective potential which amplifies tensor perturbations is nonmonotonic[59], so that there are high-frequency modes re-entering the horizonbefore the radiation era. As a consequence, the perturbation spectrum is also nonmonotonic,and the peak does not coincide any longer, in general, with the endpoint of the spectrum (see also [71]), as illustrat<strong>ed</strong> in figure 16.11.For minimal models the peak is around 100 GHz, for non-minimal modelsit could be at lower frequencies. These are good news from an experimentalpoint of view, of course, but non-minimal models seem to be less natural, atleast from a theoretical point of view. The boxes around the peak, appearingin figure 16.11, represent the uncertainty in the position of the peak due to ourpresent ignorance about the precise value of the ratio M s /M p (for the illustrativepurpose of figure 16.11, the ratio is assum<strong>ed</strong> to vary in the range 0.1–0.01). Thewavy line, in the high frequency branch of the spectrum, represents the fact thatthe spectrum associat<strong>ed</strong> with the string phase could be monotonic on average,but locally oscillating [72]. Finally, the lower strip, labell<strong>ed</strong> by δs = 99%,represents the fact that even the height of the peak could be lower than expect<strong>ed</strong>,if the produc<strong>ed</strong> gravitons have been dilut<strong>ed</strong> by some additional reheating phase,occurring well below the string scale, during the standard evolution.


314 Elementary introduction to pre-big bang cosmologylog 10( Gh 2 100)log10(/ Hz)Figure 16.12. Allow<strong>ed</strong> region for the spectrum of vacuum fluctuations in string cosmologyand in standard inflation, compar<strong>ed</strong> with other relic spectra of primordial origin.This last effect can be parametriz<strong>ed</strong> in terms of δs, which is the fraction ofpresent entropy density in radiation, due to such additional, low-scale reheating.The position of the peak then depends on δs as [33]:( ) 1/2 Msω 1 (t 0 ) ≃ T 0 (1 − δs) 1/3 ,M p( ) 2 G (ω 1 , t 0 ) ≃ 7 × 10 −5 h −2 Ms100(1 − δs) 4/3 , (16.107)M pwhere T 0 = 2.7K≃ 3.6 × 10 11 Hz. Such a dependence is not dramatic, however,because even for δs = 99% the peak keeps well above the standard inflationarypr<strong>ed</strong>iction, represent<strong>ed</strong> by the line labell<strong>ed</strong> ‘de Sitter’ in figure 16.11.Given the various theoretical uncertainties, the best we can do, at present, isto define the maximal allow<strong>ed</strong> region for the expect<strong>ed</strong> graviton background, i.e.the region spann<strong>ed</strong> by the spectrum when all its parameters are vari<strong>ed</strong>. Such aregion is illustrat<strong>ed</strong> in figure 16.12, for the phenomenologically interesting highfrequencyrange. The figure emphasizes the possible, large enhancement (of abouteight orders of magnitude) of the intensity of the background in string cosmology,with respect to the standard inflationary scenario.It may be useful to stress again the reason of such enhancement. Inthe standard inflationary scenario the graviton spectrum is decreasing, thenormalization is impos<strong>ed</strong> at low frequency, and the peak value is controll<strong>ed</strong> by


The relic graviton background 315the anisotropy of the CMB radiation, δT/T º 10 −5 . Thus, at low frequency,( ) T 2 G (t 0 ) º γ (t 0 )∼ 10 −14 . (16.108)T COBEIn string cosmology the spectrum is growing, the normalization is impos<strong>ed</strong>at high frequency, and the peak value is controll<strong>ed</strong> by the fundamental ratioM s /M p º 0.1. Thus( ) 2 Ms G (t 0 ) º γ (t 0 ) º 10 −6 . (16.109)M pThe graviton background obtain<strong>ed</strong> from the amplification of the vacuumfluctuations, in string cosmology and in standard inflation, is compar<strong>ed</strong> infigure 16.12 with other, more unconventional graviton spectra. In particular,the graviton spectrum obtain<strong>ed</strong> from cosmic strings and topological defects [73],from bubble collision at the end of a first order phase transition [74], and froma phase of parametric resonance of the inflaton oscillations [75]. Also shown infigure 16.12 is the spectrum from models of quintessential inflation [76], and athermal black body spectrum for a temperature of about one kelvin. All thesecosmological backgrounds are higher than the background expect<strong>ed</strong> from thevacuum fluctuations in standard inflation, but not in a string cosmology context.It may be interesting, at this point, to recall the expect<strong>ed</strong> sensitivities ofthe present, and near future, gravitational antennae, referr<strong>ed</strong> to the plots offigure 16.12.At present, the best direct, experimental upper bound on the energy of astochastic graviton background comes from the cross-correlation of the data ofthe two resonant bars NAUTILUS and EXPLORER [77]: G h 100 º 60, ν ≃ 907 Hz (16.110)(similar sensitivities are also reach<strong>ed</strong> by AURIGA [78]). Unfortunately, the boundis too high to be significant for the plots of figure 16.12. However, a much bettersensitivity, G ∼ 10 −4 around ν ∼ 10 3 Hz, is expect<strong>ed</strong> from the present resonantbar detectors, if the integration time of the data is extend<strong>ed</strong> to about one year. Asimilar, or slightly better sensitivity, G ∼ 10 −5 around ν ∼ 10 2 Hz, is expect<strong>ed</strong>from the first operating version of the interferometric detectors, such as LIGOand VIRGO. At high frequency, from the kilohertz to the megahertz range, apromising possibility seems to be the use of resonant electromagnetic cavities asgravity-wave detectors [79]. Work is in progress [80] to attempt to improve theirsensitivity.The present, and near future, available sensitivities of resonant bars andinterferometers, therefore, are still outside the allow<strong>ed</strong> region of figure 16.12,determin<strong>ed</strong> by the border line G h 100 ≃ 10 −6 . (16.111)


316 Elementary introduction to pre-big bang cosmologySuch sensitivities are not so far from the border, after all, but to get really insidewe have to wait for the cross-correlation of two spherical resonant-mass detectors[81], expect<strong>ed</strong> to reach G ∼ 10 −7 in the kilohertz range, or for the advanc<strong>ed</strong>interferometers, expect<strong>ed</strong> to reach G ∼ 10 −10 in the range of 10 2 Hz. At lowerfrequencies, around 10 −2 –10 −3 Hz, the space interferometer LISA [36] seems tobe able to reach very high sensitivities, up to G ∼ 10 −11 . Work is in progress,however, for a more precise computation of their sensitivity to a cosmic stochasticbackground [82].Detectors able to reach, and to cross the limiting sensitivity (16.111), couldexplore for the first time the parameter space of string cosmology and of Planckscale physics. The detection of a signal from a pre-big bang background,extrapolat<strong>ed</strong> to the gigahertz range, could give a first experimental indication onthe value of the fundamental ratio M s /M p . Even the absence of a signal, insidethe allow<strong>ed</strong> region, would be significant, as we could exclude some portion ofparameter space of the string cosmology models, obtaining in such a way directexperimental information about processes occuring at (or very near to) the stringscale.16.7 ConclusionThe conclusion of these lectures is very short and simple.There is a rich structure of stochastic, gravitational-wave backgrounds, ofcosmological origin, in the frequency range of present (or plann<strong>ed</strong> for the future)gravity-wave detectors.Among such backgrounds, the stronger one seems to be the backgroundpossibly pr<strong>ed</strong>ict<strong>ed</strong> in the context of the pre-big bang scenario, in a stringcosmology context, originating at (or very near to) the fundamental string scale.Also, the maximal pr<strong>ed</strong>ict<strong>ed</strong> intensity of the background seems to be accessibleto the sensitivity of the future advanc<strong>ed</strong> detectors.If this is the case, the future gravity wave detectors will be able to test stringtheory models, or perhaps models referring to some more fundamental unifi<strong>ed</strong>theory, such as D-brane theory, M-theory, and so on. In any case, such detectorswill give direct experimental information on Planck scale physics.Acknowl<strong>ed</strong>gmentsIt is a pleasure to thank the organizers of the Second SIGRAV School, and allthe staff of the Center ‘A. Volta’, in Villa Olmo, for their hospitality and perfectorganization of this interesting School.


Appendix A. The string effective actionAppendix A. The string effective action 317The motion of a point particle in an external gravitational field, g µν , is govern<strong>ed</strong>by the actionS =− m ∫dτ ẋ µ ẋ ν g µν (x), (16.112)2where x µ (τ) are the spacetime coordinates of the particle, and τ is an affineparameter along the particle worldline.The time evolution of a one-dimensional object like a string describes aworld-surface, or ‘world-sheet’, instead of a worldline, and the action governingits motion is given by the surface integral∫S =− M2 sdτ dσ √ −γγ ij ∂ i x µ ∂ j x ν g µν (x), (16.113)2where ∂ i ≡ ∂/∂ξ i and ξ i = (τ, σ ) are, respectively, the timelike and spacelikecoordinates on the string world-sheet (i, j = 1, 2). The coordinates x µ (τ, σ )are the fields governing the emb<strong>ed</strong>ding of the string world-sheet in the external(also call<strong>ed</strong> ‘target’) space. The parameter Ms 2 defines (in units h/2π = 1 = c)the so-call<strong>ed</strong> string tension (the mass per unit length), and its inverse defines thefundamental length scale of the theory (often call<strong>ed</strong>, for historical reasons, the α ′parameter):Ms 2 ≡ 1 λ 2 ≡ 1s 2πα ′ . (16.114)In a curv<strong>ed</strong> metric background g µν depends on x µ , and the nonlinear action(16.113) represents what is call<strong>ed</strong> a ‘σ -model’ defin<strong>ed</strong> on the string world sheet.For the point particle action (16.112) the variation with respect to x µ leadsto the well known geodesic equations of motion,ẍ µ + Ɣ αβ µẋ α ẋ β = 0. (16.115)The string equations of motion are similarly obtain<strong>ed</strong> by varying with respect tox µ the action (16.113): we get then the Euler–Lagrange equations∂ L∂ i∂(∂ i x µ ) = ∂ L∂x µ , L = √ −γγ ij ∂ i x µ ∂ j x ν g µν , (16.116)which can be written explicitly as£x µ + γ ij ∂ i x α ∂ j x β Ɣ µ αβ = 0, £ ≡ √ 1 √∂ i −γγ ij ∂ j . (16.117)−γThese equations describe the geodesic evolution of a test string in a given externalmetric. The variation with respect to γ ik imposes the so-call<strong>ed</strong> ‘constraints’, i.e.the vanishing of the world sheet stress tensor T ik ,T ij = √ 2 δS−γ δγ ij = ∂ i x µ ∂ j x ν g µν − 1 2 γ ij∂ k x µ ∂ k x µ = 0. (16.118)


318 Elementary introduction to pre-big bang cosmologyIt is important to note, at this point, that for a classical string it is alwayspossible to impose the so-call<strong>ed</strong> ‘conformal gauge’ in which the world sheetmetric is flat, γ ij = η ij . In fact, in an appropriate basis, the two-dimensionalmetric tensor can always be set in a diagonal form, γ ij = diag(a, b), and then,by using reparametrization invariance on the world sheet, b 2 dσ 2 = a 2 dσ ′2 , themetric can be set in a conformally flat form, γ ij = a 2 η ij . Since the action (16.113)is invariant under the conformal (or Weyl) transformation γ ij → 2 (ξ k )γ ij ,√ −γγij→√ 4 −2√ −γγ ij , (16.119)we can always eliminate the conformal factor a 2 in front of the Minkowski metric,by choosing = a −1 . In the conformal gauge the equations of motion (16.117)r<strong>ed</strong>uce toẍ µ − x ′′µ + Ɣ αβ µ (ẋ α + x ′α )(ẋ β − x ′β ) = 0, (16.120)where ẋ = dx/dτ, x ′ = dx/dσ , and the constraints (16.118) becomeg µν (ẋ µ ẋ ν + x ′µ x ′ν ) = 0, g µν ẋ µ x ′ν = 0. (16.121)We now come to the crucial observation which leads to the effective actiongoverning the motion of the background fields. The conformal transformation(16.119) is an invariance of classical theory. Let us require that there are no‘anomalies’, i.e. no quantum violations of this classical symmetry. By imposingsuch a constraint, we will obtain a set of differential equations to be satisfi<strong>ed</strong>by the background fields coupl<strong>ed</strong> to the string. Thus, unlike a point particlewhich does not impose any constraint on the external geometry in which it ismoving, the consistent quantization of a string gives constraints for the externalfields. The background geometry cannot be chosen arbitrarily, but must satisfy theset of equations (also call<strong>ed</strong> β-function equations) which guarantee the absenceof conformal anomalies. The string effective action us<strong>ed</strong> in this paper is theaction which reproduces such a set of equations for the background fields, andin particular for the metric.The derivation of the background equations of motion and of the effectiveaction, from the σ -model action (16.113), can be perform<strong>ed</strong> order by order byusing a perturbative expansion in powers of α ′ (inde<strong>ed</strong>, in the limit α ′ → 0the action becomes very large in natural units, so that the quantum correctionsare expect<strong>ed</strong> to become smaller and smaller). Such a proc<strong>ed</strong>ure, however, is ingeneral long and complicat<strong>ed</strong>, even to lowest order, and a detail<strong>ed</strong> derivation ofthe background equations is outside the scope of these lectures. Let us sketch herethe proc<strong>ed</strong>ure for the simplest case in which the only external field coupl<strong>ed</strong> to thestring is the metric tensor g µν .In the conformal gauge, the action (16.113) becomes:S =− 1 ∫4πα ′ d 2 ξ∂ i x µ ∂ i x ν g µν . (16.122)


Appendix A. The string effective action 319Let us formally assume a deformation of the number of world sheet dimensions,from 2 to 2 + ɛ, and perform the conformal transformation: η ij → η ij exp(ρ).Expanding we get, for small ɛ,− 14πα ′ ∫= − 14πα ′ ∫d 2+ɛ ξ e ρɛ/2 ∂ i x µ ∂ i x ν g µνd 2+ɛ ξ∂ i x µ ∂ i x ν g µν(1 + ɛ 2 ρ +···). (16.123)For ɛ → 0 the ρ dependence disappears and the classical action isconformally invariant. In order to preserve this invariance also for the quantumtheory, at the one-loop level, let us treat the σ -model as a quantum field theoryfor x µ (σ, τ), and let us consider the quantum fluctuations ˆx µ around a givenexpectation value x µ 0. For the general reparametrization invariance of the theorywe can always choose for x 0 a locally inertial frame, such that g µν (x 0 ) = η µν .By expanding the metric around x 0 , the leading corrections are of second orderin the fluctuations, because in a locally inertial frame the first derivatives of themetric (and then the Cristoffel connection) can always be set to zero (but not thecurvature). With an appropriate choice of coordinates, call<strong>ed</strong> Riemann normalcoordinates, the metric can thus be expand<strong>ed</strong> as:g µν (x) = η µν − 1 3 R µναβ(x 0 ) ˆx α ˆx β +··· (16.124)and the action for the quantum fluctuations becomes, to lowest order in thecurvature,S =− 1 ∫4πα ′ d 2+ɛ ξ[∂ i ˆx µ ∂ i ˆx µ(1 + ɛ )2 ρ− 1 ( 3 ∂ i ˆx µ ∂ i ˆx ν R µναβ (x 0 ) ˆx α ˆx β 1 + ɛ ) ]2 ρ +··· . (16.125)It must be not<strong>ed</strong> that, at the quantum level, the dependence of ρ does notdisappear in general from the action in the limit ɛ → 0, since there are oneloopterms that diverge like ɛ −1 , just to cancel the ɛ dependence and to give acontribution proportional to ρ to the effective action. By evaluating, for instance,the two-point function for the quantum operator ˆx α ˆx β , in the coincidence limitσ → σ ′ (the tadpole graph), one obtains [83]〈ˆx α (σ ) ˆx β (σ ′ )〉 σ →σ ′ ∼ η αβ∫limσ →σ ′d 2+ɛ k eik·(σ−σ ′ )k 2 ∼ η αβ ɛ −1 , (16.126)which gives the one-loop contribution to the action∫S ∼ d 2+ɛ ξ∂ i ˆx µ ∂ i ˆx ν R µν ρ. (16.127)


320 Elementary introduction to pre-big bang cosmologyThis term violates, at one loop, the conformal invariance, unless we restrict to abackground geometry satisfying the conditionR µν = 0, (16.128)which are just the usual Einstein equations in vacuum.A similar proc<strong>ed</strong>ure can be appli<strong>ed</strong> if the string moves in a richer externalbackground (not only pure gravity). Inde<strong>ed</strong>, pure gravity is not enough, as aconsistent quantum theory for clos<strong>ed</strong> bosonic strings, for instance, must containat least three massless states (besides the unphysical tachyon, remov<strong>ed</strong> bysupersymmetry) in the lowest energy level: the graviton, the scalar dilaton andthe pseudoscalar Kalb–Ramond axion. The σ -model describing the propagationof a string in such a background must thus contain the coupling to the metric, tothe dilaton φ, and to the two-form B µν =−B νµ :S =− 1 ∫4πα ′ d 2 ξ∂ i x µ ∂ j x ν ( √ −γγ ij g µν + ɛ ij B µν )− 1 ∫d 2 ξ √ −γ φ 4π2 R(2) (γ ), (16.129)where ɛ ij is the two-dimensional Levi-Civita tensor density, ɛ 12 =−ɛ 21 = 1, andR (2) (γ ) is the two-dimensional scalar curvature for the world sheet metric γ . Thecondition of conformal invariance, at the one-loop level, leads to the equationsR µν +∇ µ ∇ ν − 1 4 H µαβ H αβ ν = 0, H µνα = ∂ µ B να + ∂ ν B αµ + ∂ α B µν ,R + 2∇ 2 φ − (∇φ) 2 −12 1 H µνα 2 = 0,∇ µ (e −φ H µνα ) = 0, (16.130)which can be obtain<strong>ed</strong> by extremizing the effective actionS =− 1 ∫2λ d−1 sd d+1 x √ [|g|e −φ R + (∇φ) 2 − 1 ]12 H µνα2(16.131)(see appendix C).It should be not<strong>ed</strong> that the inclusion of the dilaton in the condition ofconformal invariance cannot be avoid<strong>ed</strong>, since the dilaton coupling in the action(16.129) breaks conformal invariance already at the classical level ( √ −γ R (2) isnot invariant under a Weyl rescaling of γ ). However, the dilaton term is of orderα ′ with respect to the other terms of the action (for dimensional reasons), so thatit is correct to sum up the classical dilaton contribution to the quantum, one-loopeffects, as they are all of the same order in α ′ . Without the dilaton, however, theworld sheet curvature density √ −γ R (2) does not contribute to the string equationsof motion, as it is a pure Eulero two-form in two dimensions (just like the Gauss–Bonnet term in four dimensions).


Appendix A. The string effective action 321Let me note, finally, that the expansion around x 0 can be continu<strong>ed</strong> to higherorders,g(x) = η + R ˆx ˆx + ∂ R ˆx ˆx ˆx + R 2 ˆx ˆx ˆx ˆx +···, (16.132)thus introducing higher curvature terms, and higher powers of α ′ , in the effectiveaction:S =− 12λ d−1 s∫d d+1 x √ |g|e −φ [R + (∇φ) 2 − α′4 R2 µναβ +···]. (16.133)At any given order, unfortunately, there is an intrinsic ambiguity in the action dueto the fact that, with an appropriate field r<strong>ed</strong>efinition of order α ′ ,g µν → g µν + α ′ (R µν + ∂ µ φ∂ ν φ +···)φ → φ + α ′ (R +∇ 2 φ +···), (16.134)we obtain a number of different actions, again of the same order in α ′ (see, forinstance, [84]). This ambiguity cannot be eliminat<strong>ed</strong> until we limit to an effectiveaction truncat<strong>ed</strong> to a given finite order.The higher curvature (or higher derivative) expansion of the effective actionis typical of string theory: it is controll<strong>ed</strong> by the fundamental, minimal lengthparameter λ s = (2πα ′ ) 1/2 , in such a way that the higher order correctionsdisappear in the point-particle limit λ s → 0. At any given order in α ′ , however,there is also the more conventional expansion in power of the coupling constantg s (i.e. the loop expansion of quantum field theory: tree-level ∼ g −2 , one-loop∼g 0 , two-loop ∼g 2 ,...). The important observation is that, in a string theorycontext, the effective coupling constant is controll<strong>ed</strong> by the dilaton. Consider,for instance, a process of graviton scattering, in four dimensions. Comparing theaction of (16.4) with the standard, gravitational Einstein action (16.1), it followsthat the effective coupling constant, to lowest order, is√8πG = λp = λ s e φ/2 (16.135)(G is the usual Newton constant). Each loop adds an integer power of the squareof the dimensionless coupling constant, which is controll<strong>ed</strong> by the dilaton asg 2 s = (λ p/λ s ) 2 = (M s /M p ) 2 = e φ . (16.136)We may thus expect, for the loop expansion of the action, the following generalscheme:∫S =− e −φ√ −g(R +∇φ 2 + α ′ R 2 +···) tree level∫ √−g(R− +∇φ 2 + α ′ R 2 +···) one-loop∫− e +φ√ −g(R +∇φ 2 + α ′ R 2 +···) two-loop... (16.137)


322 Elementary introduction to pre-big bang cosmologyUnfortunately, each term in the action, at each loop order, is multipli<strong>ed</strong> bya dilaton ‘form factor’ which is different in general for different fields andfor different orders. This difference can lead to an effective violation of theuniversality of the gravitational interactions [85] in the low-energy, macroscopicregime, and this violation can be reconcil<strong>ed</strong> with the present tests of theequivalence principle only if the dilaton is massive enough, to make short enoughthe range of the non-universal dilatonic interactions.The tree-level relation (16.136) is valid also for a higher-dimensionaleffective action, provid<strong>ed</strong> e φ represents the shift<strong>ed</strong> four-dimensional dilatonwhich includes the volume of the extra-dimensional, compact internal space, andwhich controls the grand-unification gauge coupling, α GUT , as [13]α GUT = exp〈φ〉 =(M s /M p ) 2 ∼ 0.1–0.001. (16.138)However, the relation (16.136) is no longer valid, in general, if the gaugeinteractions are confin<strong>ed</strong> in four dimensions and only gravity propagates in theextra dimensions. In that case the relation depends on the volume of the extradimensions, whose size may be allow<strong>ed</strong> to be large in Planck units [86]. Forinternal dimensions of volume V n the relation becomes, in particular,M 2 p = M2+n s V n e − , (16.139)where is the dilaton in d = 3 + n dimensions. In this case, the string massparameter could be much smaller than the value expect<strong>ed</strong> from equation (16.138),provid<strong>ed</strong> the internal volume is correspondingly larger.Appendix B. Duality symmetryNotations and conventionsIn this paper we use the metric signature (+−−−), and we define the Riemannand Ricci tensor as follows:R β µνα = ∂ µ Ɣ β να + Ɣ β µρ Ɣ ρ να − (µ ↔ ν),R να = R µ µνα .Consider the gravidilaton effective action, in the S-frame, to lowest order inα ′ and in the quantum loop expansion:S =− 1 ∫2λ d−1 d d+1 x √ |g|e −φ [R + (∇φ) 2 ]. (16.140)sFor a homogeneous, but anisotropic, Bianchi I type metric background:φ = φ(t), g 00 = N 2 (t), g ij =−a 2 i (t)δ ij, (16.141)


Appendix B. Duality symmetry 323we have (H i =ȧ i /a i ):By noting thatddt[2 e−φ(∇φ) 2 = ˙φ 2N 2 ,Ɣ 0 00 = ṄN ≡ F,R = 1 [N 2 2F ∑ i( d∏ )( ∑a kNk=1 i( d∏ )[= e−φa k 2 ∑ Nk=1 i√d∏−g = N a i ,Ɣ 0i j = H i δ ji ,H i − 2 ∑ iH i)]Ḣ i − ∑ ii=1Ɣ ij 0 = a iȧ iN 2 δ ij,( ∑) 2Hi 2 − H i]. (16.142)iḢ i − 2F ∑ H i − 2 ˙φ ∑ ( ∑) 2H i + 2 H i],iii(16.143)the action (16.140), modulo a total derivative, can be rewritten as:S =− 1 ∫ d∏2λ d−1 d d e −φ [x dt a i˙φ 2 − ∑ ( ∑) 2Hi 2 + H i − 2 ˙φ ∑ H i].sNi=1iii(16.144)We now introduce the so-call<strong>ed</strong> shift<strong>ed</strong> dilaton φ, defin<strong>ed</strong> by∫ de −φ d xd∏= a i e −φ , (16.145)λ d sfrom whichφ = φ + ∑ ln a i , ˙φ = ˙φ + ∑ H i (16.146)ii(by assuming spatial sections of finite volume, ( ∫ d d x √ |g|) t=constant < ∞,we have absorb<strong>ed</strong> into φ the constant shift − ln(λ −d ∫s d d x), requir<strong>ed</strong> to securethe scalar behaviour of φ under coordinate reparametrizations preserving thecomoving gauge). The action becomes:S =− λ ∫ (sdt e−φ ∑)˙φ2 − Hi2 . (16.147)2 NBy inverting one of the d scale factors the corresponding Hubble parameterchanges sign,a i →ã i = ai −1 , H i → ˜H i = ˙ã i dai−1= a i =−H i , (16.148)ã i dti=1i


324 Elementary introduction to pre-big bang cosmologyso that the quadratic action (16.147) is clearly invariant under the inversion of anyscale factor preserving the shift<strong>ed</strong> dilaton,a i →ã i = a −1i, φ → φ (16.149)(for ‘scale factor duality’, see [9] and the first paper of [8]).In order to derive the field equations, it is convenient to use the variablesβ i = ln a i , so that H i = ˙β i , Ḣ i = ¨β i , and the action (16.147) is cyclic in β i .By varying with respect to N,β i and φ, and subsequently fixing the cosmic timegauge N = 1, we obtain, respectively,˙φ 2 − ∑ iH 2i = 0, (16.150)Ḣ i − H i ˙φ = 0, (16.151)2 ¨φ − ˙φ 2 − ∑ iH 2i = 0. (16.152)This is a system of (d + 2) equations for the (d + 1) variables {a i ,φ}. However,only (d + 1) equations are independent (see, for instance, [21]: equation (16.150)represents a constraint on the set of initial data).The above equations are invariant under a time reversal transformationt →−t,H →−H,˙φ →−˙φ, (16.153)and also under the duality transformation (16.149). If we invert k ≤ d scalefactors, ã 1 = a1 −1 ,...,ã k = ak −1 , the shift<strong>ed</strong> dilaton is preserv<strong>ed</strong>, φ = ˜φ,provid<strong>ed</strong>from which:φ = φ −d∑ln a i = ˜φ −i=1k∑ln ã i −i=1˜φ = φ − 2d∑i=k+1ln a i , (16.154)k∑ln a i . (16.155)i=1Given an exact solution, represent<strong>ed</strong> by the set of variables{a 1 ,...,a d ,φ}, (16.156)the inversion of k ≤ d scale factors defines then a new exact solution, represent<strong>ed</strong>by the set of variables{a −11 ,...,a−1 k, a k+1 ,...,a d ,φ− 2lna 1 ,...,−2lna k }. (16.157)


Appendix B. Duality symmetry 325By inverting all the scale factors we obtain the transformation{a i ,φ}→{a −1i,φ− 2d∑ln a i } (16.158)which, in the isotropic case, corresponds in particular to the duality transformation(16.5).As a simple example, we consider here the particular isotropic solutioni=1a = t 1/√d , φ =−ln t, (16.159)which satisfies identically the set of equations (16.150)–(16.152). By applyinga duality and a time-reversal transformation we obtain the four different exactsolutions{a ± (t) = t ±1/√d , φ(t) =−ln t},{a ± (−t) = (−t) ±1/√d , φ(−t) =−ln(−t)}, (16.160)corresponding to the four branches illustrat<strong>ed</strong> in figure 16.2, and describingdecelerat<strong>ed</strong> expansion, a + (t), decelerat<strong>ed</strong> contraction, a − (t), accelerat<strong>ed</strong>contraction, a + (−t), accelerat<strong>ed</strong> expansion, a − (−t). The solution describesexpansion or contraction if the sign of ȧ is positive or negative, repectively, andthe solution is accelerat<strong>ed</strong> or decelerat<strong>ed</strong> if ȧ and ä have the same or the oppositesign, respectively.It is important to consider also the dilaton behaviour. According toequation (16.146):φ ± (±t) = φ(±t) + d ln a ± (±t) = (± √ d − 1) ln(±t). (16.161)It follows that, in a phase of growing curvature (t < 0, t → 0 − ), the dilaton isgrowing only for an expanding metric, a − (−t). This means that, in the isotropiccase, there are only expanding pre-big bang solutions, i.e. solutions evolving fromthe string perturbative vacuum (H → 0,φ →−∞), and then characteriz<strong>ed</strong> by agrowing string coupling, ġ s = (exp φ/2)˙ > 0.In the more general, anisotropic case, and in the presence of contractingdimensions, a growing curvature solution is associat<strong>ed</strong> to a growing dilaton onlyfor a large enough number of contracting dimensions. To make this point moreprecisely, consider the particular, exact solution of equations (16.150)–(16.152)with d expanding and n contracting dimensions, and scale factors a(t) and b(t),respectively:a = (−t) −1/√ d+n , b = (−t) 1/√ d+n , φ =−ln(−t), t → 0 − . (16.162)This gives, for the dilaton,φ = φ + d ln a + n ln b = n − d − √ d + n√d + nln(−t), (16.163)


326 Elementary introduction to pre-big bang cosmologyso that the dilaton is growing ifd + √ d + n > n. (16.164)For n = 6, in particular, this condition requires d > 3. This could represent apotential difficulty for the pre-big bang scenario, which might be solv<strong>ed</strong>, however,by quantum cosmology effects [87].The scale factor duality of the action (16.147) is, in general, broken by theaddition of a non-trivial dilaton potential (unless the potential depends on th<strong>ed</strong>ilaton through φ, of course). When the antisymmetric tensor B µν is includ<strong>ed</strong>in the action, however, the scale factor duality can be lift<strong>ed</strong> to a larger group ofglobal symmetry transformations. To illustrate this important aspect of the stringcosmology equations, we will consider here a set of cosmological backgroundfields {φ,g µν , B µν }, for which a synchronous frame exists where g 00 = 1,g 0i = 0, B 0µ = 0, and all the components φ,g ij , B ij do not depend on thespatial coordinates.Let us write the actionS =− 1 ∫2λ d−1 sd d+1 x √ |g|e −φ [R + (∇φ) 2 − 112 H 2 µνα](16.165)directly in the synchronous gauge, as we are not interest<strong>ed</strong> in the field equations,but only in the symmetries of the action. We set g ij =−γ ij and we find, in thisgauge,whereƔ ij 0 = 1 2 ˙γ ij,Ɣ 0i j = 1 2 g jk ġ ik = 1 2 (g−1 ġ) i j = (γ −1 ˙γ) ijR 0 0 =− 1 4 Tr(γ −1 ˙γ) 2 − 1 2 Tr(γ −1 ¨γ)− 1 2 Tr( ˙γ −1 ˙γ),R j i =− 1 2 (γ −1 ¨γ) j i − 1 4 (γ −1 ˙γ) j i Tr(γ −1 ˙γ)+ 1 2 (γ −1 ˙γγ −1 ˙γ) j i , (16.166)Tr(γ −1 ˙γ)= (γ −1 ) ij ˙γ ji = g ij ġ ji , (16.167)and so on (note also that ˙γ −1 means (γ −1 )˙).antisymmetric tensor,Similarly we find, for theH 0ij = Ḃ ij , H 0ij = g ik g jl Ḃ kl = (γ −1 Ḃγ −1 ) ij ,H µνα H µνα = 3H 0ij H 0ij =−3Tr(γ −1 Ḃ) 2 . (16.168)Let us introduce the shift<strong>ed</strong> dilaton, by absorbing into φ the spatial volume, asbefore:√| det gij |e −φ = e −φ , (16.169)from whichφ˙= ˙φ − 1 d2 dt ln(det γ)= ˙φ − 1 2 Tr(γ −1 ˙γ). (16.170)


Appendix B. Duality symmetry 327By collecting the various contributions from φ, R and H 2 , the action (16.165) canbe rewritten as:S =− λ ∫sdt e −φ [ ˙φ 2 + 124 Tr(γ −1 ˙γ) 2 − Tr(γ −1 ¨γ)− 1 2 Tr( ˙γ −1 ˙γ)+ ˙φ Tr(γ −1 ˙γ)+ 1 4 Tr(γ −1 Ḃ) 2 ]. (16.171)We can now eliminate the second derivatives, and the mix<strong>ed</strong> term (∼ ˙φ ˙γ ), bynoting thatddt [e−φ Tr(γ −1 ˙γ)] = e −φ [Tr(γ −1 ¨γ)+ Tr( ˙γ −1 ˙γ)− ˙φ Tr(γ −1 ˙γ)]. (16.172)Finally, by using the identity,(γ −1 )˙=−γ −1 ˙γγ −1 (16.173)(following from g −1 g = γ −1 γ = I), we can rewrite the action in quadratic form,modulo a total derivative, asS =− λ ∫ [sdt e −φ 1 ˙φ2 −24 Tr(γ −1 ˙γ) 2 + 1 ]4 Tr(γ −1 Ḃ) 2 . (16.174)This action can be set into a more compact form by using the (2d × 2d)matrix M, defin<strong>ed</strong> in terms of the spatial components of the metric and of theantisymmetric tensor,(G−1−GM =−1 )BBG −1 G − BG −1 ,BG = g ij ≡−γ ij , G −1 ≡ g ij , B ≡ B ij , (16.175)and using also the matrix η, representing the invariant metric of the O(d, d) groupin the off-diagonal representation,( )0 Iη =(16.176)I 0(I is the unit d-dimensional matrix). By computing Mη, Ṁη and (Ṁη) 2 we find,in fact,Tr(Ṁη) 2 = 2Tr[˙γ −1 ˙γ + (γ −1 Ḃ) 2 ] = 2Tr[−(γ −1 ˙γ) 2 + (γ −1 Ḃ) 2 ], (16.177)and the action becomesS =− λ s2∫dt e −φ [˙φ2 +18 Tr(Ṁη)2 ]. (16.178)


328 Elementary introduction to pre-big bang cosmologyWe may note, at this point, that M is a symmetric matrix of the pseudoorthogonalO(d, d) group. In fact,for any B and G. Therefore,M T ηM = η, M = M T (16.179)Mη = ηM −1 , (Ṁη) 2 = η(M −1 )˙Mη, (16.180)and the action can be finally rewritten asS =− λ ∫ [ ]sdt e −φ 1 ˙φ2 +28 Tr Ṁ(M−1 )˙ . (16.181)This form is explicitly invariant under the global O(d, d) transformations (16.10),preserving the shift<strong>ed</strong> dilaton:φ → φ, M → T M , T η = η. (16.182)In factTr ˙˜M( ˜M −1 )˙=Tr[ T Ṁ −1 (M −1 )˙( T ) −1 ] = Tr Ṁ(M −1 )˙. (16.183)In the absence of the antisymmetric tensor M is diagonal, and the special O(d, d)transformation with = η corresponds to an inversion of the metric tensor:M = diag(G −1 , G),˜M = T M = ηMη = diag(G, G −1 ) ⇒ ˜G = G −1 . (16.184)For a diagonal metric G = a 2 I, and the invariance under the scale factor dualitytransformation (16.5) is recover<strong>ed</strong> as a particular case of the global O(d, d)symmetry of the low-energy effective action.Appendix C. The string cosmology equationsIn order to derive the cosmological equations let us include in the action, forcompleteness, the antisymmetric tensor B µν , a dilaton potential V (φ), and alsothe possible contribution of other matter sources represent<strong>ed</strong> by a Lagrangiandensity L m :S = − 1 ∫d d+1 x √ [|g|e −φ R + (∇φ) 2 − 1 ]12 H µνα 2 + V (φ)2λ d−1 s∫+ d d+1 x √ |g|L m . (16.185)In a scalar–tensor model of gravity, especially in the presence of higherderivative interactions, it is often convenient to write the action in the language of


Appendix C. The string cosmology equations 329exterior differential forms, as this may simplify the variational proc<strong>ed</strong>ure (see, forinstance, [88]). Here, we will follow however the more traditional approach, byvarying the action with respect to g µν ,φ and B µν . We shall take into account th<strong>ed</strong>ynamical stress tensor T µν of the matter sources (defin<strong>ed</strong> in the usual way), aswell as the scalar source σ representing a possible direct coupling of the dilatonto the matter fields:δ g ( √ −gL m ) = 1 2√ −gTµν δg µν , δ φ ( √ −gL m ) = √ −gσδφ. (16.186)We start performing the variation with respect to the metric, using thestandard, general relativistic results:δ √ −g =− 1 2√ −ggµν δg µν ,δ( √ −gR) = √ −g(G µν δg µν + g µν ∇ 2 δg µν −∇ µ ∇ ν δg µν ), (16.187)where G µν is the usual Einstein tensor. It must be not<strong>ed</strong>, however, that the secondcovariant derivatives of δg µν , when integrat<strong>ed</strong> by parts, are no longer equivalentto a divergence (and then to a surface integral), because of the dilaton factorexp(−φ) in front of the Einstein action, which adds dilatonic gradients to thefull variation. By performing a first integration by part, and using the metricitycondition ∇ α g µν = 0, we get in fact:δ g S = 1 2∫d d+1 x √ |g|T µν δg µν − 12λ d−1 s∫d d+1 x √ |g|e −φ× [G µν +∇ α φg µν ∇ α −∇ µ φ∇ ν +∇ µ φ∇ ν φ − 1 2 g µν(∇φ) 2− 1 2 g µνV (φ) + 1 2 g µν 12 1 H αβγ 2 − 12 3 H µαβ H αβ ν ]δg µν− 1 ∫2λ d−1 d d+1 x √ |g|∇ α [e −φ g µν ∇ α δg µν − e −φ ∇ ν δg να ] = 0.s(16.188)A second integration by parts of ∇δg µν cancels the bilinear term ∇ µ φ∇ ν φ, andleads to the field equations:G µν +∇ µ ∇ ν φ + 1 2 g µν[(∇φ) 2 − 2∇ 2 φ − V (φ) + 112 H 2 αβγ ] − 1 4 H µαβ H ναβ= 1 2 eφ T µν . (16.189)We have chosen units such that 2λ d−1s = 1, so that e φ represents the (d + 1)-dimensional gravitational constant (see appendix A). Also, we have implicitlyadd<strong>ed</strong> to the action the boundary term12λ d−1 s∫∂√|g|e −φ K α d α , (16.190)


330 Elementary introduction to pre-big bang cosmologywhose variation with respect to g exactly cancels the contribution of the totaldivergence appearing in the last integral of equation (16.188):∫ √|g|e∫ √|g|eδ −φ g K α d α =−φ (g µν ∇ α δg µν −∇ ν δg να ) d α . (16.191)Here K α is a geometric term representing the so-call<strong>ed</strong> extrinsic curvature on th<strong>ed</strong>-dimensional clos<strong>ed</strong> hypersurface, of infinitesimal area d α , bounding the totalspacetime volume over which we are varying the action. Note that the integral(16.190) differs from the usual boundary term, us<strong>ed</strong> in general relativity [89]to derive the Einstein equations, only by the presence of the tree-level dilatoncoupling e −φ to the extrinsic curvature.Let us now perform the variation with respect to the dilaton, again in units2λ d−1s= 1. We get the Euler–Lagrange equations:∂ µ [−2 √ −ge −φ ∂ µ φ] = e −φ√ −g[R + (∇φ) 2 −12 1 H 2 + V ]− e −φ√ −gV ′ + √ −gσ (16.192)(where V ′ = ∂V/∂φ), from whichR + 2∇ 2 φ − (∇φ) 2 + V − V ′ − 1 12 H 2 + e φ σ = 0. (16.193)The variation with respect to B µν ,δ B∫d d+1 x √ |g|e −φ (∂ µ B να )H µνα = 0, (16.194)gives finally∂ µ ( √ |g|e −φ H µνα ) = 0 =∇ µ (e −φ H µνα ). (16.195)Equations (16.189), (16.193) and (16.195) are the equations governingthe evolution of the string cosmology background, at low energy. Note thatequation (16.189) can also be given in a simplifi<strong>ed</strong> form: if we eliminatethe scalar curvature present inside the Einstein tensor, by using the dilatonequation (16.193), we obtain:R ν µ +∇ µ ∇ ν φ − 1 2 δν µ V ′ − 1 4 H µαβ H ναβ = 1 2 eφ (T ν µ − δµ ν σ). (16.196)For the purpose of these lectures, it will be enough to derive some simplesolution of the string cosmology equations in the absence of the potential (V = 0),of the antisymmetric tensor (B = 0), and with a perfect fluid, minimally coupl<strong>ed</strong>to the dilaton (σ = 0), as the matter sources. Assuming for the background aBianchi I type metric, we can work in the synchronous gauge, by settingg µν = diag(1, −ai 2 δ ij), a i = a i (t), φ = φ(t),T ν µ = diag(ρ, −pi 2 δ ji ), p i/ρ = γ i = constant, ρ = ρ(t). (16.197)


Appendix C. The string cosmology equations 331For this background:Ɣ 0i j = H i δ ji , Ɣ ij 0 = a i ȧ i δ ij , R 0 0 =− ∑ i(Ḣ i + H 2i ),R j i =−Ḣ i δ ji− H i δ ji(∇φ) 2 = ˙φ 2 ,∑H k ,kR =− ∑ i∇ 2 φ = ¨φ + ∑ iH i ˙φ,(2Ḣ i + H 2i ) − ( ∑i∇ 0 ∇ 0 φ = ¨φ,H i) 2,∇ i ∇ j φ = H i ˙φδ ji . (16.198)The dilaton equation (16.193) gives then2 ¨φ + 2 ˙φ ∑ iH i − ˙φ 2 − ∑ i(2Ḣ i + H 2i ) − ( ∑iH i) 2= 0. (16.199)The (00) component of the equation (16.189) gives˙φ 2 − 2 ˙φ ∑ iH i − ∑ i( ∑Hi 2 +iH i) 2= e φ ρ. (16.200)The diagonal, spatial components (i, i) of equation (16.189) (the off-diagonalcomponents are trivially satisfi<strong>ed</strong>) give∑∑Ḣ i + H i H k − H i ˙φ − 1 2(2Ḣ i + Hi 2 )ki− 1 2( ∑iH i) 2− 1 2 ˙φ 2 + ¨φ + ˙φ ∑ iH i = 1 2 eφ p i . (16.201)The last five terms on the left-hand side add to zero because of the dilaton equation(16.199), and the spatial equations r<strong>ed</strong>uce to(Ḣ i − H i˙φ − ∑ )H k = 1 2 eφ p i . (16.202)kThe above equations are clearly invariant under time-reversal, t →−t. Inorder to make explicit also their duality invariance, let us introduce again theshift<strong>ed</strong> dilaton (see equation (16.146)), such thate φ = e φ / √ −g,˙φ = ˙φ − ∑ iH i , (16.203)and defineρ = ρ √ −g = ρ ∏ ia i ,p = p √ −g = p ∏ ia i . (16.204)


332 Elementary introduction to pre-big bang cosmologyIn terms of these variables, the time and space equations (16.200) and (16.202),and the dilaton equation (16.199), become, respectively:˙φ 2 − ∑ iH 2i = e φ ρ, (16.205)Ḣ i − H i ˙φ =12e φ p i , (16.206)2 ¨φ − ˙φ 2 − ∑ iH 2i = 0. (16.207)They are explicitly invariant under the scale-factor duality transformation:a i → ai −1 , φ → φ, ρ → ρ, p →−p (16.208)which implies, for a perfect fluid source, a ‘reflection’ of the equation of state,γ = p/ρ = p/ρ → −p/ρ = −γ (see the first paper in [8]). A generalO(d, d) transformation changes, however, the equation of state in a more drasticway (see [12]), introducing also shear and bulk viscosity.The above (d + 2) equations are a system of independent equations forthe (d + 2) variables {a i ,φ,ρ}. Their combination implies the usual covariantconservation of the energy density. By differentiating equation (16.205), andusing (16.206) and (16.207) to eliminate Ḣ i , ˙φ, respectively, we get in fact˙ρ + ∑ iH i p i = 0, (16.209)which, using the definitions (16.204), is equivalent to˙ρ + ∑ iH i (ρ + p i ) = 0. (16.210)In order to obtain exact solutions, it is convenient to include this energyconservation equation in the full system of independent equations.In these lectures we will present only a particular example of the matterdominat<strong>ed</strong>solution by considering a d-dimensional, isotropic backgroundcharacteriz<strong>ed</strong> by a power-law evolution,a ∼ t α , φ ∼−β ln t, p = γρ. (16.211)We use (16.205), (16.207) and (16.209) as independent equations. The integrationof equation (16.209) gives imm<strong>ed</strong>iatelyequation (16.205) is then satisfi<strong>ed</strong> provid<strong>ed</strong>ρ = ρ 0 a −dγ ; (16.212)dγα+ β = 2. (16.213)


References 333Finally, equation (16.207) provides the constraint2β − β 2 − dα 2 = 0. (16.214)We then have a system of two equations for the two parameters α, β (note that,if α is a solution for a given γ , then also −α is a solution, associat<strong>ed</strong> to −γ ).We have, in general, two solutions. The trivial flat space solution β = 2,α = 0,corresponds to dust matter (γ = 0) according to equation (16.206). For γ ̸= 0we obtain insteadα =2γ1 + dγ 2 , β = 21 + dγ 2 , (16.215)which fixes the time evolution of a and φ:2γ2a ∼ t 1+dγ 2 , φ =− ln t, (16.216)1 + dγ 2and also of the more conventional variables ρ,φ:ρ = ρa −d = ρ 0 a −d(1+γ) , φ = φ + d ln a =2(dγ − 1)ln t. (16.217)1 + dγ 2This particular solution reproduces the small curvature limit of the generalsolution with perfect fluid sources (see the last two papers of [8]), sufficientlyfar from the singularity. As in the vacuum solution (16.159) there are fourbranches, relat<strong>ed</strong> by time-reversal and by the duality transformation (16.208), andcharacteriz<strong>ed</strong> by the scale factorsa ± (±t) ∼ (±t) ±2γ/(1+dγ 2) . (16.218)The duality transformation that preserves φ and ρ, and inverts the scale factor,in this case is simply represent<strong>ed</strong> by the transformation γ → −γ . Considerfor instance the standard radiation-dominat<strong>ed</strong> solution, corresponding to d = 3,γ = 1/3, and t > 0, and associat<strong>ed</strong> to a constant dilaton, according toequation (16.217). A duality transformation gives a new solution with γ =−1/3,namely (from (16.216) and (16.217)):a ∼ t −1/2 , ρ ∼ a −2 , φ ∼−3lnt. (16.219)By performing an additional time reflection we then obtain the pre-big bangsolution ‘dual to radiation’, already report<strong>ed</strong> in eq. (16.16).References[1] Weinberg S 1972 Gravitation and Cosmology (New York: Wiley)[2] Kolb E W and Turner M S 1990 The Early Universe (R<strong>ed</strong>wood City, CA: Addison-Wesley)


334 Elementary introduction to pre-big bang cosmologyLinde A D 1990 Particle Physics and Inflationary Cosmology (Chur: Harwood)[3] Vilenkin A 1992 Phys. Rev. D 46 2355Borde A and Vilenkin A 1994 Phys. Rev. Lett. 72 3305[4] Guth A 1998 The inflationary Universe (London: Vintage) p 271[5] Vilenkin A 1984 Phys. Rev. D 30 509Linde A D 1984 Sov. Phys.–JEPT 60 211Zel’dovich Y and Starobinski A A 1984 Sov. Astron. Lett. 10 135Rubakov V A 1984 Phys. Lett. B 148 280[6] Hartle J B and Hawking S W 1983 Phys. Rev. D 28 2960Hawking S W 1984 Nucl. Phys. B 239 257Hawking S W and Page D N 1986 Nucl. Phys. B 264 185[7] Gasperini M, Maharana J and Veneziano G 1996 Nucl. Phys. B 472 394Gasperini M and Veneziano G 1996 Gen. Rel. Grav. 28 1301Gasperini M 1998 Int. J. Mod. Phys. A 13 4779Cavagliá M and Ungarelli C 1999 Class. Quantum Grav. 16 1401[8] Veneziano G 1991 Phys. Lett. B 265 287Gasperini M and Veneziano G 1993 Astropart. Phys. 1 317Gasperini M and Veneziano G 1993 Mod. Phys. Lett. A 8 3701Gasperini M and Veneziano G 1994 Phys. Rev. D 50 2519An updat<strong>ed</strong> collection of papers on the pre-big bang scenario is available athttp://www.to.infn.it/∼gasperin[9] Tseytlin A A 1991 Mod. Phys. Lett. A 6 1721[10] Sen A 1991 Phys. Lett. B 271 295Hassan S F and Sen A 1992 Nucl. Phys. B 375 103[11] Meissner K A and Veneziano G 1991 Phys. Lett. B 267 33Meissner K A and Veneziano G 1991 Mod. Phys. Lett. A 6 3397[12] Gasperini M and Veneziano G 1992 Phys. Lett. B 277 256[13] Kaplunovsky V 1985 Phys. Rev. Lett. 55 1036[14] Brustein R and Veneziano G 1994 Phys. Lett. B 329 429Kaloper N, Madden R and Olive K A 1995 Nucl. Phys. B 452 677Easther R, Ma<strong>ed</strong>a K and Wands D 1996 Phys. Rev. D 53 4247[15] Gasperini M, Maharana J and Veneziano G 1991 Phys. Lett. B 272 277[16] Gasperini M 1999 Mod. Phys. Lett. A 14 1059[17] Shadev D 1984 Phys. Lett. B 317 155Abbott R B, Barr S M and Ellis S D 1984 Phys. Rev. D 30 720Kolb E W, Lindley D and Seckel D 1984 Phys. Rev. D 30 1205Lucchin F and Matarrese S 1985 Phys. Lett. B 164 282[18] Brandenberger R and Vafa C 1989 Nucl. Phys. B 316 391Tseytlin A A and Vafa C 1992 Nucl. Phys. B 372 443[19] Gasperini M, Sánchez N and Veneziano G 1991 Nucl. Phys. B 364 365[20] Caianiello E R, Gasperini M and Scarpetta G 1991 Class. Quantum Grav. 8 659Gasperini M 1991 Proc. First Workshop on ‘Advances in Theoretical Physics’(Vietri, 1990) <strong>ed</strong> E R Caianiello (Singapore: World Scientific) p 77[21] Gasperini M, Maggiore M and Veneziano G 1997 Nucl. Phys. B 494 315[22] Rindler W 1956 Mon. Not. R. Astron. Soc. 116 6[23] Gasperini M 2000 Phys. Rev. D 61 087301[24] Turner M S and Weinberg E J 1997 Phys. Rev. D 56 4604[25] Kaloper N, Linde A and Bousso R 1999 Phys. Rev. D 59 043508


References 335[26] Maggiore M and Sturani R 1997 Phys. Lett. B 415 335[27] Buonanno A, Damour T and Veneziano G 1999 Nucl. Phys. B 543 275[28] Clancy D, Lidsey J E and Tavakol R 1999 Phys. Rev. D 59 063511[29] Antoniadis I, Rizos J and Tamvakis K 1994 Nucl. Phys. B 415 497Rey S J 1996 Phys. Rev. Lett. 77 1929Gasperini M and Veneziano G 1996 Phys. Lett. B 387 715Brustein R and Madden R 1997 Phys. Lett. B 410 110Brustein R and Madden R 1998 Phys. Rev. D 57 712Brandenberger R, Easther R and Maia J 1998 JHEP 9808 007Maggiore M and Riotto A 1999 Nucl. Phys. B 458 427Foffa S, Maggiore M and Sturani R 1999 Nucl. Phys. B 552 395[30] Gasperini M and Giovannini M 1993 Phys. Lett. B 301 334Gasperini M and Giovannini M 1993 Class. Quantum Grav. 10 L133Veneziano G 1999 Phys. Lett. B 454 22[31] Gasperini M and Giovannini M 1992 Phys. Lett. B 282 36Gasperini M and Giovannini M 1993 Phys. Rev. D 47 1519[32] Brustein R, Gasperini M, Giovannini M and Veneziano G 1995 Phys. Lett. B 36145[33] Brustein R, Gasperini M and Veneziano G 1997 Phys. Rev. D 55 3882[34] See, for instance, Fritschel P 1998 Second Amaldi Conf. on ‘<strong>Gravitational</strong> Waves’(CERN, 1997) <strong>ed</strong> E Coccia et al (Singapore: World Scientific) p 74[35] See, for instance, Brillet A 1998 Second Amaldi Conf. on ‘<strong>Gravitational</strong> Waves’(CERN, 1997) <strong>ed</strong> E Coccia et al (Singapore: World Scientific) p 86[36] See, for instance, Hough J 1998 Second Amaldi Conf. on ‘<strong>Gravitational</strong> Waves’(CERN, 1997) <strong>ed</strong> E Coccia et al (Singapore: World Scientific) p 97See also Bender P these proce<strong>ed</strong>ings[37] Durrer R, Gasperini M, Sakellariadou M and Veneziano G 1998 Phys. Lett. B 43666Durrer R, Gasperini M, Sakellariadou M and Veneziano G 1999 Phys. Rev. D 59043511[38] Gasperini M and Veneziano G 1999 Phys. Rev. D 59 043503[39] Bennet C L et al 1994 Astrophys. J. 430 423[40] Copeland E J, Easther R and Wands D 1997 Phys. Rev. D 56 874Copeland E J, Lidsey J and Wands D 1997 Nucl. Phys. B 506 407Buonanno A, Meissner K A, Ungarelli C and Veneziano G 1998 JHEP 01 004[41] Melchiorri A, Vernizzi F, Durrer R and Veneziano G 1999 Phys. Rev. Lett. 83 4464[42] See, for instance, the web site: http://map.gsfc.nasa.gov[43] See, for instance, the web site: http://astro.estec.esa.nl/SA-general/Projects/Planck[44] Gasperini M, Giovannini M and Veneziano G 1995 Phys. Rev. Lett. 75 3796[45] Turner M S and Widrow L M 1988 Phys. Rev. D 37 2743[46] Gasperini M 1994 Phys. Lett. B 327 314Gasperini M and Veneziano G 1994 Phys. Rev. D 50 2519[47] Damour T and Polyakov A M 1994 Nucl. Phys. B 423 532[48] Gasperini M 1997 Proc. 12th Int. Conf. on ‘General Relativity and <strong>Gravitational</strong>Physics’ (Rome) <strong>ed</strong> M Bassan et al (Singapore: World Scientific) p 85[49] Mukhanov V F, Feldman H A and Brandenberger R H 1992 Phys. Rep. 215 203[50] Hwang J 1991 Ap. J. 375 443[51] Brustein R, Gasperini M, Giovannini M, Mukhanov V and Veneziano G 1995 Phys.


336 Elementary introduction to pre-big bang cosmologyRev. D 51 6744[52] Ellis G F R and Bruni S 1989 Phys. Rev. D 40 1804[53] Abbott R B, B<strong>ed</strong>narz B and Ellis S D 1986 Phys. Rev. D 33 2147[54] Gasperini M 1997 Phys. Rev. D 56 4815[55] Gasperini M, Giovannini M and Veneziano G 1993 Phys. Rev. D 48 R439[56] Grishchuk L P and Sidorov Y V 1989 Class. Quantum Grav. 6 L161Grishchuk L P and Sidorov Y V 1990 Phys. Rev. D 42 3413Grishchuk L P 1993 Class. Quantum Grav. 10 2449[57] Brandenberger R, Mukhanov V and Prokopec T 1992 Phys. Rev. Lett. 69 3606Gasperini M and Giovannini M 1993 Phys. Lett. B 301 334Gasperini M and Giovannini M 1993 Class. Quantum Grav. 10 L133[58] Grishchuk L P 1975 Sov. Phys.–JETP 40 409[59] Gasperini M 1998 String Theory in Curv<strong>ed</strong> Spacetimes <strong>ed</strong>NSánchez (Singapore:World Scientific) p 333[60] Abramowitz M and Stegun I A 1972 Handbook of Mathematical Functions (NewYork: Dover)[61] Mukhanov V F and Deruelle N 1995 Phys. Rev. D 52 5549[62] Allen B 1997 Proc. Les Houches School on ‘Astrophysical Sources of <strong>Gravitational</strong>Waves’ <strong>ed</strong> J A Mark and J P Lasota (Cambridge: Cambridge University Press)p 373Allen B and Romano J D 1999 Phys. Rev. D 59 10<strong>2001</strong>[63] Allen B, Flanagan E E and Papa M A 1999 Phys. Rev. D 61 024024[64] Grishchuk L P 1998 The detectability of relic (squeez<strong>ed</strong>) gravitational <strong>waves</strong> bylaser interferometers gr-qc/9810055[65] Gasperini M 1995 Proc. Second Paris Cosmology Colloquium (Paris, 1994) <strong>ed</strong>H J De Vega and N Sanchez (Singapore: World Scientific) p 429Gasperini M 1996 String Gravity and Physics at the Planck Energy Scale <strong>ed</strong>N Sanchez and A Zichichi (Dordrecht: Kluwer) p 305Gasperini M 1998 Second Amaldi Conf. on ‘<strong>Gravitational</strong> Waves’ (CERN, 1997) <strong>ed</strong>E Coccia et al (Singapore: World Scientific) p 62[66] Maggiore M 1998 High energy physics with gravitational wave experimentsPreprint gr-qc/9803028[67] Krauss L M and White M 1992 Phys. Rev. Lett. 69 869[68] Kaspi V, Taylor J and Ryba M 1994 Ap. J. 428 713[69] Schwarztmann V F 1969 JETP Lett. 9 184[70] Copeland E J, Liddle A R, Lidsey J E and Wands D 1998 Phys. Rev. D 58 63508Copeland E J, Liddle A R, Lidsey J E and Wands D 1998 Gen. Rel. Grav. 30 1711[71] Galluccio M, Litterio F and Occhionero F 1997 Phys. Rev. Lett. 79 970[72] Buonanno A, Maggiore M and Ungarelli C 1997 Phys. Rev. D 55 3330[73] Caldwell R R and Allen B 1992 Phys. Rev. D 45 3447Battye R A and Shellard E P S 1996 Phys. Rev. D 53 1811[74] Turner M S and Wilczek F 1990 Phys. Rev. Lett. 65 3080[75] Khlebnikov S Y and Tkachev I I 1997 Phys. Rev. D 56 653Bassett B 1997 Phys. Rev. D 56 3439[76] Peebles P J E and Vilenkin A 1999 Phys. Rev. D 59 063505Giovannini M 1999 Phys. Rev. D 60 123511[77] Astone P 1998 Second Amaldi Conf. on ‘<strong>Gravitational</strong> Waves’ (CERN, 1997) <strong>ed</strong>E Coccia et al (Singapore: World Scientific) p 192


References 337[78] Prodi G 1998 Talk Given at the Second Int. Conf. ‘Around VIRGO’ (Tirrenia, Pisa)See also Cerdonio M these proce<strong>ed</strong>ings[79] Pegoraro F, Picasso E and Radicati L 1978 J. Phys. A: Math. Gen. 11 1949[80] Picasso E Private communication[81] See, for instance, Lobo J A 1998 Second Amaldi Conf. on ‘<strong>Gravitational</strong> Waves’(CERN, 1997) <strong>ed</strong> E Coccia et al (Singapore: World Scientific) p 168[82] Ungarelli C and Vecchio A in preparation[83] Green M B, Schwartz J and Witten E 1987 Superstring Theory (Cambridge:Cambridge University Press)[84] Metsaev R R and Tseytlin A A 1987 Nucl. Phys. B 293 385[85] Taylor T R and Veneziano G 1988 Phys. Lett. B 213 459[86] See, for instance, Antoniadis I and Pioline B 1999 Large dimensions and stringphysics at a TeV Preprint hep-ph/9906480 (CPHT-PC720-0699, March 1999)[87] Buonanno A, Gasperini M, Maggiore M and Ungarelli C 1997 Class. QuantumGrav. 14 L97[88] Gasperini M and Giovannini M 1992 Phys. Lett. B 287 56[89] Gibbons G W and Hawking S 1977 Phys. Rev. D 15 2752


Chapter 17Post-Newtonian computation of binaryinspiral waveformsLuc BlanchetDépartement d’Astrophysique Relativiste et de Cosmologie,Centre National de la Recherche Scientifique (UMR 8629),Observatoire de Paris, 92195 Meudon C<strong>ed</strong>ex, France17.1 IntroductionAstrophysical systems known as inspiralling compact binaries are among themost interesting sources to hunt for gravitational radiation in the future networkof laser-interferometric detectors, compos<strong>ed</strong> of the large-scale interferometersVIRGO and LIGO, and the m<strong>ed</strong>ium-scale ones GEO and TAMA (see the books[1–3] for reviews, and the contribution of B Schutz in this volume). Thesesystems are compos<strong>ed</strong> of two compact objects, i.e. gravitationally-condens<strong>ed</strong>neutron stars or black holes, whose orbit follows an inward spiral, with decreasingorbital radius r and increasing orbital frequency ω. The inspiral is driven bythe loss of energy associat<strong>ed</strong> with the gravitational-wave emission. Because th<strong>ed</strong>ynamics of a binary is essentially aspherical, inspiralling compact binaries arestrong emitters of gravitational radiation. Tidal interactions between the compactobjects are expect<strong>ed</strong> to play a little role during most of the inspiral phase; themass transfer (in the case of neutron stars) does not occur until very late, nearthe final coalescence. Inspiralling compact binaries are very clean systems,essentially dominat<strong>ed</strong> by gravitational forces. Therefore, the relevant model fordescribing the inspiral phase consists of two point-masses moving under theirmutual gravitational attraction. As a simplification for the theoretical analysis,the orbit of inspiralling binaries can be consider<strong>ed</strong> to be circular, apart from thegradual inspiral, with a good approximation. At some point in the evolution, therewill be a transition from the adiabatic inspiral to the plunge of the two objectsfollow<strong>ed</strong> by the collision and final merger. Evidently the model of point-masses338


Introduction 339breaks down at this point, and is to be replac<strong>ed</strong> by a fully relativistic numericalcomputation of the plunge and merger (see the contribution of E Seidel in thisvolume).Currently the theoretical pr<strong>ed</strong>iction from general relativity for thegravitational <strong>waves</strong> emitt<strong>ed</strong> during the inspiral phase is determin<strong>ed</strong> using thepost-Newtonian approximation (see [4, 5] for reviews). This is possible becausethe dynamics of inspiralling compact binaries, though very relativistic, is notfully relativistic: the orbital velocity v is always less than one third of c (say).However, because 1/3 is far from negligible as compar<strong>ed</strong> to 1, the gravitationalradiationwaveform should be pr<strong>ed</strong>ict<strong>ed</strong> up to a high post-Newtonian order. Inparticular, the radiation reaction onto the orbit, which triggers the inspiral, is to b<strong>ed</strong>etermin<strong>ed</strong> with the maximal precision, corresponding to at least the second andmaybe the third post-Newtonian (3PN, or 1/c 6 ) order [6,7]. Notice that the zerothorder in this post-Newtonian counting corresponds to the dominant radiationreaction force (already of the order of 2.5PN relative to the Newtonian force),which is due to the change in the quadrupole moment of the source. Actually, themethod is not to compute directly the radiation reaction force but to determinethe inspiral rate from the energy balance equation relating the mechanical loss ofenergy in the binary’s centre of mass to the total emitt<strong>ed</strong> flux at infinity.The implement<strong>ed</strong> strategy is to develop a formalism for the emission andpropagation of gravitational <strong>waves</strong> from a general isolat<strong>ed</strong> system, and only then,once some general formulae valid to some prescrib<strong>ed</strong> post-Newtonian order arein our hands, to apply the formalism to compact binaries. Hence, we consider inthis paper a particular formalism applicable to a general description of matter,under the tenet of validity of the post-Newtonian expansion, namely that thematter should be slowly moving, weakly stress<strong>ed</strong> and self-gravitating. Within thisformalism we compute the retard<strong>ed</strong> far field of the source by means of a formalpost-Minkowskian expansion, valid in the exterior of the source, and parametriz<strong>ed</strong>by some appropriately defin<strong>ed</strong> multipole moments describing the source. Fromthe post-Minkowskian expansion we obtain a relation (correct up to the prescrib<strong>ed</strong>post-Newtonian order) between the radiative multipole moments parametrizingthe metric field at infinity, and the source multipole moments. On the other hand,the source multipole moments are obtain<strong>ed</strong> as some specific integrals extendingover the distribution of matter fields in the source and the contribution of thegravitational field itself. The source moments are comput<strong>ed</strong> separately up tothe same post-Newtonian order. The latter formalism has been develop<strong>ed</strong> byBlanchet, Damour and Iyer [8–14]. More recently, a different formalism hasbeen propos<strong>ed</strong> and implement<strong>ed</strong> by Will and Wiseman [15] (see also [16, 17]).The two formalisms are equivalent at the most general level, but the details of thecomputations are quite far apart. In the second stage, one applies the formalismto a system of point-particles (modelling compact objects) by substituting for thematter stress–energy tensor that expression, involving delta-functions, which isappropriate for point-particles. This entails some divergencies due to the infiniteself-field of point-particles. Our present method is to cure them systematically


340 Post-Newtonian computation of binary inspiral waveformsby means of a variant of the Hadamard regularization (bas<strong>ed</strong> on the concept of‘partie finie’) [18, 19].In this paper, we first analyse the binary inspiral gravitational waveform atthe simplest Newtonian approximation. Notably, we spend some time describingthe relative orientation of the binary with respect to the detector. Then wecompute, still at the ‘Newtonian order’ (corresponding, in fact, to the quadrupoleapproximation), the evolution in the course of time of the orbital phase of thebinary, which is a crucial quantity to pr<strong>ed</strong>ict. Next, we review the main steps ofour general wave-generation formalism, with emphasis on the definition of thevarious types of multipole moments which are involv<strong>ed</strong>. At last, we present theresult for the binary inspiral waveform whose current post-Newtonian precision is2PN in the wave amplitude and 2.5PN in the orbital phase (that is 1/c 5 beyond thequadrupole radiation reaction). However, since our ultimate aim is to constructaccurate templates to be us<strong>ed</strong> in the data analysis of detectors, it is appropriate towarm up with a short review of the optimal filtering technique which will be us<strong>ed</strong>for hunting the inspiral binary waveform (see [20] for an extend<strong>ed</strong> review).17.2 Summary of optimal signal filteringLet o(t) be the raw output of the detector, which is made of the superposition ofthe useful gravitational-wave signal h(t) and of noise n(t):o(t) = h(t) + n(t). (17.1)The noise is assum<strong>ed</strong> to be a stationary Gaussian random variable, with zeroexpectation value,n(t) = 0, (17.2)and with (suppos<strong>ed</strong>ly known) frequency-dependent power spectral density S n (ω)satisfyingñ(ω)ñ ∗ (ω ′ ) = 2πδ(ω − ω ′ )S n (ω), (17.3)where ñ(ω) is the Fourier transform of n(t). In (17.2) and (17.3), we denote byan upper bar the average over many realizations of noise in a large ensemble ofdetectors. From (17.3), we have S n (ω) = Sn ∗(ω) = S n(−ω) > 0.Looking for the signal h(t) in the output of the detector o(t), theexperimenters construct the correlation c(t) between o(t) and a filter q(t), i.e.c(t) =∫ +∞−∞dt ′ o(t ′ )q(t + t ′ ), (17.4)and divide c(t) by the square root of its variance (or correlation noise). Thus, theexperimenters consider the ratio∫ +∞c(t)σ [q](t) =(c 2 (t) − c(t) 2 = ( ∫) 1/2 +∞−∞ dω2π õ(ω) ˜q∗ (ω)e iωt−∞ dω2π S n(ω)|˜q(ω)| 2 ) 1/2, (17.5)


Summary of optimal signal filtering 341where õ(ω) and ˜q(ω) are the Fourier transforms of o(t) and q(t). The expectationvalue (or ensemble average) of this ratio defines the filter<strong>ed</strong> signal-to-noise ratioρ[q](t) = σ [q](t) =∫ +∞−∞ 2π dω ˜h(ω) ˜q ∗ (ω)e iωt( ∫ ) +∞1/2. (17.6)−∞ 2π dω S n(ω)|˜q(ω)| 2The optimal filter (or Wiener filter) which maximizes the signal-to-noise (17.6) ata particular instant t = 0 (say), is given by the match<strong>ed</strong> filtering theorem as˜q(ω) = γ ˜h(ω)S n (ω) , (17.7)where γ is an arbitrary real constant. The optimal filter (17.7) is match<strong>ed</strong> on theexpect<strong>ed</strong> signal ˜h(ω) itself, and weight<strong>ed</strong> by the inverse of the power spectraldensity of the noise. The maximum signal to noise, corresponding to the optimalfilter (17.7), is given by( ∫ +∞ρ =−∞dω2π| ˜h(ω)| 2 ) 1/2=〈h, h〉 1/2 . (17.8)S n (ω)This is the best achievable signal-to-noise ratio with a linear filter. In (17.8), wehave us<strong>ed</strong>, for any two real functions f (t) and g(t), the notation〈 f, g〉 =∫ +∞−∞dω2π˜ f (ω) ˜g ∗ (ω)S n (ω)(17.9)for an inner scalar product satisfying 〈 f, g〉 =〈f, g〉 ∗ =〈g, f 〉.In practice, the signal h(t) or ˜h(ω) is of known form (given, for instance,by (17.27)–(17.32) later) but depends on an unknown set of parameters whichdescribe the source of radiation, and are to be measur<strong>ed</strong>. The experimenters musttherefore use a whole family of filters analogous to (17.7) but in which the signalis parametriz<strong>ed</strong> by a whole family of ‘test’ parameters which are a priori differentfrom the actual source parameters. Thus, one will have to define and use a latticeof filters in the parameter space. The set of parameters maximizing the signalto noise (17.6) is equal, by the match<strong>ed</strong> filtering theorem, to the set of sourceparameters. However, in a single detector, the experimenters maximize the ratio(17.5) rather than the signal to noise (17.6), and therefore make errors on th<strong>ed</strong>etermination of the parameters, depending on a particular realization of noise inthe detector. If the signal-to-noise ratio is high enough, the measur<strong>ed</strong> values of theparameters are Gaussian distribut<strong>ed</strong> around the source parameters, with variancesand correlation coefficients given by the covariance matrix, the computation ofwhich we now recall. Since the optimal filter (17.7) is defin<strong>ed</strong> up to an arbitrarymultiplicative constant, it is convenient to treat separately a constant amplitudeparameter in front of the signal (involving, in general, the distance of the source).We shall thus write the signal in the form˜h(ω; A,λ a ) = A ˜k(ω; λ a ), (17.10)


342 Post-Newtonian computation of binary inspiral waveformswhere A denotes some amplitude parameter. The function ˜k depends only onthe other parameters, collectively denot<strong>ed</strong> by λ a where the label a ranges on thevalues 1,...,N. The family of match<strong>ed</strong> filters (or ‘templates’) we consider isdefin<strong>ed</strong> by˜q(ω; t λ a ) = γ ′ ˜k(ω; t λ a ), (17.11)S n (ω)where t λ a is a set of test parameters, assum<strong>ed</strong> to be all independent, and γ ′ isarbitrary. By substituting (17.11) into (17.5) and choosing t = 0, we get, with thenotation of (17.9),〈o, k( t λ)〉σ( t λ) =. (17.12)〈k( t λ), k( t λ)〉 1/2(Note that σ is in fact a function of both the parameters λ a and t λ a .) Nowthe experimenters choose as their best estimate of the source parametersλ a the measur<strong>ed</strong> parameters m λ a which among all the test parameters t λ a(independently) maximize (17.12), i.e. which satisfy∂σ( m λ) = 0, a = 1,...,N. (17.13)∂ t λ aAssuming that the signal to noise is high enough, we can work out (17.13) upto the first order in the difference between the actual source parameters and themeasur<strong>ed</strong> ones,δλ a = λ a − m λ a . (17.14)As a result, we obtain{δλ a = ab −〈n, ∂h}〈n, h〉 ∂h〉+ 〈h, 〉 , (17.15)∂λ b 〈h, h〉 ∂λ bwhere a summation is understood on the dummy label b, and where the matrix ab (with a, b = 1,...,N) is the inverse of the Fisher information matrix〈 ∂h ab = , ∂h 〉− 1 〈h, ∂h 〉〈h, ∂h 〉(17.16)∂λ a ∂λ b 〈h, h〉 ∂λ a ∂λ b(we have ab bc = δ ac ). On the right-hand sides of (17.15) and (17.16), thesignal is equally (with this approximation) parametriz<strong>ed</strong> by the measur<strong>ed</strong> or actualparameters. Since the noise is Gaussian, so are, by (17.15), the variables δλ a(inde<strong>ed</strong>, δλ a result from a linear operation on the noise variable). The expectationvalue and quadratic moments of the distribution of these variables are readilyobtain<strong>ed</strong> from the facts that 〈n, f 〉 = 0 and 〈n, f 〉〈n, g〉 = 〈f, g〉 for anydeterministic functions f and g (see (17.2) and (17.3)). We then obtainδλ a = 0,δλ a δλ b = ab . (17.17)


Newtonian binary polarization waveforms 343Thus, the matrix ab (the inverse of (17.16)) is the matrix of variances andcorrelation coefficients, or covariance matrix, of the variables δλ a . Theprobability distribution of δλ a reads asP(δλ a ) ={1√(2π) N+1 det exp − 1 }2 abδλ a δλ b , (17.18)where det is the determinant of ab . A similar analysis can be done for themeasurement of the amplitude parameter A of the signal.17.3 Newtonian binary polarization waveformsThe source of gravitational <strong>waves</strong> is a binary system made of two point-massesmoving on a circular orbit. We assume that the masses do not possess any intrinsicspins, so that the motion of the binary takes place in a plane. To simplify thepresentation we suppose that the centre of mass of the binary is at rest with respectto the detector. The detector is a large-scale laser-interferometric detector likeVIRGO or LIGO, with two perpendicular arms (with length 3 km in the caseof VIRGO). The two laser beams inside the arms are separat<strong>ed</strong> by the beamsplitterwhich defines the central point of the interferometer. We introduce anorthonormal right-hand<strong>ed</strong> triad ( X, Y , Z) link<strong>ed</strong> with the detector, with X andY pointing along the two arms of the interferometer, and Z pointing toward thezenithal direction. We denote by n the direction of the detector as seen from thesource, that is, −n is defin<strong>ed</strong> as the unit vector pointing from the centre of theinterferometer to the binary’s centre of mass. We introduce some spherical anglesα and β such that−n = X sin α cos β + Y sin α sin β + Z cos α. (17.19)Thus, the plane β = constant defines the plane which is vertical, as seen fromthe detector, and which contains the source. Next, we introduce an orthonormalright-hand<strong>ed</strong> triad ( x, y, z) which is link<strong>ed</strong> to the binary’s orbit, with x and ylocat<strong>ed</strong> in the orbital plane, and z along the normal to the orbital plane. Thevector x is chosen to be perpendicular to n; thus, n is within the plane form<strong>ed</strong>by y and z. The orientation of this triad is ‘right-hand’ with respect to the senseof motion. We denote by i the inclination angle, namely the angle between th<strong>ed</strong>irection of the source or line-of-sight n and the normal z to the orbital plane.Since z is right-hand<strong>ed</strong> with respect to the sense of motion we have 0 ≤ i ≤ π.Furthermore, we define two unit vectors p and q, call<strong>ed</strong> the polarization vectors,in the plane orthogonal to n (or plane of the sky). We choose p = x and define qin such a way that the triad (n, p, q) is right-hand<strong>ed</strong>; thusn = y sin i + z cos i, (17.20)p = x, (17.21)q = y cos i − z sin i. (17.22)


344 Post-Newtonian computation of binary inspiral waveformsNotice that the direction p ≡ x is one of the ‘ascending node’ N of the binary,namely the point at which the bodies cross the plane of the sky moving towardthe detector. Thus, the polarization vectors p and q lie, respectively, along themajor and minor axis of the projection onto the plane of the sky of the (circular)orbit, with p pointing toward N using the standard practice of celestial mechanics.Finally, let us denote by ξ the polarization angle between p and the vertical planeβ = constant; that is, ξ is the angle between the vertical and the direction of thenode N. Wehaven =−X sin α cos β − Y sin α sin β − Z cos α, (17.23)p = X(cos ξ cos α cos β + sin ξ sin β)+ Y (cos ξ cos α sin β − sin ξ cos β) − Z cos ξ sin α, (17.24)q = X(− sin ξ cos α cos β + cos ξ sin β)+ Y (− sin ξ cos α sin β − cos ξ cos β) + Z sin ξ sin α. (17.25)Defining all these angles, the relative orientation of the binary with respect to theinterferometric detector is entirely determin<strong>ed</strong>. Inde<strong>ed</strong> using (17.22) and (17.25)one relates the triad (x, y, z) associat<strong>ed</strong> with the source to the triad ( X, Y , Z)link<strong>ed</strong> with the detector.The gravitational wave as it propagates through the detector in the wavezone of the source is describ<strong>ed</strong> by the so-call<strong>ed</strong> transverse and traceless (TT)asymptotic waveform h TTij= (g ij −δ ij ) TT , where g ij denotes the spatial covariantmetric in a coordinate system adapt<strong>ed</strong> to the wave zone, and δ ij is the Kroneckermetric. Neglecting terms dying out like 1/R 2 in the distance to the source, thetwo polarization states of the wave, customarily denot<strong>ed</strong> h + and h × , are given byh + = 1 2 (p i p j − q i q j )h TTij , (17.26)h × = 1 2 (p iq j + p j q i )h TTij , (17.27)where p i and q i are the components of the polarization vectors. The detector isdirectly sensitive to a linear combination of the polarization waveforms h + andh × given byh(t) = F + h + (t) + F × h × (t), (17.28)where F + and F × are the so-call<strong>ed</strong> beam-pattern functions of the detector, whichare some given functions (for a given type of detector) of the direction of thesource α, β and of the polarization angle ξ. This h(t) is the gravitational-<strong>waves</strong>ignal look<strong>ed</strong> for in the data analysis of section 17.2, and us<strong>ed</strong> to construct theoptimal filter (17.10). In the case of the laser-interferometric detector we haveF + = 1 2 (1 + cos2 α) cos 2β cos 2ξ + cos α sin 2β sin 2ξ, (17.29)F × = − 1 2 (1 + cos2 α) cos 2β sin 2ξ + cos α sin 2β cos 2ξ. (17.30)The orbital plane and the direction of the node N are fix<strong>ed</strong> so the polarizationangle ξ is constant (in the case of spinning particles, the orbital plane precesses


Newtonian binary polarization waveforms 345around the direction of the total angular momentum, and angle ξ varies). Thus,the gravitational wave h(t) depends on time only through the two polarizationwaveforms h + (t) and h × (t). In turn, these waveforms depend on time throughthe binary’s orbital phase φ(t) and the orbital frequency ω(t) = dφ(t)/dt. Theorbital phase is defin<strong>ed</strong> as the angle, orient<strong>ed</strong> in the sense of motion, between theascending node N and the direction of one of the particles, conventionally particle1 (thus φ = 0 modulo 2π when the two particles lie along p, with particle 1 at theascending node). In the absence of any radiation reaction, the orbital frequencywould be constant, and so the phase would evolve linearly with time. Becauseof the radiation reaction forces, the actual variation of φ(t) is nonlinear, and theorbit spirals in and shrinks to zero-size to account, via the Kepler third law, forthe gravitational-radiation energy loss. The main problem of the constructionof accurate templates for the detection of inspiralling compact binaries is thepr<strong>ed</strong>iction of the time variation of the phase φ(t). Inde<strong>ed</strong>, because of theaccumulation of cycles, most of the accessible information allowing accuratemeasurements of the binary’s intrinsic parameters (such as the two masses) iscontain<strong>ed</strong> within the phase, and rather less accurate information is available in thewave amplitude itself. For instance, the relative precision in the determination ofthe distance R to the source, which affects the wave amplitude, is less than forthe masses, which strongly affect the phase evolution [6, 7]. Hence, we can oftenneglect the higher-order contributions to the amplitude, which means retainingonly the dominant harmonics in the waveform, which corresponds to a frequencyat twice the orbital frequency.Once the functions φ(t) and ω(t) are known they must be insert<strong>ed</strong> into thepolarization waveforms comput<strong>ed</strong> by means of some wave-generation formalism.For instance, using the quadrupole formalism, which neglects all the harmonicsbut the dominant one, we findh + =− 2Gµc 2 Rh × =− 2Gµc 2 R( ) Gmω 2/3c 3 (1 + cos 2 i) cos 2φ, (17.31)( ) Gmω 2/3c 3 (2 cos i) sin 2φ (17.32)where R denotes the absolute luminosity distance of the binary’s centre of mass;the mass parameters are given bym = m 1 + m 2 ; µ = m 1m 2m ; ν = µ m . (17.33)This last parameter ν, introduc<strong>ed</strong> for later convenience, is the ratio between ther<strong>ed</strong>uc<strong>ed</strong> mass and the total mass, and is such that 0


346 Post-Newtonian computation of binary inspiral waveforms17.4 Newtonian orbital phase evolutionLet y 1 (t) and y 2 (t) be the two trajectories of the masses m 1 and m 2 , andy = y 1 − y 2 be their relative position, and denote r =|y|. The velocities arev 1 (t) = dy 1 /dt, v 2 (t) = dy 2 /dt and v(t) = dy/dt. The Newtonian equations ofmotion read asdv 1dt=− Gm 2r 3 y;dv 2dt= Gm 1r 3 y. (17.34)The difference between these two equations yields the relative acceleration,dvdt=− Gm y. (17.35)r 3We place ourselves into the Newtonian centre-of-mass frame defin<strong>ed</strong> bym 1 y 1 + m 2 y 2 = 0, (17.36)in which frame the individual trajectories y 1 and y 2 are relat<strong>ed</strong> to the relative oney byy 1 = m 2m y; y 2 =− m 1y.m(17.37)The velocities are given similarly byv 1 = m 2m v; v 2 =− m 1v. (17.38)mIn principle, the binary’s phase evolution φ(t) should be determin<strong>ed</strong> froma knowl<strong>ed</strong>ge of the radiation reaction forces acting locally on the orbit. Atthe Newtonian order, this means considering the ‘Newtonian’ radiation reactionforce, which is known to contribute to the total acceleration only at the 2.5PNlevel, i.e. 1/c 5 smaller than the Newtonian acceleration (where 5 = 2s + 1, withs = 2 the helicity of the graviton). A simpler computation of the phase is tod<strong>ed</strong>uce it from the energy balance equation between the loss of centre-of-massenergy and the total flux emitt<strong>ed</strong> at infinity in the form of <strong>waves</strong>. In the case ofcircular orbits one ne<strong>ed</strong>s only to find the decrease of the orbital separation r andfor that purpose the balance of energy is sufficient. Relying on an energy balanceequation is the method we follow for computing the phase of inspiralling binariesin higher post-Newtonian approximations (see section 17.6). Thus, we writ<strong>ed</strong>Edt=−Ä, (17.39)where E is the centre-of-mass energy, given at the Newtonian order byE =− Gm 1m 2, (17.40)2r


Newtonian orbital phase evolution 347and where Ä denotes the total energy flux (or gravitational ‘luminosity’), d<strong>ed</strong>uc<strong>ed</strong>to the Newtonian order from the quadrupole formula of Einstein:Ä = G d 3 Q ij d 3 Q ij5c 5 dt 3 dt 3 . (17.41)The quadrupole moment is merely the Newtonian (trace-free) quadrupole of thesource, which reads in the case of the point-particle binary asIn the mass-centr<strong>ed</strong> frame (17.36) we getQ ij = m 1 (y1 i y j 1 − 1 3 δij y 2 1 ) + 1 ↔ 2. (17.42)Q ij = µ(y i y j − 1 3 δij r 2 ). (17.43)The third time derivative of Q ij ne<strong>ed</strong><strong>ed</strong> in the quadrupole formula (17.41) iseasily obtain<strong>ed</strong>. When an acceleration is generat<strong>ed</strong> we replace it by the Newtonianequation of motion (17.35). In the case of a circular orbit we getd 3 Q ijdt 3=−4 Gmµr 3 (y i v j + y j v i ) (17.44)(this is automatically trace-free because y · v = 0). Replacing (17.44) into (17.41)leads to the ‘Newtonian’fluxÄ = 32 G 3 m 2 µ 25 c 5 r 4 v 2 . (17.45)A better way to express the flux is in terms of some dimensionless quantities,namely the mass ratio ν given in (17.33), and a very convenient post-Newtonianparameter defin<strong>ed</strong> from the orbital frequency ω byx =( Gmωc 3 ) 2/3. (17.46)Notice that x is of formal order O(1/c 2 ) in the post-Newtonian expansion.Thanks to the Kepler law Gm = r 3 ω 2 we transform (17.45) and arrive atÄ = 32 c 55 G ν2 x 5 . (17.47)In this form the only factor having a dimension isc 5G ≈ 3.63 × 1052 W, (17.48)which is the Planck unit of a power, which turns out to be independent of thePlanck constant. (Notice that instead of c 5 /G the inverse ratio G/c 5 appears as


348 Post-Newtonian computation of binary inspiral waveformsa factor in the quadrupole formula (17.41).) On the other hand, we find that Ereads simplyE =− 1 2 µc2 x. (17.49)Next we replace (17.47) and (17.49) into the balance equation (17.39), and findin this way an ordinary differential equation which is easily integrat<strong>ed</strong> for theunknown x. We introduce for later convenience the dimensionless time variableτ = c3 ν5Gm (t c − t), (17.50)where t c is a constant of integration. Then the solution readsx(t) = 1 4 τ −1/4 . (17.51)It is clear that t c represents the instant of coalescence, at which (by definition)the orbital frequency diverges to infinity. Then a further integration yieldsφ = ∫ ω dt =−ν5 ∫ x 2/3 dτ, and we get the look<strong>ed</strong> for resultφ c − φ(t) = 1 ν τ 5/8 , (17.52)where φ c denotes the constant phase at the instant of coalescence. It is oftenuseful to consider the number Æ of gravitational-wave cycles which are left untilthe final coalescence starting from some frequency ω:Æ = φ c − φπ= 132πν x −5/2 . (17.53)As we see the post-Newtonian order of magnitude of Æ is c +5 , that is the inverseof the order c −5 of radiation reaction effects. As a matter of fact, Æ is a largenumber, approximately equal to 1.6×10 4 in the case of two neutron stars between10 and 1000 Hz (roughly the frequency bandwidth of the detector VIRGO). Dataanalysts of detectors have estimat<strong>ed</strong> that, in order not to suffer a too severer<strong>ed</strong>uction of signal to noise, one should monitor the phase evolution with anaccuracy comparable to one gravitational-wave cycle (i.e. δÆ ∼ 1) or better.Now it is clear, from a post-Newtonian point of view, that since the ‘Newtonian’number of cycles given by (17.53) is formally of order c +5 , any post-Newtoniancorrection therein which is larger than order c −5 is expect<strong>ed</strong> to contribute to thephase evolution more than that allow<strong>ed</strong> by the previous estimate. Therefore, oneexpects that in order to construct accurate templates it will be necessary to includeinto the phase the post-Newtonian corrections up to at least the 2.5PN or 1/c 5order. This expectation has been confirm<strong>ed</strong> by various studies [21–24] whichshow<strong>ed</strong> that in advanc<strong>ed</strong> detectors the 2.5PN or, better, the 3PN approximation isrequir<strong>ed</strong> in the case of inspiralling neutron star binaries. Notice that 3PN heremeans 3PN in the centre-of-mass energy E, which is d<strong>ed</strong>uc<strong>ed</strong> from the 3PNequations of motion, as well as in the total flux Ä, which is comput<strong>ed</strong> from a 3PNwave-generation formalism. For the moment the phase has been complet<strong>ed</strong> to the2.5PN order [15, 25–27]; the 3PN order is still incomplete (but, see [13, 28, 29]).


17.5 Post-Newtonian wave generation17.5.1 Field equationsPost-Newtonian wave generation 349We consider a general compact-support stress–energy tensor T µν describing theisolat<strong>ed</strong> source, and we look for the solutions, in the form of a (formal) post-Newtonian expansion, of the Einstein field equations,R µν − 1 2 gµν R = 8πGc 4 T µν , (17.54)and thus also of their consequence, the equations of motion ∇ ν T µν = 0 of thesource. We impose the condition of harmonic coordinates, i.e. the gauge condition∂ ν h µν = 0; h µν = √ −gg µν − η µν , (17.55)where g and g µν denote the determinant and inverse of the covariant metric g µν ,and where η µν is a Minkowski metric: η µν = diag(−1, 1, 1, 1). Then the Einsteinfield equations (17.54) can be replac<strong>ed</strong> by the so-call<strong>ed</strong> relax<strong>ed</strong> equations, whichtake the form of simple wave equations,£h µν = 16πGc 4 τ µν , (17.56)where the box operator is the flat d’Alembertian £ = η µν ∂ µ ∂ ν , and where thesource term τ µν can be view<strong>ed</strong> as the stress–energy pseudotensor of the matterand gravitational fields in harmonic coordinates. It is given byτ µν =|g|T µν +c416πG µν . (17.57)τ µν is not a generally-covariant tensor, but only a Lorentz tensor relative to theMinkowski metric η µν . As a consequence of the gauge condition (17.55), τ µν isconserv<strong>ed</strong> in the usual sense,∂ ν τ µν = 0 (17.58)(this is equivalent to ∇ ν T µν = 0). The gravitational source term µν is a quitecomplicat<strong>ed</strong>, highly nonlinear (quadratic at least) functional of h µν and its firstandsecond-spacetime derivatives.We supplement the resolution of the field equations (17.55) and (17.56) bythe requirement that the source does not receive any radiation from other sourceslocat<strong>ed</strong> very far away. Such a requirement of ‘no-incoming radiation’ is to beimpos<strong>ed</strong> at Minkowskian past null infinity (taking advantage of the presence ofthe Minkowski metric η µν ); this corresponds to the limit r =|x| →+∞witht + r/c = constant. (Please do not confuse this r with the same r denoting theseparation between the two bodies in section 17.4.) The precise formulation ofthe no-incoming radiation condition is[ ∂limr→+∞ ∂r (rhµν ) + ∂ ]c∂t (rhµν ) (x, t) = 0. (17.59)t+ r c =constant


350 Post-Newtonian computation of binary inspiral waveformsIn addition, r∂ λ h µν should be bound<strong>ed</strong> in the same limit. Actually we often adopt,for technical reasons, the more restrictive condition that the field is stationarybefore some finite instant −Ì in the past (refer to [8] for details). With theno-incoming radiation condition (17.58) or (17.59) we transform the differentialEinstein equation (17.56) into the equivalent integro-differential systemh µν = 16πG £ −1c 4 R τ µν , (17.60)where £ −1Rdenotes the standard retard<strong>ed</strong> inverse d’Alembertian given by(£ −11Rτ)(x, t) =−4π17.5.2 Source moments∫d 3 x ′|x − x ′ | τ(x′ , t −|x − x ′ |/c). (17.61)In this section we shall solve the field equations (17.55) and (17.56) in the exteriorof the isolat<strong>ed</strong> source by means of a multipole expansion, parametriz<strong>ed</strong> by someappropriate source multipole moments. The particularity of the moments we shallobtain, is that they are defin<strong>ed</strong> from the formal post-Newtonian expansion of thepseudotensor τ µν , supposing that the latter expansion can be iterat<strong>ed</strong> to any order.Therefore, these source multipole moments are physically valid only in the caseof a slowly-moving source (slow internal velocities; weak stresses). The generalstructure of the post-Newtonian expansion involves besides the usual powers of1/c some arbitrary powers of the logarithm of c, sayτ µν (t, x, c) = ∑ p,q(ln c) qc p τ pq µν (t, x), (17.62)where the overbar denotes the formal post-Newtonian expansion, and whereτ pq µν are the functional coefficients of the expansion (p, q are integers, includingzero). Now, the general multipole expansion of the metric field h µν , denot<strong>ed</strong>by Å(h µν ), is found by requiring that when re-develop<strong>ed</strong> into the near-zone,i.e. in the limit where r/c → 0 (this is equivalent with the formal re-expansionwhen c →∞), it matches with the multipole expansion of the post-Newtonianexpansion h µν (whose structure is similar to (17.62)) in the sense of themathematical technics of match<strong>ed</strong> asymptotic expansions. We find [11, 14] thatthe multipole expansion Å(h µν ) satisfying the matching is uniquely determin<strong>ed</strong>,and is compos<strong>ed</strong> of the sum of two terms,(−) ll!l=0(17.63)The first term, in which £ −1Ris the flat retard<strong>ed</strong> operator (17.61), is a particularsolution of the Einstein field equations in vacuum (outside the source), i.e. itÅ(h µν ) = finite part£ −1R [Å(µν )] − 4Gc 4 +∞ ∑∂ L{ 1r µνL (t − r/c) }.


Post-Newtonian wave generation 351satisfies £h µνpart = Å( µν ). The second term is a retard<strong>ed</strong> solution of thesource-free homogeneous wave equation, i.e. £h µνhom = 0. We denote ∂ L =∂ i1 ...∂ il where L = i 1 ...i l is a multi-index compos<strong>ed</strong> of l indices; the lsummations over the indices i 1 ...i l are not indicat<strong>ed</strong> in (17.63). The ‘multipolemoments’parametrizing this homogeneous solution are given explicitly by (withu = t − r/c)∫ µνL(u) = finite part∫ 1d 3 x ˆx L dz δ l (z)τ µν (x, u + z|x|/c), (17.64)−1where the integrand contains the post-Newtonian expansion of the pseudostress–energy tensor τ µν , whose structure reads like (17.62). In (17.64), we denote thesymmetric-trace-free (STF) projection of the product of l vectors x i with a hat,so that ˆx L = STF(x L ), with x L = x i 1...x i land L = i 1 ...i l ; for instance,ˆx ij = x i x j − 1 3 δ ij x 2 . The function δ l (z) is given byand satisfies the propertiesδ l (z) =(2l + 1)!!2 l+1 (1 − z 2 ) l , (17.65)l!∫ 1−1dzδ l (z) = 1;lim δ l(z) = δ(z) (17.66)l→+∞(where δ(z) is the Dirac measure). Both terms in (17.63) involve an operationof taking a finite part. This finite part can be defin<strong>ed</strong> precisely by means of ananalytic continuation (see [14] for details), but it is in fact basically equivalent totaking the finite part of a divergent integral in the sense of Hadamard [18]. Notice,in particular, that the finite part in the expression of the multipole moments (17.64)deals with the behaviour of the integral at infinity: r →∞(without the finite partthe integral would be divergent because of the factor x L = r l n L in the integrandand the fact that the pseudotensor τ µν is not of compact support).The result (17.63)–(17.64) permits us to define a very convenient notion ofthe source multipole moments (by opposition to the radiative moments defin<strong>ed</strong>below). Quite naturally, the source moments are construct<strong>ed</strong> from the tencomponents of the tensorial function µνL(u). Among these components fourcan be eliminat<strong>ed</strong> using the harmonic gauge condition (17.55), so in the endwe find only six independent source multipole moments. Furthermore, it can beshown that by changing the harmonic gauge in the exterior zone one can furtherr<strong>ed</strong>uce the number of independent moments to only two. Here we shall report theresult for the ‘main’ multipole moments of the source, which are the mass-typemoment I L and current-type J L (the other moments play a small role starting onlyat highorder in the post-Newtonian expansion). We have [14]


352 Post-Newtonian computation of binary inspiral waveforms∫ ∫ 1{I L (u) = finite part d 3 4(2l + 1)x dz δ l ˆx L −−1c 2 (l + 1)(2l + 3) δ l+1 ˆx iL ∂ t i}2(2l + 1)+c 4 (l + 1)(l + 2)(2l + 5) δ l+2 ˆx ijL ∂t 2 ij , (17.67)∫ ∫ 1J L (u) = finite part d 3 x dz ε aba b−1}2l + 1−c 2 (l + 2)(2l + 3) δ l+1 ˆx L−1>ac ∂ t bc . (17.68)Here the integrand is evaluat<strong>ed</strong> at the instant u + z| x|/c, ε abc is the Levi-Civitasymbol, 〈L〉 is the STF projection, and we employ the notation = τ 00 + τ iic 2 ; i = τ 0ic ; ij = τ ij (17.69)(with τ ii = δ ij τ ij ). The multipole moments I L , J L are valid formally up to anypost-Newtonian order, and constitute a generalization in the nonlinear theory ofthe usual mass and current Newtonian moments (see, [14] for details). It can becheck<strong>ed</strong> that, when consider<strong>ed</strong> at the 1PN order, these moments agree with th<strong>ed</strong>ifferent expressions obtain<strong>ed</strong> in [9] (case of mass moments) and in [10] (currentmoments).17.5.3 Radiative momentsIn lineariz<strong>ed</strong> theory, where we can neglect the gravitational source term µν in(17.57), as well as the first term in (17.63), the source multipole moments coincidewith the so-call<strong>ed</strong> radiative multipole moments, defin<strong>ed</strong> as the coefficients of themultipole expansion of the 1/r term in the distance to the source at retard<strong>ed</strong> timest − r/c = constant. However, in full nonlinear theory, the first term in (17.63)will bring another contribution to the 1/r term at future null infinity. Therefore,the source multipole moments are not the ‘measur<strong>ed</strong>’ ones at infinity, and sothey must be relat<strong>ed</strong> to the real observables of the field at infinity which areconstitut<strong>ed</strong> by the radiative moments. It has been known for a long time thatthe harmonic coordinates do not belong to the class of Bondi coordinate systemsat infinity, because the expansion of the harmonic metric when r → ∞ witht − r/c = constant involves, in addition to the normal powers of 1/r, somepowers of the logarithm of r. Let us change the coordinates from harmonic tosome Bondi-type or ‘radiative’ coordinates ( X, T ) such that the metric admits apower-like expansion without logarithms when R →∞with T − R/c = constantand R =|X| (it can be shown that the condition to be satisfi<strong>ed</strong> by the radiativecoordinate system is that the retard<strong>ed</strong> time T − R/c becomes asymptotically nullat infinity). For the purpose of deriving the formula (17.73) below it is sufficient


to transform the coordinates according toT − R c = t − r c − 2GMc 3Post-Newtonian wave generation 353( ) rln , (17.70)r 0where M denotes the ADM mass of the source and r 0 is a gauge constant. Inradiative coordinates it is easy to decompose the 1/R term of the metric intomultipoles and to define in that way the radiative multipole moments U L (masstype;where L = i 1 ...i l with l ≥ 2) and V L (current-type; with l ≥ 2).(Actually, it is often simpler to bypass the ne<strong>ed</strong> for transforming the coordinatesfrom harmonic to radiative by considering directly the TT projection of the spatialcomponents of the harmonic metric at infinity.) The formula for the definition ofthe radiative moments ish TTij= − 4G+∞∑c 2 R È ijab(N)l=2−{1c l N L−2 U abL−2 (T − R/c)l!}2lc(l + 1) N cL−2ε cd(a V b)dL−2 (T − R/c)( ) 1+ OR 2(17.71)where N is the vector N i = N i = X i /R (for instance N L−2 = N i1 ...N il−2 ), andÈ ijab denotes the TT projectorÈ ijab = (δ ia − N i N a )(δ jb − N j N b ) − 1 2 (δ ij − N i N j )(δ ab − N a N b ). (17.72)In the limit of lineariz<strong>ed</strong> gravity the radiative multipole moments U L , V L agreewith the lth time derivatives of the source moments I L , J L . Let us give, withoutproof, the result for the expression of the radiative mass-quadrupole moment U ijincluding relativistic corrections up to the 3PN or 1/c 6 order inclusively [12, 13].The calculation involves implementing explicitly a post-Minkowskian algorithmdefin<strong>ed</strong> in [8] for the computation of the nonlinearities due to the first term of(17.63). We find (U ≡ T − R/c)U ij (U) = M (2)ij(U) + 2 GM+ G c 5 {− 2 7∫ +∞0c 3 ∫ +∞0[ ( )dv M (4)cvij(U − v) ln2r 0+ 11 ]12dv [M (3)aa ](U − v) − 2 7 M(3) aa (U)− 5 7 M(4) aa (U) + 1 7 M(5) aa(U) + 1 3 ε aba J b(U)( GM+ 2[ ( cv× ln 2+ Oc 3 ) 2 ∫ +∞0dv M (5)ij(U − v))+ 57 ( ) ]cv2r 0 70 ln 124 627+2r 0 44 100( ) 1c 7 . (17.73)}


354 Post-Newtonian computation of binary inspiral waveformsThe superscript (n) denotes n time derivations. The quadrupole moment M ijentering this formula is closely relat<strong>ed</strong> to the source quadrupole I ij ,M ij = I ij + 2G( )3c 5 {K (3) I ij − K (2) I (1) 1ij}+Oc 7 , (17.74)where K is the Newtonian moment of inertia (see equation (4.24) in [27]; weare using here a mass-centr<strong>ed</strong> frame so that the mass-dipole moment I i is zero).The Newtonian term in (17.73) corresponds to the quadrupole formalism. Next,there is a quadratic nonlinear correction term with multipole interaction M × M ijwhich represents the effect of tails of gravitational <strong>waves</strong> (scattering of linear<strong>waves</strong> off the spacetime curvature generat<strong>ed</strong> by the mass M). This correction isof order 1/c 3 or 1.5PN and takes the form of a non-local integral with logarithmickernel [30]. It is responsible notably for the term proportional to πτ 1/4 in theformula for the phase (17.87) below. The next correction, of order 1/c 5 or2.5PN, is constitut<strong>ed</strong> by quadratic interactions between two mass-quadrupoles,and between a mass-quadrupole and the constant current dipole [12]. This termcontains also a non-local integral, which is due to the radiation of gravitational<strong>waves</strong> by the distribution of the stress–energy of linear <strong>waves</strong> [12,30–32]. Finally,at the 3PN order in (17.73) the first cubic nonlinear interaction appears, which isof the type (M × M × M ij ) and corresponds to the tails generat<strong>ed</strong> by the tailsthemselves [13].17.6 Inspiral binary waveformTo conclude, let us give (without proof) the result for the two polarizationwaveforms h + (t) and h × (t) of the inspiralling compact binary develop<strong>ed</strong> to 2PNorder in the amplitude and to 2.5PN order in the phase. The calculation wasdone by Blanchet, Damour, Iyer, Will and Wiseman [15, 25–27, 33], bas<strong>ed</strong> on theformalism review<strong>ed</strong> in section 17.5 and, independently, on that defin<strong>ed</strong> in [15].Following [33] we present the polarization waveforms in a form which is ready foruse in the data analysis of binary inspirals in the detectors VIRGO and LIGO (theanalysis will be bas<strong>ed</strong> on the optimal filtering technique review<strong>ed</strong> in section 17.2).We find, extending the Newtonian formulae in section 17.3,h +,× = 2Gµc 2 R( Gmωc 3 ) 2/3×{H +,× (0) + x 1/2 H (1/2)+,× + xH(1) +,× + x 3/2 H (3/2)+,× + x 2 H +,× (2) }, (17.75)where the various post-Newtonian terms, order<strong>ed</strong> by x, are given for the pluspolarization byH (0)+ =−(1 + c2 iH (1/2)+ =− s i8) cos 2ψ, (17.76)δmm [(5 + c2 i ) cos ψ − 9(1 + c2 i ) cos 3ψ], (17.77)


Inspiral binary waveform 355H + (1) = 1 6 [(19 + 9c2 i − 2ci 4 ) − ν(19 − 11c2 i − 6ci 4 )] cos 2ψ− 4 3 s2 i (1 + c2 i )(1 − 3ν)cos 4ψ, (17.78)H (3/2)+ = s i δm192 m {[(57 + 60c2 i − ci 4 ) − 2ν(49 − 12c2 i − ci 4 )] cos ψ− 27 2 [(73 + 40c2 i − 9ci 4 ) − 2ν(25 − 8c2 i − 9ci 4 )] cos 3ψ+ 6252 (1 − 2ν)s2 i (1 + c2 i ) cos 5ψ}−2π(1 + c2 i ) cos 2ψ, (17.79)H + (2) = 120 1 [(22 + 396c2 i + 145ci 4 − 5ci 6 ) + 5 3 ν(706 − 216c2 i − 251ci 4 + 15ci 6 )− 5ν 2 (98 − 108ci 2 + 7ci 4 + 5ci 6 )] cos 2ψ+15 2 s2 i [(59 + 35c2 i − 8ci 4 ) − 5 3 ν(131 + 59c2 i − 24ci 4 )+ 5ν 2 (21 − 3ci 2 − 8ci 4 )] cos 4ψ− 8140 (1 − 5ν + 5ν2 )si 4 (1 + c2 i ) cos 6ψ+ s i δm40 m {[11 + 7c2 i + 10(5 + ci 2 ) ln 2] sin ψ − 5π(5 + c2 i ) cos ψ− 27[7 − 10 ln(3/2)](1 + ci 2 ) sin 3ψ + 135π(1 + c2 i ) cos 3ψ}, (17.80)and for the cross-polarization byH × (0) =−2c i sin 2ψ, (17.81)H (1/2)× =− 3 4 s δmi c i [sin ψ − 3 sin 3ψ], (17.82)mH × (1) = c i3 [(17 − 4c2 i ) − ν(13 − 12c2 i )] sin 2ψ − 8 3 (1 − 3ν)c isi 2 sin 4ψ,(17.83)H (3/2)× = s ic i δm96 m {[(63 − 5c2 i ) − 2ν(23 − 5c2 i )] sin ψ− 27 2 [(67 − 15c2 i ) − 2ν(19 − 15c2 i )] sin 3ψ+ 6252 (1 − 2ν)s2 i sin 5ψ}−4πc i sin 2ψ, (17.84)H × (2) = c i60 [(68 + 226c2 i − 15ci 4 ) + 5 3 ν(572 − 490c2 i + 45ci 4 )− 5ν 2 (56 − 70ci 2 + 15ci 4 )] sin 2ψ+15 4 c i si 2 [(55 − 12c2 i ) − 5 3 ν(119 − 36c2 i ) + 5ν2 (17 − 12ci 2 )] sin 4ψ− 8120 (1 − 5ν + 5ν2 )c i si 4 sin 6ψ− 320 s δmi c i {[3 + 10 ln 2] cos ψ + 5π sin ψm− 9[7 − 10 ln(3/2)] cos 3ψ − 45π sin 3ψ}. (17.85)The notation is consistent with sections 17.3 and 17.4. In particular, the post-Newtonian parameter x is defin<strong>ed</strong> by (17.46). We use the shorthands c i = cos i


356 Post-Newtonian computation of binary inspiral waveformsand s i = sin i where i is the inclination angle. The basic phase variable ψ enteringthe waveforms is defin<strong>ed</strong> byψ = φ − 2Gmω ( ) ωc 3 ln , (17.86)ω 0where φ is the actual orbital phase of the binary, and where ω 0 can be chosen asthe seismic cut-off of the detector (see [33] for details). As for the phase evolutionφ(t), it is given up to 2.5PN order, generalizing the Newtonian formula (17.52),byφ(t) = φ 0 − 1 { ( 3715τ 5/8 +ν 8064 + 55 )96 ν τ 3/8 − 3 4 πτ1/4( 9 275 495 284 875++14 450 688 258 048 ν + 1855 )2048 ν2 τ 1/8(38 645+ −172 032 − 15 ) }2048 ν π ln τ , (17.87)where φ 0 is a constant and where we recall that the dimensionless time variableτ was given by (17.50). The frequency is equal to the time derivative of (17.87),hencec3{ ( 743τ −3/8 +ω(t) =8Gm( 1 855 099+14 450 688+2688 + 1132 ν )τ −5/8 − 3 10 πτ−3/4+56 975258 048 ν + 3712048 ν2 )τ −7/8(− 772921 504 − 3256 ν )πτ −1 }. (17.88)We have check<strong>ed</strong> that both waveforms (17.76)–(17.81) and phase/frequency(17.87)–(17.88) agree in the test mass limit ν → 0 with the results of linearblack hole perturbations as given by Tagoshi and Sasaki [34].References[1] Ciufolini I and Fidecaro F (<strong>ed</strong>) 1996 <strong>Gravitational</strong> Waves, Sources and Detectors(Singapore: World Scientific)[2] Marck J-A and Lasota J-P (<strong>ed</strong>) 1997 Relativistic Gravitation and <strong>Gravitational</strong>Radiation (Cambridge: Cambridge University Press)[3] Davier M and Hello P (<strong>ed</strong>) 1997 <strong>Gravitational</strong> Wave Data Analysis (Gif-sur-Yvette:Edition Frontières)[4] Will C M 1994 Proc. 8th Nishinomiya–Yukawa Symposium on RelativisticCosmology <strong>ed</strong> M Sasaki (Universal Academic)[5] Blanchet L 1997 Relativistic Gravitation and <strong>Gravitational</strong> Radiation <strong>ed</strong> J-A Marckand J-P Lasota (Cambridge: Cambridge Univerity Press)


References 357[6] Cutler C et al 1993 Phys. Rev. Lett. 70 2984[7] Cutler C and Flanagan E 1994 Phys. Rev. D 49 2658[8] Blanchet L and Damour T 1986 Phil. Trans. R. Soc. A 320 379[9] Blanchet L and Damour T 1989 Ann. Inst. H. Poincaré (Phys. Théorique) 50 377[10] Damour T and Iyer B R 1991 Ann. Inst. H. Poincaré (Phys. Théorique) 54 115[11] Blanchet L 1995 Phys. Rev. D 51 2559[12] Blanchet L 1998 Class. Quantum Grav. 15 89[13] Blanchet L 1998 Class. Quantum Grav. 15 113[14] Blanchet L 1998 Class. Quantum Grav. 15 1971[15] Will C M and Wiseman A G 1996 Phys. Rev. D 54 4813[16] Epstein R and Wagoner R V 1975 Astrophys. J. 197 717[17] Thorne K S 1980 Rev. Mod. Phys. 52 299[18] Hadamard J 1932 Le problème de Cauchy et les Équations aux Dérivées PartiellesLinéaires Hyperboliques (Paris: Hermann)[19] Blanchet L and Faye G 2000 J. Math. Phys. 41 7675Blanchet L and Faye G Preprint gr-qc 0004008[20] Wainstein L A and Zubakov L D 1962 Extraction of Signals from Noise (EnglewoodCliffs, NJ: Prentice-Hall)[21] Cutler C, Finn L S, Poisson E and Sussman G J 1993 Phys. Rev. D 47 1511[22] Tagoshi H and Nakamura T 1994 Phys. Rev. D 49 4016[23] Poisson E 1995 Phys. Rev. D 52 5719Poisson E 1997 Phys. Rev. D 55 7980[24] Damour T, Iyer B R and Sathyaprakash B S 1998 Phys. Rev. D 57 885[25] Blanchet L, Damour T, Iyer B R, Will C M and Wiseman A G 1995 Phys. Rev. Lett.74 3515[26] Blanchet L, Damour T and Iyer B R 1995 Phys. Rev. D 51 5360[27] Blanchet L 1996 Phys. Rev. D 54 1417[28] Jaranowski P and Schäfer G 1998 Phys. Rev. D 57 7274[29] Blanchet L and Faye G 2000 Phys. Lett. A to appearBlanchet L and Faye G Preprint gr-qc 0004009[30] Blanchet L and Damour T 1992 Phys. Rev. D 46 4304[31] Wiseman A G and Will C M 1991 Phys. Rev. D 44 R2945[32] Thorne K S 1992 Phys. Rev. D 45 520[33] Blanchet L, Iyer B R, Will C M and Wiseman A G 1996 Class. Quantum Grav. 13575[34] Tagoshi H and Sasaki M 1994 Prog. Theor. Phys. 92 745


PART 5NUMERICAL RELATIVITYEdward SeidelMax-Planck-Institut für Gravitationsphysik,Albert-Einstein-Institut, Am Mühlenberg 5, D-14473, Potsdam,Germany and University of Illinois, NCSA and Departments ofPhysics and Astronomy, Champaign, IL 61820, USA


Chapter 18Numerical relativityGeneral relativity is the fundamental theory of gravity, which is govern<strong>ed</strong>by an extremely complex set of coupl<strong>ed</strong>, nonlinear, hyperbolic-elliptic partialdifferential equations. General solutions to these equations, ne<strong>ed</strong><strong>ed</strong> to fullyunderstand their implications as a fundamental theory of physics, are elusive.Additionally, the astrophysics of compact objects, which requires Einstein’stheory of general relativity for understanding phenomena such as black holesand neutron stars, is attracting increasing attention. The largest parallelsupercomputers are finally approaching the spe<strong>ed</strong> and memory requir<strong>ed</strong> to solvethe complete set of Einstein’s equations for the first time since they were writtenover 80 years ago, allowing one to attempt full 3D simulations of such excitingevents as colliding black holes and neutron stars. In this paper we review thecomputational effort in this direction, and discuss a new 3D multipurpose parallelcode call<strong>ed</strong> ‘Cactus’ for general relativistic astrophysics. Directions for furtherwork are indicat<strong>ed</strong> where appropriate.18.1 OverviewThis article is intend<strong>ed</strong> to provide an introduction and overview to numericalrelativity, following a set of lectures given at the Como SIGRAV School ongravitational <strong>waves</strong> in Spring, 1999. We have bas<strong>ed</strong> them heavily on severalprevious articles, especially [1, 2], but we have tri<strong>ed</strong> to extend and update themwhere significant new work has been done.The Einstein equations for the structure of spacetime have remain<strong>ed</strong>essentially unchang<strong>ed</strong> since their discovery nearly a century ago, providing theunderpinnings of modern theories of gravity, astrophysics and cosmology. Thetheory is essential in describing phenomena such as black holes, compact objects,supernovae, and the formation of structure in the universe. Unfortunately, theequations are a set of highly complex, coupl<strong>ed</strong>, nonlinear partial differentialequations involving ten functions of four independent variables. They are amongthe most complicat<strong>ed</strong> equations in mathematical physics. For this reason, in361


362 Numerical relativityspite of more than 80 years of intense study, the solution space to the full set ofequations is essentially unknown. Most of what we know about this fundamentaltheory of physics has been glean<strong>ed</strong> from lineariz<strong>ed</strong> solutions, highly idealiz<strong>ed</strong>solutions possessing a high degree of symmetry (e.g., static, or spherically oraxially symmetric), or from perturbations of these solutions.Over the last several decades a growing research area, call<strong>ed</strong> numericalrelativity, has develop<strong>ed</strong>, where computers are employ<strong>ed</strong> to construct numericalsolutions to these equations. Although much has been learn<strong>ed</strong> through thisapproach, progress has been slow due to the complexity of the equations andinadequate computer power. For example, an important astrophysical applicationis the 3D spiralling coalescence of two black holes (BH) or neutron stars(NS), which will generate strong sources of gravitational <strong>waves</strong>. As has beenemphasiz<strong>ed</strong> by Flanagan and Hughes, one of the best candidates for earlydetection by the laser interferometer network is increasingly consider<strong>ed</strong> to be BHmergers [3, 4]. The imminent arrival of data from the long await<strong>ed</strong> gravitationalwaveinterferometers (see, e.g., [3] and references therein) has provid<strong>ed</strong> a senseof urgency in understanding these strong sources of gravitational <strong>waves</strong>. Suchunderstanding can be obtain<strong>ed</strong> only through large-scale computer simulationsusing the full machinery of numerical relativity.Furthermore, the gravitational-wave signals are likely to be so weak by thetime they reach the detectors that reliable detection may be difficult without priorknowl<strong>ed</strong>ge of the merger waveform. These signals can be properly interpret<strong>ed</strong>, orperhaps even detect<strong>ed</strong>, only with a detail<strong>ed</strong> comparison between the observationaldata and a set of theoretically determin<strong>ed</strong> ‘waveform templates’. In most cases,these waveform templates ne<strong>ed</strong><strong>ed</strong> for gravitational-wave data analysis have tobe generat<strong>ed</strong> by large-scale computer simulations, adding to the urgency ofdeveloping numerical relativity. However, a realistic 3D simulation bas<strong>ed</strong> on thefull Einstein equations is a highly non-trivial task—one can estimate the timerequir<strong>ed</strong> for a reasonably accurate 3D simulation of, say, the coalescence of acompact object binary, to be at least 100 000 Cray Y-MP hours!However, there is good reason for optimism that such problems can be solv<strong>ed</strong>within the next decade. Scalable parallel computers, and efficient algorithms thatexploit them, are quickly revolutionizing computational science, and numericalrelativity is a great beneficiary of these developments. Over the last years thecommunity has develop<strong>ed</strong> 3D codes design<strong>ed</strong> to solve the complete set of Einsteinequations that run very efficiently on large-scale parallel computers. We willdescribe below one such code, call<strong>ed</strong> ‘Cactus’, that has achiev<strong>ed</strong> 142 GFlops ona 1024 node Cray T3E-1200, which is more than 2000 times faster than 2D codesof a few years ago running on a Cray Y-MP (which also had only about 0.5%the memory capacity of the large T3E). Such machines are expect<strong>ed</strong> to scaleup rapidly as faster processors are connect<strong>ed</strong> together in even higher numbers,achieving Teraflop performance on real applications in a few years.Numerical relativity requires not only large computers and efficient codes,but also a wide variety of numerical algorithms for evolving and analysing


Einstein equations for relativity 363the solution. Because of this richness and complexity of the equations,and the interesting applications to problems such as black holes and neutronstars, natural collaborations have develop<strong>ed</strong> between appli<strong>ed</strong> mathematicians,physicists, astrophysicists and computational scientists in the development of asingle code to attack these problems. There are various large-scale collaborativeefforts in recent years in this direction, including the NSF Black Hole GrandChallenge Project (recently conclud<strong>ed</strong>), the NASA Neutron Star Grand ChallengeProject, the NCSA/Potsdam/Wash U numerical relativity collaboration, and mostrecently a large European collaboration of ten institutions fund<strong>ed</strong> by the EU [5].We will describe the Cactus Computational Toolkit, along with some ofits algorithms and capabilities of this code, and a number of its applications toproblems of black holes, gravitational <strong>waves</strong>, and neutron stars. In the nextsections we will first give a brief description of the numerical formulation ofthe theory of general relativity, and discuss particular difficulties associat<strong>ed</strong> withnumerical relativity. The discussion will necessarily be brief. Examples aremostly drawn from work carri<strong>ed</strong> out by the NCSA/Potsdam/Wash U numericalrelativity collaboration. We also provide URL addresses for web pages containinggraphics and movies of some of our results.To conclude this brief introduction, a statement of where we stand in termsof simulating general relativistic compact objects is in order. The NSF black holegrand challenge project and relat<strong>ed</strong> work achiev<strong>ed</strong> long term stable evolution ofsingle black hole spacetimes under certain conditions [6–8], but there is still along way to go before the spiralling coalescence can be comput<strong>ed</strong>. The presentlyon-going NASA neutron star grand challenge project recently succe<strong>ed</strong><strong>ed</strong> inevolving grazing collision of two neutron stars using the full Einstein-relativistichydrodynamic system of equations, with a simple equation of state, and theJapanese groups also report preliminary success in evolving several orbits witha fully relativistic GR-hydro code [9]. The recently fund<strong>ed</strong> EU Network [5] willcontinue on the momentum of these projects. However, the final goal of a fullsolution of the problem including radiation transport and magnetohydrodynamicsfor comparison between numerical simulations and observations in gravitationalwaveastronomy (waveform templates) and high-energy astronomy (γ -ray bursts)will take many more years, hopefully building on the effort describ<strong>ed</strong> in thispaper. So although much work has been done, much more remains, and newcommunity tools, including Cactus, are being made available to all groups. Weare very optimistic about the future, and hope to see more involvement in thiseffort across the relativity and astrophysics communities.18.2 Einstein equations for relativityThe generality and complexity of the Einstein equations make them an excellentand fertile testing ground for a variety of broadly significant computing issues.They form a system of dozens of coupl<strong>ed</strong>, nonlinear equations, with thousands of


364 Numerical relativityterms, of mix<strong>ed</strong> hyperbolic-elliptic type, and even undefin<strong>ed</strong> types, depending oncoordinate conditions. This rich and general structure of the equations implies thatthe techniques develop<strong>ed</strong> to solve our problems will be imm<strong>ed</strong>iately applicable toa large family of diverse scientific applications.The system of equations breaks up naturally into a set of constraintequations, which are elliptic in nature, evolution equations, which are‘hyperbolic’ in nature (more on this below), and gauge equations, which canbe chosen arbitrarily (often leading to more elliptic equations). The evolutionequations guarantee (mathematically) that the elliptic constraints are satisfi<strong>ed</strong> atall times provid<strong>ed</strong> the initial data satisfi<strong>ed</strong> them. This implies that the initialdata are not freely specifiable. Moreover, although the constraints are satisfi<strong>ed</strong>mathematically during evolution, it will not be so numerically. These problemsare discuss<strong>ed</strong> in turn below. First, however, we point out that a much simplertheory, familiar to many, has all of these same features. Maxwell’s equationsdescribing electromagnetic radiation have: (a) elliptic constraint equations,demanding that in vacuum the divergence of the electric and magnetic fieldsvanish at all times; (b) evolution equations, determining the time development ofthese fields, given suitable initial data satisfying the elliptic constraint equations;and (c) gauge conditions that can be appli<strong>ed</strong> freely to certain variables in thetheory, such as some components of the vector potential. Some choices of vectorpotential lead to hyperbolic evolution equations for the system, and some do not.We will find all of these features present in the much more complicat<strong>ed</strong> Einsteinequations, so it is useful to keep Maxwell’s equations in mind when reading thenext sections.In the standard 3 + 1 ADM approach to general relativity [10], the basicbuilding block of the theory—the spacetime metric—is written in the formds 2 =−(α 2 − β a β a ) dt 2 + 2β a dx a dt + γ ab dx a dx b , (18.1)using geometriz<strong>ed</strong> units such that the gravitational constant G and the spe<strong>ed</strong> oflight c are both equal to unity. Throughout this paper, we use Latin indices tolabel spatial coordinates, running from 1 to 3. The ten functions (α, β a ,γ ab )are functions of the spatial coordinates x a and time t. Indices are rais<strong>ed</strong> andlower<strong>ed</strong> by the ‘spatial 3-metric’ γ ab . Notice that the geometry on a 3D spacelikehypersurface of constant time (i.e. dt = 0) is determin<strong>ed</strong> by γ ab . As we will seebelow, the Einstein equations control the evolution in time of this 3D geometrydescrib<strong>ed</strong> by γ ab , given appropriate initial conditions. The lapse function αand the shift vector β a determine how the slices are thread<strong>ed</strong> by the spatialcoordinates. Together, α and β a represent the coordinate degrees of fre<strong>ed</strong>ominherent in the covariant formulation of Einstein’s equations, and can therefore bechosen, in some sense, ‘freely’, as discuss<strong>ed</strong> below.This formulation of the equations assumes that one begins with aneverywhere spacelike slice of spacetime, that should be evolv<strong>ed</strong> forward in time.Due to limit<strong>ed</strong> space, we will not discuss promising alternate treatments, bas<strong>ed</strong>


Einstein equations for relativity 365on either characteristic, or null foliations of spacetime [11], or on asymptoticallynull slices of spacetime [12–15].18.2.1 Constraint equationsThe constraints can be consider<strong>ed</strong> as the relativistic generalization of the Poissonequation of Newtonian gravity, but instead of a single linear elliptic equationthere are now four, coupl<strong>ed</strong>, highly nonlinear elliptic equations, known as theHamiltonian and momentum constraints. Under certain conditions, the equationsdecouple and can be solv<strong>ed</strong> independently and more easily, and this is how theyhave usually been treat<strong>ed</strong>. Recently, techniques have been develop<strong>ed</strong> that allowone to solve the constraints in a more general setting, without making restrictiveassumptions that lead to decoupling [16–19]. In such a system the four constraintequations are solv<strong>ed</strong> simultaneously. This may prove useful in generating newclasses of initial data. However, at present there is no satisfactory algorithm forcontrolling the physics content of the data generat<strong>ed</strong>. The major remaining workin this direction is to develop a scheme that is capable of constructing the initialdata that describe a given physical system. That is, although we have schemesavailable to solve many variations on the initial value problem, it is difficult tospecify in advance, for example, what are the precise spins and momenta of twoblack holes in orbit, or even if the holes are in orbit. This can generally only b<strong>ed</strong>etermin<strong>ed</strong> after the equations have been solv<strong>ed</strong> and analys<strong>ed</strong>.The elliptic operators for these equations are usually symmetric, but they areotherwise the most general type, with all first and mix<strong>ed</strong> second derivative termspresent. The boundary conditions, which can break the symmetry, are usuallylinear conditions that involve derivatives of the fields being solv<strong>ed</strong>. In any case,once the initial value equations have been solv<strong>ed</strong>, initial data for the evolutionproblem result.We illustrate the central idea of constructing initial data with vacuumspacetimes for simplicity. The application of the algorithm present<strong>ed</strong> here to ageneral spacetime with matter source is currently routine in numerical relativity.The full 4D Einstein equations can be decompos<strong>ed</strong> into six evolution equationsand four constraint equations. The constraints may be subdivid<strong>ed</strong>, in turn, intoone Hamiltonian (or energy) constraint equation,R + (tr K ) 2 − K ab K ab = 0, (18.2)and three momentum constraint equations (or one vector equation),D b (K ab − γ ab tr K ) = 0. (18.3)In these equations K ab is the extrinsic curvature of the slice, relat<strong>ed</strong> to the tim<strong>ed</strong>erivative of γ ab byK ab =− 12α (∂ t γ ab − D a β b − D b β a ). (18.4)


366 Numerical relativityHere we have introduc<strong>ed</strong> the 3D spatial covariant derivative operator D aassociat<strong>ed</strong> with the 3-metric γ ab (i.e. D a γ bc = 0), and the 3D scalar curvatureR comput<strong>ed</strong> from γ ab . These four constraint equations can be us<strong>ed</strong> to determineinitial data for γ ab and K ab , which are to be evolv<strong>ed</strong> with the evolution equationsdiscuss<strong>ed</strong> below. These equations (18.2) and (18.3) are referr<strong>ed</strong> to as constraintsbecause, as in the case of electrodynamics, they contain no time derivatives of thefundamental fields γ ab and K ab , and hence do not propagate the solution in time.Next, we will sketch the standard method for obtaining a solution to theseconstraint equations. We follow York and coworkers (e.g., [20]) by writing the3-metric and extrinsic curvature in ‘conformal form’, and also make use of thesimplifying assumption tr K = 0 which causes the Hamiltonian and momentumconstraints to completely decouple (note that actually the equations decoupl<strong>ed</strong>with tr K = constant but we will discuss only the simplest case here). We writeγ ab = 4 ˆγ ab , K ab = −2 ˆK ab , (18.5)where ˆγ ab and the transverse-tracefree part of ˆK ab is regard<strong>ed</strong> as given, i.e. chosento represent the physical system that we want to study. Under the conformaltransformation, with tr K = 0wefind that the momentum constraint becomesˆD b ˆK ab = 0, (18.6)where ˆD a is the 3D covariant derivative associat<strong>ed</strong> with ˆγ ab (i.e. ˆD a ˆγ ab = 0).In vacuum, black hole spacetimes ˆK ab can often be solv<strong>ed</strong> analytically. Formore details on how to solve the momentum constraints in complicat<strong>ed</strong> situations,see [10, 21, 22].The remaining unknown function , must satisfy the Hamiltonianconstraint. The conformal transformation of the scalar curvature isR = −4 ˆR − 8 −5 ˆ, (18.7)where ˆ = ˆγ ab ˆD a ˆD b and ˆR is the scalar curvature of the known metric ˆγ ab .Plugging this back into the Hamiltonian constraint and dividing through by−8 −5 , we obtainˆ − 1 8 ˆR + 1 8 −7 ( ˆK ab ˆK ab ) = 0, (18.8)an elliptic equation for the conformal factor .To summarize, one first specifies ˆγ ab and the transverse-tracefree part ofˆK ab ‘at will’, choosing them to be something ‘closest’ to the spacetime onewants to study. Then one solves (18.6) for the conformal extrinsic curvatureˆK ab . Finally, (18.8) is solv<strong>ed</strong> for the conformal factor , so the full solution γ aband K ab can be reconstruct<strong>ed</strong>. In this process the elliptic equations are solv<strong>ed</strong>by standard techniques, for example, the conjugate gradient [23] or multigridmethods [24]. In situations where there is a black hole singularity, there couldbe add<strong>ed</strong> complications in solving the elliptic equations, and special treatments


Einstein equations for relativity 367would have to be introduc<strong>ed</strong>, for example, the ‘puncture’ treatment of [25], oremploying an ‘isometry’ operation to provide boundary conditions on black holethroats, ensuring identical spatial geometries inside and outside the throat (see,e.g., [21, 26], or [27] for more details).While this is a well-establish<strong>ed</strong> process for generating an initial data setfor numerical study, there is a fundamental difficulty in using this approach togenerate initial data corresponding to a physical system one wants to evolve, forexample, a coalescing binary system. It is not clear how to choose the ‘closest’ˆγ ab , and the corresponding free components in ˆK ab , so that the resulting γ ab andK ab represent the inspiralling system at its late stage of inspiral. This late stage isthe so-call<strong>ed</strong> ‘interm<strong>ed</strong>iate challenge problem’ of binary black holes [28], an areaof much current interest.18.2.2 Evolution equations18.2.2.1 The standard evolution systemWith the initial data γ ab and K ab specifi<strong>ed</strong>, we now consider their evolution intime. There are six evolution equations for the 3-metric γ ab that are secondorder in time, resulting from projections of the full 4D Einstein equations ontothe 3D spacelike slice [10]. These are most often written as a first-order in-timesystem of twelve evolution equations, usually referr<strong>ed</strong> to as the ‘ADM’ evolutionsystem [10, 29]:∂ t γ ab =−2αK ab + D a β b + D b β a (18.9)∂ t K ab =−D a D b α + α[R ab + (tr K )K ab − 2K ac K c b]+ β c D c K ab + K ac D b β c + K cb D a β c . (18.10)Here R ab is the Ricci tensor of the 3D spacelike slice labell<strong>ed</strong> by a constant valueof t. Note that these are quantities defin<strong>ed</strong> only on a t = constant hypersurface,and require only the 3-metric γ ab in their construction. Do not confuse them withthe conventional 4D objects! The complete set of Einstein equations are contain<strong>ed</strong>in constraint equations (18.2), (18.3) and the evolution equations (18.10), (18.9).Note that (18.9) is simply the definition of the extrinsic curvature K ab (18.4).These equations are analogous to the evolution equations for the electric andmagnetic fields of electrodynamics. Given the ‘lapse’ α and ‘shift’ β a , discuss<strong>ed</strong>below, they allow one to advance the system forward in time.18.2.2.2 Hyperbolic evolution systemsThe evolution equations (18.10) and (18.9) have been present<strong>ed</strong> in the ‘standardADM form’, which has serv<strong>ed</strong> numerical relativity well over the last fewdecades. However, the equations are enormously complicat<strong>ed</strong>; the complicationis hidden in the definition of the curvature tensor R ab and the covariantdifferentiation operator D a . In particular, although they describe physical


368 Numerical relativityinformation propagating with a finite spe<strong>ed</strong>, the system does not form a hyperbolicsystem, and is not necessarily the best for numerical evolution. Other fieldsof physics, in particular hydrodynamics, have develop<strong>ed</strong> very mature numericalmethods that are specially design<strong>ed</strong> to treat the well studi<strong>ed</strong> flux conservative,hyperbolic system of balance laws having the form∂ t u + ∂ k F k − u = S − u (18.11)where the vector u displays the set of variables and both ‘fluxes’ F k and ‘sources’S are vector valu<strong>ed</strong> functions. In hydrodynamic systems, it often turns out thatthe characteristic matrix ∂ F/∂ u project<strong>ed</strong> into any spacelike direction can oftenbe diagonaliz<strong>ed</strong>, so that fields with definite propagation spe<strong>ed</strong>s can be identifi<strong>ed</strong>(the eigenvectors and the eigenvalues of the project<strong>ed</strong> characteristic matrix). Oneimportant point is that in (18.11) all spatial derivatives are contain<strong>ed</strong> in the fluxterms, with the source terms in the equations containing no derivatives of theeigenfields. All of these features can be exploit<strong>ed</strong> in numerical finite differenceschemes that treat each term in an appropriate way to preserve important physicalcharacteristics of the solution.Amazingly, the complete set of Einstein equations can also be put inthis ‘simple’ form (the source terms still contain thousands of terms however).Building on earlier work by Choquet-Bruhat and Ruggeri [30], Bona and Massóbegan to study this problem in the late 1980s, and by 1992 they had develop<strong>ed</strong>a hyperbolic system for the Einstein equations with a certain specific gaugechoice [31] (see below). Here by hyperbolic, we mean simply that the project<strong>ed</strong>characteristic matrix has a complete set of eigenfields with real eigenvalues.This work was generaliz<strong>ed</strong> recently to apply to a large family of gauge choices[32, 33]. The Bona–Massó system of equations is available in the 3D ‘Cactus’code [34,35], as is the standard ADM system, where both are test<strong>ed</strong> and compar<strong>ed</strong>on a number of spacetimes.The Bona–Massó system is now one among many hyperbolic systems,as other independent hyperbolic formulations of Einstein’s equations wer<strong>ed</strong>evelop<strong>ed</strong> [36–41] at about the same time as [42]. Among these otherformulations only the one originally devis<strong>ed</strong> in [38] has been appli<strong>ed</strong> tospacetimes containing black holes [43], although still only in the sphericallysymmetry 1D case (a 3D version is under development [44].) Hence, of themany hyperbolic variants, only the Bona–Massó family and the formulationsof York and co-workers have been test<strong>ed</strong> in any detail in 3D numerical codes.Notably among the differences in the formulations, the Bona–Massó and Frittellifamilies contain terms equivalent to second time derivatives of the three metricγ ab , while many other formulations go to a higher time derivative to achievehyperbolicity. Another comment worth making is that for harmonic slicing, boththe Bona–Massó and York families have characteristic spe<strong>ed</strong>s of either zero, orlight spe<strong>ed</strong>. For maximal slicing, they both r<strong>ed</strong>uce to a coupl<strong>ed</strong> elliptic-hyperbolicsystem. The Bona–Massó system (at least) also allows for an additional familyof explicit algebraic slicings, with the lapse proportional to an explicit function


Still newer formulations: towards a stable evolution system 369of the determinant of the 3-metric, and in those cases one can also identify gaugespe<strong>ed</strong>s which can be different from light spe<strong>ed</strong> (harmonic slicing is one exampleof this family where the gauge spe<strong>ed</strong> corresponds to light spe<strong>ed</strong>). Some ofthese slicings, such as ‘1 + log’ [45], have been found to be very useful in 3Dnumerical evolutions. This information about the spe<strong>ed</strong> of gauge and physicalpropagation can be very helpful in understanding the system, and can also beuseful in developing numerical methods. Only extensive numerical studies willtell if the various hyperbolic formulations live up to their promise.Reula has recently review<strong>ed</strong>, from the mathematical point of view, most ofthe recent hyperbolic formulations of the Einstein equations [46] (This article, inthe online journal Living Reviews in Relativity, will be periodically updat<strong>ed</strong>). Itis important to realize that the mathematical relativity field has been interest<strong>ed</strong>in hyperbolic formulations of the Einstein equations for many years and somesystems that could have been suitable for numerical relativity were alreadypublish<strong>ed</strong> in the 1980s [30, 47]. However, these developments were generallynot recogniz<strong>ed</strong> by the numerical relativity community until recently.18.3 Still newer formulations: towards a stable evolutionsystemSomewhere in between the standard ADM formulation and hyperbolicformulations are a class of formulations that have been getting significantattention lately, and which seem to be very promising and very stable.As discuss<strong>ed</strong> above, the 3D evolution of Einstein’s equations has prov<strong>ed</strong> verydifficult, with instabilities developing on rather short timescales, even in cases ofweakly gravitating, vacuum systems, such as low amplitude gravitational <strong>waves</strong>,as summariz<strong>ed</strong> in an important paper by Baumgarte and Shapiro [48]. In thiswork, it was shown how one can achieve highly improv<strong>ed</strong> stability by making afew key changes to the formulation of the ADM equations, most notably througha conformal decomposition and by rewriting certain terms in the 3D Ricci tensorto eliminate terms that spoil its elliptic nature. In fact, essentially these sametricks were already notic<strong>ed</strong> a few years earlier by Shibata and Nakamura [49].Hence, we refer to these formulations collectively as ‘BSSN’ after the fourauthors. These subtle changes to the standard ADM formalism have a verypowerful stabilizing effect on the evolutions. Evolutions of weak <strong>waves</strong> thatwould develop instabilities and crash with the standard ADM formulation runmuch longer with the new system, and as shown in Alcubierre et al [50], the newsystem and variations allow for the first time the successful evolution of highlynonlinear gravitational <strong>waves</strong> to form a black hole in 3D while the standard ADMtreatment would fail well in advance of black hole formation. Further work by thePalma group, show<strong>ed</strong> the deep connection between the BSSN formulations andthe Bona–Massó family of formulations [51], leading to the possibility of a fullyhyperbolic, very stable formulation that shares advantages from many sides.


370 Numerical relativityThe Palma, Potsdam and Wash U groups also show<strong>ed</strong> that these newformulations lead to much more stable black hole evolutions as well. Whilestandard ADM formulations can evolve black holes very accurately for a shortperiod of time, as describ<strong>ed</strong> above, large peaks in metric functions caus<strong>ed</strong> by socall<strong>ed</strong>‘grid-stretching’ develop instabilities, which cause the codes to crash fartoo soon to study orbits of black holes. The new formulations can significantlyextend the evolution times (by factors of two or much more) that can be achiev<strong>ed</strong>.In all cases, the evolutions are convergent, but seem to have larger error than thestandard ADM or Bona–Massó system. These effects were recently analys<strong>ed</strong> in apaper by Alcubierre et al [52]. We are now in the process of applying these newformulations to a series of interesting spacetimes, including pure gravitational<strong>waves</strong>, black holes, and neutron stars, some results of which are report<strong>ed</strong> below.In this and a companion paper [52] we focus on an alternative approach bas<strong>ed</strong>on a conformal decomposition of the metric and the trace-free components ofthe extrinsic curvature. The conformal-tracefree (CT) approach was first devis<strong>ed</strong>by Nakamura in the 1980s in 3D calculations [22, 53], and then modifi<strong>ed</strong> andappli<strong>ed</strong> to work on gravitational <strong>waves</strong> [49], and on neutron stars [9, 54]. Thisapproach was not taken up by others in the community until a recent paper byBaumgarte and Shapiro [48], where a similar formulation was compar<strong>ed</strong> with thestandard ADM approach and shown to be superior, in terms of both accuracy andstability, on tests involving weak gravitational <strong>waves</strong>, with geodesic and harmonicslicing. In a follow-up paper, Baumgarte, Hughes, and Shapiro [55] appli<strong>ed</strong> thesame formulation to systems with given (analytically prescrib<strong>ed</strong>) matter sources,and found similar stability properties. More recently fully hydrodynamicalsimulations employing the CT approach have been report<strong>ed</strong> in [56–58] in thecontext of collapse of rapidly-rotating (isolat<strong>ed</strong>) neutron stars and coalescenceand merge of binary neutron stars.In the companion paper [52] we perform an analytic investigation of thestability properties of the ADM and the CT evolution equations. Using alineariz<strong>ed</strong> plane wave analysis, we identify features of the equations that webelieve are responsible for the difference in their stability properties.18.3.0.1 Numerical techniques for the evolution equationsMost of what has been attempt<strong>ed</strong> in numerical relativity evolution schemes is builton explicit finite difference schemes. Implicit and iterative evolution schemeshave been occasionally attempt<strong>ed</strong>, but the extra cost associat<strong>ed</strong> has made themless popular. We now describe the basic approach that has been tri<strong>ed</strong> for boththe standard ADM formulation and more recent hyperbolic formulations of theequations.


ADM evolutionsStill newer formulations: towards a stable evolution system 371The ADM system of evolution equations is often solv<strong>ed</strong> using some variationof the leapfrog method, similar to that describ<strong>ed</strong> in [59]. The most extensivelytest<strong>ed</strong> is the ‘stagger<strong>ed</strong> leapfrog’, detail<strong>ed</strong> in axisymmetric cases in [59] and in 3Din [45], but other successful versions include full leapfrog implementations us<strong>ed</strong>in 3D by [60] and [34]. For the ADM system, the basic strategy is to use centr<strong>ed</strong>spatial differences everywhere, march forward according to some explicit timescheme, and hope for the best! Generally, this technique has work<strong>ed</strong> surprisinglywell until large gradients are encounter<strong>ed</strong>, at which time the methods often breakdown. The problem is that the equations in this ADM form are difficult to analyse,and hence ad hoc numerical schemes are often tri<strong>ed</strong> without detail<strong>ed</strong> knowl<strong>ed</strong>geof how to treat specific terms in the equations, or how to treat instabilities whenthey arise. A recent development is that of the ‘delous<strong>ed</strong>’ leapfrog, which amountsto filtering the solution [61].Also recently, the iterative Crank–Nicholson (ICN) scheme has been foundeffective in suppressing some instabilities that occur [62–64]. ICN is an iterative,explicit version of the standard implicit Crank–Nicholson (CN) scheme [61, 65].The idea behind this method is to solve the implicit equations by an iterativeproc<strong>ed</strong>ure, where each iteration is an explicit operation depending only onpreviously comput<strong>ed</strong> data. Normally, this process is stopp<strong>ed</strong> after a certainnumber of iterations, or until some tolerance is achiev<strong>ed</strong>. For a linear equation(and, in particular, in one dimension), the iterative proc<strong>ed</strong>ure can easily be muchmore computationally expensive than the matrix inversion requir<strong>ed</strong> to solve theoriginal implicit scheme. For a nonlinear system, however, solving the implicitscheme directly can prove to be extremely difficult.The stability properties of the ICN scheme have been studi<strong>ed</strong>, with theseimportant results.• In order to obtain a stable scheme one must do at least three iterations, andnot just the two one would normally expect (two iterations are enough toachieve second-order accuracy, but they are unstable!).• The iterative scheme itself is only convergent if the standard Courant–Fri<strong>ed</strong>richs–Lewy (CFL) stability condition is satisfi<strong>ed</strong>, otherwise theiterations diverge.These two results taken together imply that there is no reason (at least fromthe point of view of stability) to ever do more that three ICN iterations. Threeiterations are already second-order accurate, and provide us with a (conditionally)stable scheme. Increasing the number of iterations will not improve the stabilityproperties of the scheme any further. In particular, we will never achieve theunconditional stability properties of the full implicit CN scheme, since if weviolate the CFL condition the iterations will diverge.


372 Numerical relativityHyperbolic evolutionsThe hyperbolic formulations are on a much firmer footing numerically than theADM formulation, as the equations are in a much simpler form that has beenstudi<strong>ed</strong> for many years in computational fluid dynamics. However, the applicationof such methods to relativity is quite new, and hence the experience with suchmethods in this community is relatively limit<strong>ed</strong>. Furthermore, the treatment of thehighly nonlinear source terms that arise in relativity is very much unexplor<strong>ed</strong>, andthe source terms in Einstein’s equations are much more complicat<strong>ed</strong> than those inhydrodynamics. Here we will just discuss the basic ideas in such schemes.A standard technique for equations having flux conservative form is to splitequation (18.11) into two separate processes. The transport part is given by theflux terms∂ t u + ∂ k F− k u = 0. (18.12)The source contribution is given by the following system of ordinary differentialequations∂ t u = S − u. (18.13)Numerically, this splitting is perform<strong>ed</strong> by a combination of both flux and sourceoperators. Denoting by E(t) the numerical evolution operator for system(18.11) in a single timestep, we implement the following combination sequenceof subevolution steps:E(t) = S(t/2)T (t)S(t/2) (18.14)where T , S are the numerical evolution operators for systems (18.12) and (18.13),respectively. This is known as ‘Strang splitting’ [66]. As long as both operatorsT and S are second-order accurate in t, the overall step of operator E is alsosecond-order accurate in time.This choice of splitting allows easy implementation of different numericaltreatments of the principal part of the system without having to worry aboutthe sources of the equations. Additionally, there are numerous computationaladvantages to this technique, as discuss<strong>ed</strong> in [67].The sources can be updat<strong>ed</strong> using a variety of ODE integrators, and in‘Cactus’ the usual technique involves second-order pr<strong>ed</strong>ictor-corrector methods.Higher order methods for source integration can be easily implement<strong>ed</strong>, but thiswill not improve the overall order of accuracy. However, in special cases wherethe evolution is largely source driven [68], it may be important to use higher ordersource operators, and this method allows such generalizations. The details can befound in [34].The implementation of numerical methods for the flux operator is muchmore involv<strong>ed</strong>, and there are many possibilities, ranging from standard choicesto advanc<strong>ed</strong> shock capturing methods [33, 69, 70]. Among standard methods, theMacCormack method, which has proven to be very robust in the computationalfluid dynamics field (see, e.g., [71] and references therein), and a directionally


Still newer formulations: towards a stable evolution system 373split Lax-Wendroff method have been implement<strong>ed</strong> and test<strong>ed</strong> extensively in‘Cactus’. These schemes are fully second order in space and time. Shockcapturing methods have been shown to work extremely well in 1D problemsin numerical relativity [32, 70], but their application in 3D is an active researcharea full of promise, but as yet, unfulfill<strong>ed</strong>. The details of these methods, asthey are appli<strong>ed</strong> to the Bona–Massó formulation of the equations, can be foundin [34, 70].18.3.0.2 The role of constraintsIf the constraints are satisfi<strong>ed</strong> on the initial hypersurface, the evolution equationsthen guarantee that they remain satisfi<strong>ed</strong> on all subsequent hypersurfaces. Thus,once the initial value problem has been solv<strong>ed</strong>, one may advance the solutionforward in time by using only the evolution equations. This is the same situationencounter<strong>ed</strong> in electrodynamics as discuss<strong>ed</strong> before. However, in a numericalsolution, the constraints will be violat<strong>ed</strong> at some level due to numerical error.They hence provide useful indicators for the accuracy of the numerical spacetimesgenerat<strong>ed</strong>. Traditional alternatives to this approach, which is often referr<strong>ed</strong> toas ‘free evolution’, involve solving some or all of the constraint equations oneach slice for certain metric and extrinsic curvature components, and then simplymonitoring the ‘left over’ evolution equations. This issue is discuss<strong>ed</strong> further byChoptuik in [72], and in detail for the Schwarzschild spacetime in [73]. Newapproaches to this problem of constraint versus evolution equations are currentlybeing pursu<strong>ed</strong> by Lee [74], among others [75]. This approach is to advancethe system forward using the evolution equations, and then adjust the variablesslightly so that the constraints are satisfi<strong>ed</strong> (to some tolerance), i.e. the solutionis project<strong>ed</strong> onto the constraint surface. Because there are many variables thatgo into the constraints, there is not a unique way to decide which ones to adjustand by how much. However, one can compute the ‘minimum’ perturbation tothe system, which corresponds to projecting to the closest point on the constraintsurface. Other approaches, similar in spirit to each other, have been suggest<strong>ed</strong>by Detweiler [76] and Brodbeck et al [77]. The Detweiler approach restricts thenumerical evolution to the constraint surface by adding terms to the evolutionequations (18.9) and (18.10) terms which are proportional to the constraints.Numerical tests of the scheme using gravitational wave spacetimes have recentlybeen carri<strong>ed</strong> out, showing promising results [78].18.3.0.3 Gauge conditionsKinematic conditions for the lapse function α and shift vector β i have to bespecifi<strong>ed</strong> for the evolution equations (18.9) and (18.10). With γ ab and K absatisfying the constraint on the initial slice, the lapse and shift can be chosenarbitrarily on the initial slice and thereafter. These are referr<strong>ed</strong> to as gaugechoices, analogous to the choice of the gauge function in electrodynamics.


374 Numerical relativityEinstein did not specify these quantities; they are up to the numerical relativist tochoose at will!LapseThe choice of lapse corresponds to how one chooses 3D spacelike hypersurfacesin 4D spacetime. The ‘lapse’ of proper time along the normal vector of one sliceto the next is given by α dt, where dt is the coordinate time interval betweenslices. As α(x, y, z) can be chosen at will on a given slice, some regions ofspacetime can be made to evolve farther into the future than others.There are many motivations for particular choices of lapse. A primaryconcern is to ensure that it leads to a stable long-term evolution. It is easy tosee that a naive choice of the lapse, for example, α = 1, the so-call<strong>ed</strong> geodesicslicing, suffers from a strong tendency to produce coordinate singularities [79,80].A relat<strong>ed</strong> concern is that one would like to cover the region of interest in anevolution, say, where gravitational <strong>waves</strong> generat<strong>ed</strong> by some process could b<strong>ed</strong>etect<strong>ed</strong>, while avoiding troublesome regions, say, inside black holes wheresingularities lurk (the so-call<strong>ed</strong> ‘singularity avoiding’ time slicings). Anotherimportant motivation is that some choices of α allow one to write the evolutionequations in forms that are especially suit<strong>ed</strong> to numerical evolution. Finally,computational considerations also play an important role in the choice of thelapse; one prefers a condition for α that does not involve great computationalexpense, while also providing smooth, stable evolution.Some ‘traditional’ choices of the lapse us<strong>ed</strong> in the numerical constructionof spacetimes are [81]: (1) Lagrangian slicing, in which the coordinates arefollowing the flow of the matter in the simulation. This choice simplifies thematter evolution equations, but it is not always applicable, for example, in avacuum spacetime or when the fluid flow pattern becomes complicat<strong>ed</strong>. (2)Maximal slicing, [79, 80] in which the trace of the extrinsic curvature is requir<strong>ed</strong>to be zero always, i.e, K (t = 0) = 0 = ∂ t K . The evolution equations of theextrinsic curvature then lead to an elliptic equation for the lapseD a D a α − α(R + K 2 ) = 0. (18.15)The maximal slicing has the nice property of causing the lapse to ‘collapse’ toa small value at regions of strong gravity, hence avoiding the region where acurvature singularity is forming. It is one of the so-call<strong>ed</strong> ‘singularity avoidingslicing conditions’. Maximal slicing is easily the most studi<strong>ed</strong> slicing conditionin numerical relativity. (3) Constant mean curvature, where we let K = constantdiffer from zero, a choice often us<strong>ed</strong> in constructing cosmological solutions.(4) Algebraic slicing, where the lapse is given as an algebraic function of th<strong>ed</strong>eterminant of the 3-metric. Algebraic slicing can also be singularity avoiding[82]. As there is no ne<strong>ed</strong> to solve an elliptic equation as in the case of maximalslicing, algebraic slicing is computationally efficient. Some algebraic slicings(e.g., the harmonic slicing in which α is set proportional to the square root of


Still newer formulations: towards a stable evolution system 375the determinant of the 3-metric g ab ) also make the mathematical structure of theevolution equations simpler. However, the local nature of the choice of the lapsecould lead to noise in the lapse [45] and the formation of ‘shock’-like features innumerical evolutions [83, 84]. The former problem can be dealt with by turningthe algebraic slicing equation to an evolution equation with a diffusion term [45],but the latter problem does not seem to have a simple solution.In addition to these most widely us<strong>ed</strong> ‘traditional’ choices of the lapse,there are also some newly develop<strong>ed</strong> slicing conditions whose use in numericalrelativity though promising remain largely unexplor<strong>ed</strong> [85]: (5) K-driver. This isa generalization of the maximal slicing in which the extrinsic curvature, insteadof being set to zero, is requir<strong>ed</strong> to satisfy the condition∂ t K =−cK, (18.16)where c is some positive constant. This was first brought up by Eppley [86] andrecently investigat<strong>ed</strong> in [87]. In this way the trace of the extrinsic curvature, whennumerical inaccuracy causes it to drift away from zero, is ‘driven’ back to zeroexponentially. When combin<strong>ed</strong> with the evolution equations, (18.16) again leadsto an elliptic equation for the lapse. This choice of the lapse is shown in [87] tolead to a much more stable numerical evolution in cases where one wants to avoidlarge values of the extrinsic curvature. The optimal choice of the constant c aswell as a number of variations on this ‘driver’ scheme are presently being studi<strong>ed</strong>.(6) γ -driver. This is another use of the ‘driver’ idea. In this case, the time rateof change of the determinant of the three metric det(g ab ) is driven to zero [87].In the absence of a shift vector or if the shift has zero divergence, this r<strong>ed</strong>ucesto the K-driver. This choice of the lapse, which has the unique property of beingable to respond to the choice of the shift, demands extensive investigations andevaluations.ShiftThe shift vector describes the ‘shifting’ of the coordinates from the normal vectoras one moves from one slice to the next. If the shift vanishes, the coordinate point(x, y, z) will move normal to a given 3D time slice to the next slice in the future.(Refer to York [10] or Cook [88], for details and diagrams.) The choice of shiftis perhaps less well develop<strong>ed</strong> than the choice of lapse in numerical relativity,and many choices ne<strong>ed</strong> to be explor<strong>ed</strong>, particularly in 3D. The main purpose ofthe shift is to ensure that the coordinate description of the spacetime remainswell behav<strong>ed</strong> throughout the evolution. With an inappropriate or poorly chosenshift, coordinate lines may move toward each other, or become very stretch<strong>ed</strong>or shear<strong>ed</strong>, leading to pathological behaviour of the metric functions that may b<strong>ed</strong>ifficult to handle numerically. It may even cause the code to crash if, for example,two coordinate lines ‘touch’ each other creating a ‘coordinate singularity’ (i.e.the metric becomes singular as the distance ds between two coordinate linesgoes to zero). Two important considerations for appropriate shift conditions


376 Numerical relativityare the ability to prevent large shearing or drifting of coordinates during anevolution, and the ability to control the coordinate location of a physical object,for example, the horizon of a black hole. These considerations are discuss<strong>ed</strong>below. The development of appropriate shift conditions for full 3D evolution,for systems without symmetries, is an important research area that ne<strong>ed</strong>s muchattention. Geometrical shift conditions that can be formulat<strong>ed</strong> without referenceto specific coordinate systems or symmetries seem to be desirable. The basic ideais to develop a condition that minimizes the stretching, shearing, and driftingof coordinates in a general way. A few examples have been devis<strong>ed</strong> whichpartially meet these goals, such as ‘minimal distortion’, ‘minimal strain’ andvariations [10], but much more investigations are ne<strong>ed</strong><strong>ed</strong>. New gauge conditions,bas<strong>ed</strong> on these earlier proposals, have recently been propos<strong>ed</strong> but not yet test<strong>ed</strong>in numerical simulations [28].It is important to emphasize that the lapse and shift only change the wayin which the slices are chosen through a spacetime and where coordinates arelaid down on every slice, and do not, in principle, affect any physical resultswhatsoever. They will affect the value of the metric quantities, but not the physicsderiv<strong>ed</strong> from them. In this respect the fre<strong>ed</strong>om of choice in the lapse and shift isanalogous to the fre<strong>ed</strong>om of gauge in electromagnetic systems.On the other hand, it is also important to emphasize that proper choices oflapse and shift are crucial for the numerical construction of a spacetime in theEinstein theory of general relativity, in particular in a general 3D setting. In ageneral 3D simulation without symmetry assumption, there is no preferr<strong>ed</strong> choiceof the form of the metric (e.g., a diagonal 3-metric, or g θθ = r 2 as in sphericalsymmetry), hence forcing us to deal with the gauge degree of fre<strong>ed</strong>om in relativityin full. This, when coupl<strong>ed</strong> with the inevitable lower resolution in 3D simulations,often leads to development of coordinate singularities, when evolv<strong>ed</strong> without asophisticat<strong>ed</strong> choice of lapse and shift. Inde<strong>ed</strong> the success of the ‘driver’ ideasuggest<strong>ed</strong> [87] that in order to obtain a stable evolution over a long timescale, itis important to ensure that the coordinate conditions us<strong>ed</strong> are not only suitablefor the geometry of the spacetime being evolv<strong>ed</strong>, but also that the conditionsthemselves are stable. That is, when the condition is perturb<strong>ed</strong>, for example,by numerical inaccuracy, there is no long-term secular drifting. We regard theconstruction of an algorithm for choosing a suitable lapse and shift for a general3D numerical simulation to be one of the most important issues facing numericalrelativity at present.18.3.1 General relativistic hydrodynamicsIn order to make numerical relativity a tool for computational general relativisticastrophysics, it is important to combine numerical relativity with traditional toolsof computational astrophysics, and, in particular, relativistic hydrodynamics.While a large amount of 3D studies in numerical relativity have been devot<strong>ed</strong>to the vacuum Einstein equations, the spacetime dynamics with a non-vanishing


Still newer formulations: towards a stable evolution system 377source term remains a large unchart<strong>ed</strong> territory. As astrophysics of compactobjects that ne<strong>ed</strong>s general relativity for its understanding is attracting increasingattention, general relativistic hydrodynamics will become an increasinglyimportant subject as astrophysicists begin to study more relativistic systems,as relativists become more involv<strong>ed</strong> in studies of astrophysical sources. Thispromises to be one of the most exciting and important areas of research inrelativistic astrophysics in the coming years.Previously, most work in relativistic hydrodynamics has been done on fix<strong>ed</strong>metric backgrounds. In this approximation the fluid is allow<strong>ed</strong> to move in arelativistic manner in strong gravitational fields, say around a black hole, but itseffect on the spacetime is not consider<strong>ed</strong>. Over the last years very sophisticat<strong>ed</strong>methods for general relativistic hydrodynamics have been develop<strong>ed</strong> by theValencia group l<strong>ed</strong> by José MIbáñez [89–92]. These methods are bas<strong>ed</strong> ona hyperbolic formulation of the hydrodynamic equations, and are shown to besuperior to traditional artificial viscosity methods for highly relativistic flows andstrong shocks.However, just fix<strong>ed</strong> background approximation is inadequate in describinga large class of problems which are of most interest to gravitational-waveastronomy, namely those with substantial matter motion generating gravitationalradiation, like the coalescences of neutron star binaries. We are constructinga multipurpose 3D code for the NASA Neutron Star Grand Challenge Project[93] that contains the full Einstein equations coupl<strong>ed</strong> to general relativistichydrodynamics. The hydrodynamic part consists of both an artificial viscositymodule, [94] and a module bas<strong>ed</strong> on modern shock capturing schemes [95],containing three hydroevolution methods [95]: a flux split method, Roe’sapproximate Riemann solver [96] and Marquina’s approximate Riemann solver[92, 97]. All are bas<strong>ed</strong> on finite-difference schemes employing approximateRiemann solvers to account explicitly for the characteristic information of theequations. These schemes are particularly suitable for astrophysics simulationsthat involve matter in (ultra)relativistic spe<strong>ed</strong>s and strong shock <strong>waves</strong>.In the flux split method, the flux is decompos<strong>ed</strong> into the part contributingto the eigenfields with positive eigenvalues (fields moving to the right) and thepart with negative eigenvalues (fields moving to the left). These fluxes are thendiscretiz<strong>ed</strong> with one sid<strong>ed</strong> derivatives (which side depends on the sign of theeigenvalue). The flux split method presupposes that the equation of state of thefluid has the form P = P(ρ, ɛ) = ρ f (ɛ), which includes, for example, theadiabatic equation of state. The second scheme, Roe’s approximate Riemannsolver [96] is by now a ‘traditional’ method for the integration of nonlinearhyperbolic systems of conservation laws [90, 91, 98]. This method makes noassumption on the equation of state, and, is more flexible than the flux splitmethods. The third method, the Marquina’s method, is a promising new scheme[97]. It is bas<strong>ed</strong> on a flux formula which is an extension of Shu and Osher’sentropy-satisfying numerical flux [99] to systems of hyperbolic conservationlaws. In this scheme there are no artificial interm<strong>ed</strong>iate states construct<strong>ed</strong> at


378 Numerical relativityeach cell interface. This implies that there are no Riemann solutions involv<strong>ed</strong>(either exact or approximate); moreover, the scheme has been prov<strong>ed</strong> to alleviateseveral numerical pathologies associat<strong>ed</strong> to the introduction of an averag<strong>ed</strong> state(as Roe’s method does) in the local diagonalization proc<strong>ed</strong>ure (see [92, 97]).For a detail<strong>ed</strong> comparison of the three schemes and their coupling to dynamicalevolution of spacetimes, see [95].The availability of the hyperbolic hydrotreatment and its coupling to thespacetime evolution code is particularly noteworthy. With the development ofa hyperbolic formulation of the Einstein equations describ<strong>ed</strong> above, the entiresystem can be treat<strong>ed</strong> as a single system of hyperbolic equations, rather thanartificially separating the spacetime part from the fluid part.18.3.2 Boundary conditionsAppropriate conditions for the outer boundary have yet to be deriv<strong>ed</strong> for 3Dnumerical relativity. In 1D and 2D relativity codes, the outer boundary isgenerally plac<strong>ed</strong> far enough away that the spacetime is nearly flat there, and staticor flat (i.e. copying data from the next-to-last zone to the outer <strong>ed</strong>ge) boundaryconditions can usually be specifi<strong>ed</strong> for the evolv<strong>ed</strong> functions. However, due tothe constraints plac<strong>ed</strong> on us by limit<strong>ed</strong> computer memory, this is not currentlypossible in 3D. Adaptive mesh refinement will be of great use in this regard, butwill not substitute for proper physical treatment. Most results to date have beencomput<strong>ed</strong> with the evolv<strong>ed</strong> functions kept static at the outer boundary, even if theboundaries are too close for comfort in 3D!There are several other approaches under development that promise toimprove this situation greatly that we will not have room to explore in detail here,but should be mention<strong>ed</strong>. Generally, one has in mind using Cauchy evolution inthe strong field, interior region where, say, black holes are colliding. The outerpart of this region will be match<strong>ed</strong> to some exterior treatment design<strong>ed</strong> to handlewhat is primarily expect<strong>ed</strong> to be outgoing radiation.Two major approaches have been develop<strong>ed</strong> by the NSF Black Hole GrandChallenge Alliance, a large USA collaboration working to solve the black holecoalescence problem, and other groups. First, by using perturbation theory, it ispossible to identify quantities in the numerically evolv<strong>ed</strong> metric functions thatobey the Regge–Wheeler and Zerilli wave equations that describe gravitational<strong>waves</strong> propagating on a black hole background. These can be us<strong>ed</strong> to provideboundary conditions on the metric and extrinsic curvature functions in an actualevolution, as describ<strong>ed</strong> in a recent paper [100]. This is an excellent stepforward in outer boundary treatments that should work to minimize reflectionsof the outgoing wave signals from the outer boundary. In tests with weak<strong>waves</strong>, a full 3D Cauchy evolution code has been successfully match<strong>ed</strong> to theperturbative treatment at the boundary, permitting <strong>waves</strong> to escape from theinterior region with very little reflection. Alternatively, ‘Cauchy-characteristicmatching’ attempts to match spacelike slices in the Cauchy region to null slices


Still newer formulations: towards a stable evolution system 379at some finite radius, and the null slices can be carri<strong>ed</strong> out to null infinity. 3Dcharacteristic evolution codes have progress<strong>ed</strong> dramatically in recent years, andalthough the full 3D matching remains to be complet<strong>ed</strong>, tests of the schemein specializ<strong>ed</strong> settings show promise [11]. One can also use the hyperbolicformulations of the Einstein equations to find eigenfields, for which outgoingconditions can in principle be appli<strong>ed</strong> [32] in 1D. In 3D this technique is stillunder development, but it shows promise for future work. Less sophisticat<strong>ed</strong>approaches that seem nonetheless rather successful are discuss<strong>ed</strong> in [63]. Finally,there is another hyperbolic approach which uses conformal rescaling to movethe boundary to infinity [12–15]. These methods have different strengths andweaknesses, but all promise to improve boundary treatments significantly, helpingto enable longer evolutions than are presently possible.18.3.3 Special difficulties with black holesThe techniques describ<strong>ed</strong> so far are generic in their application in numericalrelativity. However, in this section we describe a few problems that arecharacteristic of black holes, and special algorithms under development to handlethem. Black hole spacetimes all have in common one problem: singularities lurkwithin them, which must be handl<strong>ed</strong> numerically. Developing suitable techniquesfor doing so is one of the major research priorities of the community at present.If one attempts to evolve directly into the singularity, infinite curvature will beencounter<strong>ed</strong>, causing any numerical code to break down.Traditionally, the singularity region is avoid<strong>ed</strong> by the use of ‘singularityavoiding’ time slices, that wrap up around the singularity. Consider the evolutionshown in figure 18.1. A star is collapsing, a singularity is forming, and time slicesare shown which avoid the interior while still covering a large fraction of thespacetime where <strong>waves</strong> will be seen by a distant observer. However, these slicingconditions by themselves do not solve the problem; they merely serve to delay theonset of instabilities. As shown in figure 18.1, in the vicinity of the singularitythese slicings inevitably contain a region of abrupt change near the horizon, and aregion in which the constant time slices dip back deep into the past in some sense.This behaviour typically manifests itself in the form of sharply peak<strong>ed</strong> profilesin the spatial metric functions [80], ‘grid stretching’ [101] or large coordinateshift [73] on the BH throat, etc. Numerical simulations will eventually crash dueto these pathological properties of the slicing.18.3.3.1 Apparent horizon boundary conditions (AHBC)Cosmic censorship suggests that in physical situations, singularities are hiddeninside BH horizons. Because the region of spacetime inside the horizon is causallydisconnect<strong>ed</strong> from the region of interest outside the horizon, one is tempt<strong>ed</strong>numerically to cut away the interior region containing the singularity, and evolveonly the singularity-free region outside, as originally suggest<strong>ed</strong> by Unruh [102].


380 Numerical relativityFigure 18.1. A spacetime diagram showing the formation of a BH, and time slicestraditionally us<strong>ed</strong> to foliate the spacetime in traditional numerical relativity with singularityavoiding time slices. As the evolution proce<strong>ed</strong>s, pathologically warp<strong>ed</strong> hypersurfacesdevelop, leading to unresolvable gradients that cause numerical codes to crash.This has the consequence that there will be a region inside the horizon that simplyhas no numerical data. To an outside observer no information will be lost sincethe regions cut away are unobservable. Because the time slices will not ne<strong>ed</strong>such sharp bends to the past, this proc<strong>ed</strong>ure will drastically r<strong>ed</strong>uce the dynamicrange, making it easier to maintain accuracy and stability. Since the singularityis remov<strong>ed</strong> from the numerical spacetime, there is in principle no physical reasonwhy BH codes cannot be made to run indefinitely without crashing.We spoke innocently about the BH horizon, but did not distinguish betweenthe apparent and event horizon. These are very different concepts! Whilethe event horizon, which is roughly a null surface that never reaches infinityand never hits the singularity, may hide singularities from the outside world inmany situations, there is no guarantee that the apparent horizon, which is the(outermost) surface that has instantaneously zero expansion everywhere, evenexists on a given slice! (By ‘zero expansion’ we mean that the surface area ofoutgoing bundles of photons normal to the surface is constant. Hence, the surfaceis ‘trapp<strong>ed</strong>’.) Methods for finding event horizons in numerical spacetimes are nowknown, and will be discuss<strong>ed</strong> below. However, event horizons can only be foundafter examining the history of an evolution that has been already been carri<strong>ed</strong>


Still newer formulations: towards a stable evolution system 381out to sufficiently late times [103, 104]. Hence they are useless in providingboundaries as one integrates forward in time. On the other hand the apparenthorizon, if it exists, can be found on any given slice by searching for clos<strong>ed</strong>two-surfaces with zero expansion. Although one should worry that in a genericBH collision, one may evolve into situations where no apparent horizon actuallyexists, let us cross that bridge if we come to it! Methods for finding apparenthorizons will also be discuss<strong>ed</strong> below, but for now we assume that such a methodexists.Given these considerations, there are two basic ideas behind theimplementation of the apparent horizon boundary condition (AHBC), also knownas black hole excision:(a) It is important to use a finite-differencing scheme which respects thecausal structure of the spacetime. Since the horizon is a one-way membrane,quantities on the horizon can be affect<strong>ed</strong> only by quantities outside but not insidethe horizon: all quantities on the horizon can in principle be updat<strong>ed</strong> solely interms of known quantities residing on or outside the horizon. There are varioustechnical details and variations on this idea, which is call<strong>ed</strong> ‘causal differencing’[105] or ‘causal reconnection’ [106], but here we focus primarily on the basicideas and results obtain<strong>ed</strong> to date.(b) A shift is us<strong>ed</strong> to control the motion of the horizon, and the behaviour ofthe grid points outside the BH, as they tend to fall into the horizon if uncontroll<strong>ed</strong>.An additional advantage to using causal differencing is that it allows oneto follow the information flow to create grid points with proper data on them,as ne<strong>ed</strong><strong>ed</strong> inside the horizon, even if they did not exist previously. (Rememberabove that we have cut away a region inside the horizon, so in fact we have no datathere.) One example is to let a BH move across the computational grid. If a BHis moving physically, it may also be desirable to have it move through coordinatespace. Otherwise, all physical movement will be represent<strong>ed</strong> by the ‘motion’ ofthe grid points. For a single BH moving in a straight line, this may be possible(though complicat<strong>ed</strong>), but for spiralling coalescence this will lead to hopelesslycontort<strong>ed</strong> grids. The imm<strong>ed</strong>iate consequence of this is that as a BH moves acrossthe grid, regions in the wake of the hole, now in its exterior, must have previouslybeen inside it where no data exist! However, with AHBC and causal differencingthis ne<strong>ed</strong> not be a problem.Does the AHBC idea work? Preliminary indications are very promising. Inspherical symmetry (1D), numerous studies show that one can locate horizons, cutaway the interior, and evolve for essentially unlimit<strong>ed</strong> times (t ∝ 10 3−4 M, whereM is the black hole mass). The growth of metric functions can be completelycontroll<strong>ed</strong>, errors are r<strong>ed</strong>uc<strong>ed</strong> to a very low level, and the results can be obtain<strong>ed</strong>with a large variety of shift and slicing conditions, and with matter falling in theBH to allow for true dynamics even in spherical symmetry [105, 107–109].In 3D, the basic ideas are similar but the implementation is much mor<strong>ed</strong>ifficult. The first successful test of these ideas to a Schwarzschild BH in 3D us<strong>ed</strong>horizon excision and a shift provid<strong>ed</strong> from similar simulations carri<strong>ed</strong> out with a


382 Numerical relativity1D code [45]. The errors were found to be greatly r<strong>ed</strong>uc<strong>ed</strong> when compar<strong>ed</strong> even tothe 1D evolution with singularity avoiding slicings. Another 3D implementationof the basic technique was provid<strong>ed</strong> by Brügmann [60].This was a proof of principle, but more general treatments are following.Daues extend<strong>ed</strong> this work to a full range of shift conditions [6], including the full3D minimal distortion shift [10]. He also appli<strong>ed</strong> these techniques to dynamicBHs, and collapse of a star to form a BH, at which point the horizon is detect<strong>ed</strong>,the region interior to the horizon excis<strong>ed</strong>, and the evolution continu<strong>ed</strong> withAHBC. The focus of this work has been on developing general gauge conditionsfor single BHs without movement through a grid. Under these conditions, BHshave been accurately evolv<strong>ed</strong> well beyond t = 100M. The NSF Black HoleGrand Challenge Alliance has been focus<strong>ed</strong> on the development of 3D AHBCtechniques for moving Schwarzschild BHs [7]. In this work, analytic gaugeconditions are provid<strong>ed</strong>, which are chosen to make the evolution static, althoughthe numerical evolution is allow<strong>ed</strong> to proce<strong>ed</strong> freely. This moving hole is thefirst successful 3D test of populating grid points with data as they emerge in theBH wake. The recent, as yet unpublish<strong>ed</strong>, work of Huq et al has successfullyevolv<strong>ed</strong> full 3D grazing collisions through about t = 10M ADM , including thetopology change from two excision regions to a single one, while Alcubierre et alhave recently evolv<strong>ed</strong> Schwarzschild black holes in 3D, with rather general gaugeconditions, for well over t = 4000M ADM with less than a few per cent error!These new results are significant achievements, and show that the basictechniques outlin<strong>ed</strong> above are not only sound, but are also practically realizablein a 3D numerical code. However, there is still a significant amount of work to b<strong>ed</strong>one! The techniques have yet to be appli<strong>ed</strong> carefully to distort<strong>ed</strong> BHs, with testsof the waveforms emitt<strong>ed</strong>. There are still clearly many steps to be taken beforethe techniques will be successfully appli<strong>ed</strong> to the general BH merger problem.18.4 Tools for analysing the numerical spacetimesWe now turn to the description of several important tools that have been develop<strong>ed</strong>to analyse the results of a numerical evolution, carri<strong>ed</strong> out by some numericalevolution scheme. The evolution will generally provide metric functions ona grid, but as describ<strong>ed</strong> above these functions are highly dependent on boththe coordinate system and gauge in which the system is evolv<strong>ed</strong>. Determiningphysical information, such as whether a black hole exists in the data, or whatgravitational waveforms have been emitt<strong>ed</strong>, are the subjects of this section.18.4.1 Horizon findersAs describ<strong>ed</strong> earlier, black holes are defin<strong>ed</strong> by the existence of an event horizon(EH), the surface of no return from which nothing, not even light, can escape.The event horizon is the boundary that separates those null geodesics that reachinfinity from those that do not. The global character of such a definition implies


Tools for analysing the numerical spacetimes 383that the position of an EH can only be found if the whole history of the spacetimeis known. For numerical simulations of black hole spacetimes in particular, thisimplies that in order to locate an EH one ne<strong>ed</strong>s to evolve sufficiently far into thefuture, up to a time where the spacetime has basically settl<strong>ed</strong> down to a stationarysolution. Recently, methods have been develop<strong>ed</strong> to locate and analyse blackhole horizons in numerically generat<strong>ed</strong> spacetimes, with a number of interestingresults obtain<strong>ed</strong> [103, 104, 110–113].In contrast, an apparent horizon (AH) is defin<strong>ed</strong> locally in time as the outermostmarginally trapp<strong>ed</strong> surface [114], i.e. a surface on which the expansion ofout-going null geodesics is everywhere zero. An AH can therefore be defin<strong>ed</strong> ona given spatial hypersurface. A well-known result [114] guarantees that if an AHis found, an EH must exist somewhere outside of it and hence a black hole hasform<strong>ed</strong>.18.4.2 Locating the apparent horizonsThe expansion of a congruence of null rays moving in the outward normaldirection to a clos<strong>ed</strong> surface can be shown to be [20] =∇ i s i + K ij s i s j − tr K, (18.17)where ∇ i is the covariant derivative associat<strong>ed</strong> with the 3-metric γ ij , s i is thenormal vector to the surface, K ij is the extrinsic curvature of the time slice, andtr K is its trace. An AH is then the outermost surface such that = 0. (18.18)This equation is not affect<strong>ed</strong> by the presence of matter, since it is purely geometricin nature. We can use the same horizon finders without modification for vacuumas well as non-vacuum spacetimes. The key is to find a clos<strong>ed</strong> surface with normalvector s i satisfying this equation.18.4.2.1 Minimization algorithmsAs apparent horizons are defin<strong>ed</strong> by the vanishing of the expansion on a surface,a fairly obvious algorithm to find such a surface involves minimizing a suitablenorm of the expansion below some tolerance while adjusting test surfaces.Minimization algorithms for finding apparent horizons were among the firstmethods develop<strong>ed</strong> [115, 116]. More recently, a 3D minimization algorithmwas develop<strong>ed</strong> and implement<strong>ed</strong> by the Potsdam/NCSA/Wash U group, appli<strong>ed</strong>to a variety of black hole initial data and 3D numerically evolv<strong>ed</strong> black holespacetimes [117–121]. Essentially the same algorithm was also implement<strong>ed</strong>independently by Baumgarte et al [122].The basic idea behind a minimization algorithm is to assume the surface canbe represent<strong>ed</strong> by a function F(x i ) = 0, expand it in terms of some set of basis


384 Numerical relativityfunctions, and then minimize the integral of the square of the expansion 2 overthe surface. For example, one can parameterize a surface asF(r,θ,φ)= r − h(θ, φ). (18.19)The surface under consideration will be taken to correspond to the zero level ofF. The function h(θ, φ) is then expand<strong>ed</strong> in terms of spherical harmonics:h(θ, φ) =l max ∑l∑l=0 m=−la lm Y lm (θ, φ). (18.20)Similar techniques were develop<strong>ed</strong> by [123].At an AH the expansion integral over the surface should vanish, and we willhave a global minimum. Of course, since numerically we will never find a surfacefor which the integral vanishes exactly, one must set a given tolerance level belowwhich a horizon is assum<strong>ed</strong> to have been found.Minimization algorithms for finding AHs have a few drawbacks: First, thealgorithm can easily settle down on a local minimum for which the expansion isnot zero, so a good initial guess is often requir<strong>ed</strong>. Moreover, when more thanone marginally trapp<strong>ed</strong> surface is present, as is often the case, it is very difficultto pr<strong>ed</strong>ict which of these surfaces will be found by the algorithm. The algorithmcan often settle on an inner horizon instead of the true AH. Again, a good initialguess can help point the finder towards the correct horizon. Finally, minimizationalgorithms tend to be very slow when compar<strong>ed</strong> with ‘flow’ algorithms of the typ<strong>ed</strong>escrib<strong>ed</strong> in the next section. Typically, if N is the total number of terms in thespectral decomposition, a minimization algorithm requires of the order of a fewtimes N 2 evaluations of the surface integrals (where in our experience ‘afew’ cansometimes be as high as ten).This algorithm has been implement<strong>ed</strong> in the ‘Cactus’ code for 3D numericalrelativity [34]. For more details of the application of this algorithm, see[117–119, 122].18.4.2.2 3D fast flow algorithmA second method that has been implement<strong>ed</strong> in the ‘Cactus’ code is the ‘fast flow’method propos<strong>ed</strong> by Gundlach [124]. Starting from an initial guess for the a lm ,itapproaches the apparent horizon through the iterationa (n+1)lm= a (n)lm − A1 + Bl(l + 1) (ρ)(n) lm(18.21)where (n) labels the iteration step, ρ is some positive definite function (‘aweight’), and (ρ) lm are the harmonic components of the function (ρ). Variouschoices for the weight ρ and the coefficients A and B parametrize a family of suchmethods. The fast flow algorithm in Cactus usesρ = 2r 2 |∇ F|[(g ij − s i s j )(ḡ ij −∇ i r∇ j r)] −1 , (18.22)


Tools for analysing the numerical spacetimes 385where ḡ ij is the flat background metric associat<strong>ed</strong> with the coordinates (r,θ,φ),andαA =l max (l max + 1) + β, B = β (18.23)αwith α = c and β = c/2. Here c is a variable step size, with a typical valueof c ∼ 1. l max is the maximum value of l one chooses to use in describingthe surface. The iteration proc<strong>ed</strong>ure is a finite-difference approximation to aparabolic flow, and the adaptive step size is chosen to keep the finite-differenceapproximation roughly close to the flow limit to prevent overshooting of the trueapparent horizon. The adaptive step size is determin<strong>ed</strong> by a standard method us<strong>ed</strong>in ODE integrators: we take one full step and two half steps and compare theresulting a lm . If the two results differ too much one from another, the step size isr<strong>ed</strong>uc<strong>ed</strong>.Other methods for finding apparent horizons, bas<strong>ed</strong> directly on computingthe Jacobian of the finite differenc<strong>ed</strong> horizon equation, have been develop<strong>ed</strong>[125, 126] and successfully us<strong>ed</strong> in 3D. For details, see these references.18.4.3 Locating the event horizonsThe AH is defin<strong>ed</strong> locally in time and hence is much easier to locate than theevent horizon (EH) in numerical relativity. The EH is a global object in time;it is trac<strong>ed</strong> out by the path of outgoing light rays that never propagate to futurenull infinity, and never hit the singularity. (It is the boundary of the causal pastof future null infinity Â˙− (Á + ).) In principle, one ne<strong>ed</strong>s to know the entire timeevolution of a spacetime in order to know the precise location of the EH. However,in spite of the global properties of the EH, hope is not lost for finding it veryaccurately, even in a numerical simulation of finite duration. Here we discussa method to find the EH, given a numerically construct<strong>ed</strong> black hole spacetimethat eventually settles down to an approximately stationary state at late times. Inprinciple, as the EH is a null surface, it can be found by tracing the path of nullrays through time. Outward going light rays emitt<strong>ed</strong> just outside the EH willdiverge away from it, escaping to infinity, and those emitt<strong>ed</strong> just inside the EHwill fall away from it, towards the singularity. In a numerical integration it isdifficult to follow accurately the evolution of a horizon generator forward in time,as small numerical errors cause the ray to drift and diverge rapidly from the trueEH. It is a physically unstable process. However, we can actually use this propertyto our advantage by considering the time-revers<strong>ed</strong> problem. In a global sense intime, any outward going photon that begins near the EH will be attract<strong>ed</strong> to thehorizon if integrat<strong>ed</strong> backward in time [103, 117]. In integrating backwards intime, it turns out that it suffices to start the photons within a fairly broad regionwhere the EH is expect<strong>ed</strong> to reside. Such a horizon-containing region, as we callit, is often easy to determine after the spacetime has settl<strong>ed</strong> down to approximatestationarity. The crucial point is that when integrat<strong>ed</strong> backward in time along nullgeodesics, this horizon-containing region shrinks rapidly in ‘thickness’, leading


386 Numerical relativityto a very accurate determination of the location of the EH at earlier times. Notethat it is the earlier time when the black hole is under highly dynamical evolutionthat we are really interest<strong>ed</strong> in.Although one can integrate individual null geodesics backwards in time,we find that there are various advantages to integrating the entire null surfacebackwards in time. A null surface, if defin<strong>ed</strong> by f (t, x i ) = 0 satisfies theconditiong µν ∂ µ f ∂ ν f = 0. (18.24)Hence the evolution of the surface can be obtain<strong>ed</strong> by a simple integration,√∂ t f = −gti ∂ i f + (g ti ∂ i f ) 2 − g tt g ij ∂ i f ∂ j fg tt . (18.25)The inner and outer boundary of the horizon containing region when integrat<strong>ed</strong>backwards in time, will rapidly converge to practically a single surface to withinthe resolution of the numerically construct<strong>ed</strong> spacetime, i.e. a small fraction ofa grid point. An accurate location of the event horizon is hence obtain<strong>ed</strong>. Wehenceforth shall represent the horizon surface as the function f H (t, x i ). Asidefrom the simplicity of this method, there are a number of technical advantagesas discuss<strong>ed</strong> in [103]. One particularly noteworthy point is that this method iscapable of giving the caustic structure of the event horizon if there is any; fordetails see [103].The function f H (t, x i ) provides the complete coordinate location of theEH through the spacetime (or a very good approximation of it, as shown in[104]). This function by itself directly gives us the topology and location of theEH. When combin<strong>ed</strong> with the induc<strong>ed</strong> metric function on the surface, which isrecord<strong>ed</strong> throughout the evolution, it gives the intrinsic geometry of the EH. Whenfurther combin<strong>ed</strong> with the spacetime metric, all properties of the EH including itsemb<strong>ed</strong>ding can be obtain<strong>ed</strong>. Moreover, as the normal of f H (t, x i ) = 0givesthe null generators of the horizon, it is an easy further step to determine the nullgenerators, and hence the complete dynamics of the horizon in this formulation.As describ<strong>ed</strong> in a series of papers, the event horizon, once found with such amethod, can be analys<strong>ed</strong> to provide important information about the dynamics ofblack holes in a numerically generat<strong>ed</strong> spacetime [103, 104, 110–113].18.4.4 Wave extractionThe gravitational radiation emitt<strong>ed</strong> is one of the most important quantities ofinterest in many astrophysical processes. The radiation is generat<strong>ed</strong> in regionsof strong and dynamic gravitational fields, and then propagat<strong>ed</strong> to regions faraway where it will som<strong>ed</strong>ay be detect<strong>ed</strong>. We take the approach of computing thegeneration and evolution of the fields in a fully nonlinear way, while analysing theradiation with a perturbation formulation in the regions where it can be so treat<strong>ed</strong>.


Tools for analysing the numerical spacetimes 387The theory of black hole perturbations is well develop<strong>ed</strong>. One identifiescertain perturb<strong>ed</strong> metric quantities that evolve according to wave equationson the black hole background. These perturb<strong>ed</strong> metric functions are alsodependent of the gauge in which they are comput<strong>ed</strong>. We use a gauge-invariantprescription for isolating wave modes on black hole background, develop<strong>ed</strong> firstby Moncrief [127]. The basic idea is that although the perturb<strong>ed</strong> metric functionstransform under coordinate transformations (gauge transformations), one canidentify certain linear combinations of these functions that are invariant to firstorder of the perturbation. These gauge-invariant functions are the quantities thatcarry true physics, which does not depend on coordinate systems. They obey thewave equations describing <strong>waves</strong> propagating on the fix<strong>ed</strong> black hole background.There are two independent wave modes, even- and odd-parity, corresponding tothe two degrees of fre<strong>ed</strong>om, or polarization modes, of the <strong>waves</strong>.A waveform extraction proc<strong>ed</strong>ure has been develop<strong>ed</strong> that allows oneto process the metric and to identify the wave modes. The gravitationalwavefunction (often call<strong>ed</strong> the ‘Zerilli function’ for even-parity or the ‘Regge–Wheeler function’ for odd-parity) can be comput<strong>ed</strong> by writing the metric as thesum of a background black hole part and a perturbation:g αβ = o g αβ +h αβ (Y l,m ), (18.26)where the perturbation h αβ is expand<strong>ed</strong> in spherical harmonics and theirotensor generalizations and the background part g αβ is spherically symmetric.To compute the elements of h αβ in a numerical simulation, one integratesthe numerically evolv<strong>ed</strong> metric components g αβ against appropriate sphericalharmonics over a coordinate 2-sphere surrounding the black hole, makinguse of the orthogonality properties of the tensor harmonics. This process isperform<strong>ed</strong> for each l, m mode for which waveforms are desir<strong>ed</strong>. The resultingfunctions h αβ (Y l,m ) can then be combin<strong>ed</strong> in a gauge-invariant way, followingthe prescription given by Moncrief [127]. For each l, m mode, this gaugeinvariantgravitational waveform can be extract<strong>ed</strong> when the wave passes through‘detectors’ at some fix<strong>ed</strong> radius in the computational grid. This proc<strong>ed</strong>ure hasbeen describ<strong>ed</strong> in detail in [128–130], and more generally in [121, 131, 132]. Itworks amazingly well, allowing extraction of <strong>waves</strong> that carry very small energies(of order 10 −7 M or less, with M being the mass of the source) away from thesource. The proc<strong>ed</strong>ure should apply to any isolat<strong>ed</strong> source of <strong>waves</strong>, such ascolliding black holes, neutron stars, etc. The spherical perturbation theory (witha few minor modifications) has also been appli<strong>ed</strong> to distort<strong>ed</strong> rotating black holeswith satisfactory results [128–130].


388 Numerical relativity18.5 Computational science, numerical relativity, and the‘Cactus’ code18.5.1 The computational challenges of numerical relativityBefore we describe our computational methods in the following subsections, wesummarize the computational challenges of numerical relativity discuss<strong>ed</strong> above.It is in response to these challenges that we have devis<strong>ed</strong> the computationalmethods.• Computational challenges due to the complexity of the physics involv<strong>ed</strong>.The Einstein equations are probably the most complex partial differentialequations in all physics, forming a system of dozens of coupl<strong>ed</strong>, nonlinearequations, with thousands of terms, of mix<strong>ed</strong> hyperbolic, elliptic, and evenundefin<strong>ed</strong> types in a general coordinate system. The evolution has ellipticconstraints that should be satisfi<strong>ed</strong> at all times. In simulations withoutsymmetry, as would be the case for realistic astrophysical processes, codescan involve hundr<strong>ed</strong>s of 3D arrays, and ten of thousands of operationsper grid point per update. Moreover, for simulations of astrophysicalprocesses, we will ultimately ne<strong>ed</strong> to integrate numerical relativity withtraditional tools of computational astrophysics, including hydrodynamics,nuclear astrophysics, radiation transport and magnetohydrodynamics, whichgovern the evolution of the source terms (i.e. the right-hand side) of theEinstein equations. This complexity requires us to push the frontiers ofmassively parallel computation.• Challenge in collaborative technology. The integration of numerical relativityinto computational astrophysics is a multidisciplinary development,partly due to the complexity of the Einstein equations, and partly due to thephysical systems of interest. Solving the Einstein equations on massivelyparallel computers involves gravitational physics, computational science, numericalalgorithm and appli<strong>ed</strong> mathematics. Furthermore, for the numericalsimulations of realistic astrophysical systems; many physics disciplines,including relativity, astrophysics, nuclear physics, and hydrodynamics areinvolv<strong>ed</strong>. It is therefore essential to have the numerical code software engineer<strong>ed</strong>to allow codevelopment by different research groups and groups withdifferent expertise.• The multiscale problem. Astrophysics of strongly gravitating systemsinherently involves many length and timescales. The microphysics of theshortest scale (the nuclear force), controls macroscopic dynamics on thestellar scale, such as the formation and collapse of neutron stars (NSs). Onthe other hand, the stellar scale is at least ten times less than the wavelengthof the gravitational <strong>waves</strong> emitt<strong>ed</strong>, and many orders of magnitude less thanthe astronomical scales of their accretion disk and jets; these larger scalesprovide the directly observ<strong>ed</strong> signals. Numerical studies of these systems,aiming at direct comparison with observations, fundamentally require the


Cactus computational toolkit 389capability of handling a wide range of dynamical time and length scales.• Challenge in interactive computational science. In spite of the incr<strong>ed</strong>ibleadvances in computational science, most simulations are still done in a veryold-fashion<strong>ed</strong> way. Jobs are submitt<strong>ed</strong> in batch mode, data are output,results are studi<strong>ed</strong>, and the process starts over again. This is a very timeconsuming and cumbersome process. In order to really take advantage oflarge-scale simulation as a tool for computational scientists, it is necessaryto develop new techniques that allow one to conveniently make use of theircomputational resources, wherever they may be, to interactively monitor thesimulations with advanc<strong>ed</strong> visualization tools, perhaps in conjunction withtheir colleagues in different parts of the world, and to interactively adjust thesimulation bas<strong>ed</strong> on the observ<strong>ed</strong> results.All of these issues lead to important research questions in computationalscience. Here we give an overview of some of our effort in these directions,focusing on performance and coding issues on parallel machines, and on th<strong>ed</strong>evelopment of a community code that incorporates all the mathematical andcomputational techniques describ<strong>ed</strong> above (and many more), in a collaborativeinfrastructure for numerical relativity.18.6 Cactus computational toolkitThe computational and collaborative ne<strong>ed</strong>s of numerical relativity are clearlyimmense. To develop a basic 3D code with all the different modules, includingparallel layers, adaptive mesh refinement, elliptic solvers, initial value solvers,gauge conditions, black hole excision modules, analysis tools, wave extraction,hydrodynamics modules, visualization tools, etc, require dozens of person yearsof effort from many different disciplines (in fact, such a feat has still not beendone by the entire community!). Different groups often ne<strong>ed</strong>lessly repeat eachother’s effort, further slowing the progress of the field. The NSF Black HoleGrand Challenge was a first attempt to address this problem, and an outgrowthof that effort l<strong>ed</strong> to the development of the ‘Cactus’ Computational Toolkit(CCTK), develop<strong>ed</strong> by the Potsdam group, in collaboration first with NCSAand Washington University, and now with a growing number of internationalcollaborators in various disciplines. Originally design<strong>ed</strong> to solve Einstein’sequations, the CCTK has grown into a general purpose parallel environment forsolving complex PDEs [133–135] that is being pick<strong>ed</strong> up by various communitiesin computational science. Here we focus on its application to Einstein’s equations.Cactus is design<strong>ed</strong> to minimize barriers to the community developmentand use of the code, including the complexity associat<strong>ed</strong> with both the codeitself and the network<strong>ed</strong> supercomputer environments in which simulations anddata analysis are perform<strong>ed</strong>. This complexity is particularly noticeable in largemultidisciplinary simulations such as ours, because of the range of disciplines thatmust contribute to code development (relativity, hydrodynamics, astrophysics,


390 Numerical relativitynumerics and computer science) and because of the geographical distribution ofthe people and computer resources involv<strong>ed</strong> in simulation and data analysis.The name Cactus comes from the design of a central core (or flesh) whichconnects to application modules (or thorns) through an extensible interface.Thorns can implement custom develop<strong>ed</strong> scientific or engineering applications,such as the Einstein solvers, or other applications such as computational fluiddynamics. Other thorns from a standard computational toolkit provide a range ofcapabilities, such as parallel I/O, data distribution, or checkpointing.Cactus runs on many architectures. Applications, develop<strong>ed</strong> on standardworkstations or laptops, can be seamlessly run on clusters or supercomputers.Parallelism and portability are achiev<strong>ed</strong> by hiding the driver layer and featuressuch as the I/O system and calling interface under a simple abstraction API. TheCactus API supports C/C++ and F77/F90 programming languages for the thorns.Thus, thorn programmers can work in the language they find most convenient,and are not requir<strong>ed</strong> to master the latest and greatest computing paradigms. Thismakes it easier for scientists to turn existing codes into thorns which can thenmake use of the complete Cactus infrastructure, and in turn be us<strong>ed</strong> by otherthorns within Cactus.Cactus provides easy access to many cutting <strong>ed</strong>ge software technologies beingdevelop<strong>ed</strong> in the academic research community, such as the Globus MetacomputingToolkit, HDF5 parallel file I/O, the PETSc scientific computing library,adaptive mesh refinement, web interfaces, and advanc<strong>ed</strong> visualization tools.So, how does a user use the code? A detail<strong>ed</strong> user guide is available withthe code (see http://www.cactuscode.org), but in a nutshell, one specifies whichphysics modules, and which computational/parallelism modules, are desir<strong>ed</strong> in aconfiguration file, and makes the code on the desir<strong>ed</strong> architecture, which can beany one of a number of machines from SGI/Cray Origin or T3E, Dec Alpha, Linuxworkstations or clusters, NT clusters, and others. The make system automaticallydetects the architecture and configures the code appropriately. Control of runparameters is then provid<strong>ed</strong> through an input file. Additional modules can beselect<strong>ed</strong> from a large community-develop<strong>ed</strong> library, or new modules may bewritten and us<strong>ed</strong> in conjunction with the pre-develop<strong>ed</strong> modules.Our experiences with Cactus up to now suggest that these techniques areeffective. It allows a code of many tens of thousands of lines, but with acompact flesh that is possible to maintain despite the large number of peoplecontributing to it. The common code base has enhanc<strong>ed</strong> the collaborativeprocess, having important and beneficial effects on the flow of ideas betweenremote groups. This flexible, open code architecture allows, for example, arelativity expert to contribute to the code without knowing the details of, say,the computational layers (e.g., message passing or AMR libraries) or othercomponents (e.g., hydrodynamics). We encourage users from throughout therelativity and astrophysics communities to make use of this freely downloadablecode infrastructure and physics modules, either for their own use, or as acollaborative tool to work with other groups in the community.


Cactus computational toolkit 39118.6.1 Adaptive mesh refinement3D simulations of Einstein’s equations are very demanding computationally. Inthis section we outline the computational ne<strong>ed</strong>s, and techniques design<strong>ed</strong> tor<strong>ed</strong>uce them. We will ne<strong>ed</strong> to resolve <strong>waves</strong> with wavelengths of order 5M or less,where M is the mass of the BH or the neutron star. Although for Schwarzschildblack holes, the fundamental l = 2 quasinormal mode wavelength is 16.8M,higher modes, such as l = 4 and above, have wavelengths of 8M and below. TheBH itself has a radius of 2M. More importantly, for very rapidly rotating KerrBHs, which are expect<strong>ed</strong> to be form<strong>ed</strong> in realistic astrophysical BH coalescence,the modes are shift<strong>ed</strong> down to significantly shorter wavelengths [3, 4]. As we


392 Numerical relativityinitial distribution may not even see the final BH. Further, as pulses of radiationpropagate back out from the origin, they too may have to be resolv<strong>ed</strong> in regionswhere there was previously a coarse grid. Choptuik’s AMR system, built on earlywork of Berger and Oliger [138], was able to track dynamically features thatdevelop, enabling him to discover and accurately measure BH critical phenomenathat have now become so widely studi<strong>ed</strong> [139].Bas<strong>ed</strong> on this success and others, and on the general considerations discuss<strong>ed</strong>above, full 3D AMR systems are under development to handle the much greaterne<strong>ed</strong>s of solving the full set of 3D Einstein equations. A large collaboration,begun by the NSF Black Hole Grand Challenge Alliance, has been developing asystem for distributing computing on large parallel machines, call<strong>ed</strong> Distribut<strong>ed</strong>Adapt<strong>ed</strong> Grid Hierarchies, or DAGH. DAGH was develop<strong>ed</strong> to provide MPIbas<strong>ed</strong>parallelism for the kinds of computations ne<strong>ed</strong><strong>ed</strong> for many PDE solvers,and it also provides a framework for parallel AMR. It is one of the majorcomputational science accomplishments to come out of the Alliance. Develop<strong>ed</strong>by Manish Parashar and Jim Browne, in collaboration with many subgroupswithin and without the Alliance, it is now being appli<strong>ed</strong> to many problemsin science and engineering. One can find information about DAGH online athttp://www.cs.utexas.<strong>ed</strong>u/users/dagh/.At least two other 3D software environments for AMR have been develop<strong>ed</strong>for relativity: one is call<strong>ed</strong> HLL, or Hierarchical Link<strong>ed</strong> Lists, develop<strong>ed</strong> byWild and Schutz [140]; another, call<strong>ed</strong> BAM, was the first AMR application in3D relativity develop<strong>ed</strong> by Brügmann [60]. The HLL system has recently beenappli<strong>ed</strong> to the test problem of the Zerilli equation (discuss<strong>ed</strong> above) describingperturbations of black holes [141]. This nearly 30 year old linear equation is stillproviding a powerful model for studying BH collisions, and it is also being us<strong>ed</strong>as a model problem for 3D AMR. In this work, the 1D Zerilli equation is recastas a 3D equation in Cartesian coordinates, and evolv<strong>ed</strong> within the AMR systemprovid<strong>ed</strong> by HLL. Even though the 3D Zerilli equation is a single linear equation,it is quite demanding in terms of resolution requirements, and without AMR it isextremely difficult to resolve both the initial pulse of radiation, the blue shiftingof <strong>waves</strong> as they approach the horizon, and the scattering of radiation, includingthe normal modes, far from the hole.18.7 Recent applications and progress18.7.1 Evolving pure gravitational <strong>waves</strong>With the new formulations of the Einstein equations discuss<strong>ed</strong> above, we are nowable to study the nonlinear dynamics of pure gravitational <strong>waves</strong> with much morestability than ever before. This allows us to use numerical relativity to probegeneral relativity in highly nonlinear regime. Can one form a black hole infull 3D from pure gravitational <strong>waves</strong>? Does one see critical phenomena in full3D? These inherently nonlinear phenomena have been investigat<strong>ed</strong> in 1D and 2D


Recent applications and progress 393studies, but little is known about generic 3D behaviour.In our investigations, we take as initial data a pure Brill-type gravitationalwave [142], later studi<strong>ed</strong> by Eppley [86, 116] and others [143]. The metric takesthe formds 2 = 4 [e 2q (dρ 2 + dz 2 ) + ρ 2 dφ 2 ] = 4 ˆds 2 , (18.27)where q is a free function subject to certain boundary conditions. Following[120, 132, 144], we choose q of the form]q = aρ 2 ρe[1 2−r2 + c(1 + ρ 2 ) cos2 (nφ) , (18.28)where a, c are constants, r 2 = ρ 2 + z 2 and n is an integer. For c = 0, thes<strong>ed</strong>ata sets r<strong>ed</strong>uce to the Holz [143] axisymmetric form, recently studi<strong>ed</strong> in full 3DCartesian coordinates [145]. Taking this form for q, we impose the condition oftime-symmetry, and solve the Hamiltonian constraint numerically in Cartesiancoordinates. An initial data set is thus characteriz<strong>ed</strong> only by the parameters(a, c, n). For the case (a, 0, 0), we found in [145] that no AH exists in initialdata for a < 11.8, and we also studi<strong>ed</strong> the appearance of an AH for other valuesof c and n.We have survey<strong>ed</strong> a large range of this parameter space, but here we discusstwo cases of interest: (a) a subcritical (but highly nonlinear) case where after aviolent collapse of the self-gravitating <strong>waves</strong>, there is a subsequent rebound andafter a few oscillations the <strong>waves</strong> all disperse; and (b) a supercritical case wherethe <strong>waves</strong> collapse in on themselves and imm<strong>ed</strong>iately form a black hole.The subcritical case studi<strong>ed</strong> in [50] has parameters (a = 4, c = 0, n = 0)in the notation above. It is a rather strong axisymmetric Brill wave (BW). Theevolution of this data set shows that part of the wave propagates outward whilepart implodes, re-expanding after passing through the origin. However, due tothe nonlinear self-gravity, not all of it imm<strong>ed</strong>iately disperses out to infinity; againpart re-collapses and bounces again. After a few collapses and bounces the wavecompletely disperses out to infinity. This behaviour is shown in figure 18.2(a),where the evolution of the central value of the lapse is given for simulationswith three different grid sizes: x = y = z = 0.16 (low resolution), 0.08(m<strong>ed</strong>ium resolution) and 0.04 (high resolution), using 32 3 ,64 3 and 128 3 gridpoints respectively. At late times, the lapse returns to 1 (the log returns to 0).Figure 18.2(b) shows the evolution of the log of the central value of the Riemanninvariant J for the same resolutions. At late times J settles on a constant valuethat converges rapidly to zero as we refine the grid. With these results, and directverification that the metric functions become stationary at late times, we concludethat spacetime returns to flat (in non-trivial spatial coordinates; the metric isdecid<strong>ed</strong>ly non-flat in appearance!).The same simulation carri<strong>ed</strong> out with the standard ADM systems crashesfar earlier than in the present case with the BSSN systems, which essentially run


394 Numerical relativityFigure 18.2. (a) Evolution of the log of the lapse α at r = 0 for the axisymmetricdata (4, 0, 0). The dash<strong>ed</strong>/dott<strong>ed</strong>/full curves represent simulations at low/m<strong>ed</strong>ium/highresolution. (b) Evolution of the Riemann invariant J at r = 0. The wave disperses afterdynamic evolution, leaving flat space behind.forever. With this experience, we next try the case of an even stronger amplitudewave, which in this case will actually collapse on itself and form a black hole.In figure 18.3, we study the development of the data set (a = 6, c = 0.2,n = 1), a full 3D data set, and watch it collapse to form a black hole (the firstsuch 3D simulation). The figure also compares this black hole formation to resultsobtain<strong>ed</strong> with an axisymmetric data set. The system clearly collapses on itself andrapidly forms a black hole. The waveform extraction shows that the newly form<strong>ed</strong>hole then rings at its quasinormal mode frequency. High quality images andmovies of these simulations can be found at http://jean-luc.aei-potsdam.mpg.de.These results are exciting examples of how numerical relativity can actas a laboratory to probe the nonlinear aspects of Einstein’s equations. Puregravitational <strong>waves</strong> are clearly a rich and exciting research area that allows oneto study Einstein’s equations as a nonlinear theory of physics. With these newcapabilities of accurate 3D evolution that can follow the implosion of <strong>waves</strong>to a black hole, there is much more physics to study, including the structureof horizons, full 3D studies of critical phenomena, and much more. Further,this study of pure vacuum <strong>waves</strong> has help<strong>ed</strong> us to understand the importance ofdeveloping and testing new formulations of Einstein’s equations for numericalpurposes. Without the new formulations, these results simply could not beobtain<strong>ed</strong>. Further, we have run literally hundr<strong>ed</strong>s of simulations like these in orderto determine which variation on the ‘BSSN’ families of formulations performbest. With this new knowl<strong>ed</strong>ge, we turn to the problem of 3D black holes.18.7.2 Black holesHaving test<strong>ed</strong> these new formulations of Einstein’s equations on the problem ofpure gravitational <strong>waves</strong>, we now apply what we have learn<strong>ed</strong> to the considerablymore complex problem of black hole evolutions. We first appli<strong>ed</strong> these newformulations to black hole spacetimes that have been very carefully test<strong>ed</strong> in


Recent applications and progress 395Figure 18.3. (a) The full (dott<strong>ed</strong>) curve is the AH for the full 3D data set (6, 0.2, 1)((6, 0, 0)) att = 9inthex–z plane. (b) The {l = 2, m = 0} waveform for the 3D(6, 0.2, 1) case at r = 4 (full curve) is compar<strong>ed</strong> to the axisymmetric (6, 0, 0) case (dott<strong>ed</strong>curve). The chain curve shows the fit of the 3D case to the two lowest lying QNMs fora BH of mass 0.99. (c) Same comparison for the {l = 4, m = 0} waveform. (d) Samecomparison for the non-axisymmetric {l = 2, m = 2} waveform.axisymmetry and with 3D nonlinear numerical codes, but with standard 3 + 1formulations, as well as with perturbative methods. In summary, these newformulations are able to extend the evolutions by a considerable amount, sinceinstabilities that develop during grid stretching caus<strong>ed</strong> by singularity avoidingslicings are highly suppress<strong>ed</strong>. 3D black hole evolutions that crash<strong>ed</strong> previouslyby t = 20–30M now routinely extend several times longer. However, by suchlate times the pathological peaks in metric functions associat<strong>ed</strong> with singularityavoiding slicings cannot be resolv<strong>ed</strong>, and then of course evolutions become veryinaccurate. This is actually a major improvement! Previously, with traditionalformulations, instabilities would develop as large gradients were creat<strong>ed</strong> near theblack hole, causing the codes to crash prematurely. As discuss<strong>ed</strong> above and shownin many papers [121,131,132,146], the evolutions can actually be very accurate—allowing the extraction of very delicate waveforms from a large spectrum ofmodes—and remain so until the code crashes. With the new formulations, wefind that we can break far through the former crash barrier, but as features becomeunder-resolv<strong>ed</strong> the results naturally become less accurate. For a fuller discussion


396 Numerical relativityof these results, and mathematical analysis giving insight into the improv<strong>ed</strong>behaviour of the new formulations, see [52].As an example of what we are now able to do with these new formulations,we now turn to an advanc<strong>ed</strong> application of a fully 3D ‘grazing collision’ of twoblack holes.18.7.2.1 True 3D grazing black hole collisionIn the previous sections we have shown that using the standard ADM formulationsof Einstein’s equations in 3D Cartesian coordinates, it is possible to perform veryaccurate evolutions of gravitational-wave and black hole spacetimes. However,with these formulations, there are instabilities that appear when large gradientsdevelop, leading to premature crashing of the code. On the other hand, the new‘BSSN’ family of formulations are much more robust. Having allow<strong>ed</strong> us to breakthrough the barriers seen in evolving pure nonlinear gravitational <strong>waves</strong>, we nowapply these formulations to full 3D grazing collisions of black holes of the typeoriginally consider<strong>ed</strong> by Brügmann a few years ago.The initial data sets we use here for binary BH systems were develop<strong>ed</strong>originally by Brandt and Brügmann [25]. They are very convenient, since noisometry is ne<strong>ed</strong><strong>ed</strong> and hence the elliptic solver can be appli<strong>ed</strong> on a standardCartesian grid without the ne<strong>ed</strong> to apply boundary conditions on strangely shap<strong>ed</strong>(e.g., non-planar!) surfaces.A few of these data sets were first evolv<strong>ed</strong> by Brügmann [147] using thestandard ADM formulation. This was a first pioneering attempt to go beyondthe highly symmetric black hole collisions that had been studi<strong>ed</strong> previously,combining for the first time unequal mass, spinning black holes with linear andorbital angular momentum. Brügmann was able to use nest<strong>ed</strong> grids to providereasonable resolution near the holes, while putting the boundary reasonably faraway. The result was that for select<strong>ed</strong> data sets he was able to carry out theevolution far enough to observe a merger of the two apparent horizons. However,the difficulties of the ADM formulation, discuss<strong>ed</strong> above, coupl<strong>ed</strong> with poorresolution achievable at that time limit<strong>ed</strong> these evolutions to about t = 7M, and itwas not possible to extract detail<strong>ed</strong> physics, such as horizon masses, waveforms,energies, spins, etc.Now we apply all we have learn<strong>ed</strong> in the last few years, together with theadvanc<strong>ed</strong> computational infrastructure develop<strong>ed</strong> in Cactus and the accessibilityof much larger computers, and revisit this same problem. This work is still inprogress, and calculations are presently underway to refine the waveforms andthe energy accounting, but we can report the following preliminary results.For initial data of the type describ<strong>ed</strong> in [147], we follow Brügmann andchoose individual mass parameters M 1 = 1.5, and M 2 = 1.0, and linear and spinmomenta on each hole such that the overall mass and angular momentum of theinitial slice are measur<strong>ed</strong> from asymptotic properties to beM ADM = 3.1 (18.29)


Recent applications and progress 397J = 6.7 so that a/m = J/MADM 2 = 0.70. (18.30)On a 256 processor Origin 2000 machine at NCSA we are able to runsimulations of 387 3 , which take roughly 100 Gb of memory. Still, with sufficientresolution to carry out long-term evolutions, the boundaries are still rather close,at roughly x = 12M. We use the ‘BSSN’ formulations to carry out the evolutions,coupl<strong>ed</strong> with vanishing shift and either maximal or algebraic slicings (of the‘1 + log’ family [45]) and with a three-step Crank–Nicholson method. Furtherdetails are in preparation for publication. Under these conditions, we find that weare able to evolve the black hole merger far beyond the time at which the horizonsmerge, beyond t = 30M, at which time the simulations become fairly inaccurate.(We must point out that we have to date only studi<strong>ed</strong> the apparent horizons. Theevent horizons can also be locat<strong>ed</strong> by techniques develop<strong>ed</strong> in [103]. At presentit is not known whether a single event horizon is present on the initial slice in thisdata set.) Depending on computational parameters, the simulations can be carri<strong>ed</strong>out far beyond this time without crashing, in stark contrast to earlier attemptswhich were doom<strong>ed</strong> to crash far earlier.Of course, the ‘time to crash’ is not a measure of success of a code! Whatwe are really interest<strong>ed</strong> in is whether we are able to extract meaningful physicsfrom such simulations. We are in the process of analysing such simulations ingreat detail, and the results are very encouraging. First, for the example discuss<strong>ed</strong>above, we begin with qualitative measurements of the physics we extract. Infigure 18.4 we show a sequence of visualizations of simulations near the time justbefore, during, and after the merger of the two holes. The coordinate locationsof the apparent horizons (AH) are shown as colour<strong>ed</strong> surfaces. The colourmaprepresents the local Gaussian curvature of the surface, comput<strong>ed</strong> from the induc<strong>ed</strong>2-metric on the horizon. As the holes approach each other and merge, a global AHdevelops. Meanwhile, a burst of gravitational <strong>waves</strong>, indicat<strong>ed</strong> by the colour<strong>ed</strong>wisps emanating from the BH system develops and propagates away. TheNewman–Penrose quantity 4 , comput<strong>ed</strong> fully nonlinearly, is us<strong>ed</strong> to indicate thegravitational <strong>waves</strong>. As this system has no symmetries, and includes rotation, alll–m-modes and both even- and odd-parity polarizations of the <strong>waves</strong> are present,leading to a much more complex structure in the wave patterns than one is us<strong>ed</strong> toseeing in such simulations. However, this is now moving much closer to what oneexpects to see in nature, and it, too, will be rather complicat<strong>ed</strong>! A full multipolaranalysis of the <strong>waves</strong> is in progress, and it is clear that quasinormal mode ringingof the final BH is present, as expect<strong>ed</strong>.These results are preliminary, but indicate that for the first time we are inde<strong>ed</strong>now able to simulate the late merger stages of two black holes colliding, withrather general spin, mass and momenta, and that we can now begin to studythe fine details of the physics. A quantitative analysis of the horizon evolution,mass of the final black hole, the energy emitt<strong>ed</strong>, the total angular momentum,etc, are underway, and preliminary results indicate that much detail<strong>ed</strong> physicscan be accurately extract<strong>ed</strong> from these simulations. Without more advanc<strong>ed</strong>


398 Numerical relativityFigure 18.4. We show a sequence of visualizations of the merger of two black holeswith unequal mass and spin. The apparent horizons are shown as the surfaces at thecentre of the image, and the colours represent the Gaussian curvature. The <strong>waves</strong>, shownemanating from the merger, are visualizations of the real part of the Newman–Penrosequantity 4 . The top left-hand panel shows the system just before the merger, while thebottom right-hand panel shows the system much later.techniques, such as black hole excision, these simulations will be limit<strong>ed</strong> to thefinal merger phase of black hole coalescence. Hence, it is important that thecommunity continue to focus on this long term solution. However, while that isunder development, we can take advantage of our capabilities and explore thisphase of the inspiral now. Our goal is several fold: (a) to explore new black holephysics of the ‘final plunge’ phase of the binary BH merger; (b) to try to determinesome useful information relevant for gravitational-wave astronomy; and (c) toprovide a strong foundation of knowl<strong>ed</strong>ge for this process that will be useful whenmore advanc<strong>ed</strong> techniques, such a black hole excision, are fully develop<strong>ed</strong>. When


Summary 399these techniques are us<strong>ed</strong> to extend the ability of the community to handle theearlier orbital phase, it will be important to have an understanding of details ofthis most violent phase in advance, both as a testb<strong>ed</strong> to ensure that results arecorrect, and because the understanding we gain may be useful in devising theappropriate techniques for longer-term evolution.18.8 SummaryIn this article we have attempt<strong>ed</strong> to review the essential mathematical andcomputational elements ne<strong>ed</strong><strong>ed</strong> for a full-scale numerical relativity code thatcan treat a variety of problems in relativistic astrophysics and gravitation.Various formulations of the Einstein equations for evolving spacelike time slices,techniques for providing initial data, the basic ideas of gauge conditions, severalimportant analysis tools for discovering the physics contain<strong>ed</strong> in a simulation,and the numerical algorithms for each of these items have been review<strong>ed</strong>.Unfortunately, we have only been able to cover the basics of such a program,and in addition many promising alternative approaches have necessarily been leftout.As one can see, the solution to a single problem in numerical relativityrequires a huge range of computational and mathematical techniques. It is trulya large-scale effort, involving experts in computer and computational science,mathematical relativity, astrophysics, and so on. For these reasons, aid<strong>ed</strong> bycollaborations such as the NSF Black Hole Grand Challenge Alliance and theNCSA/Potsdam/Wash U collaboration, there has been a great focusing of effortover the last years.A natural byproduct of this focusing has been the development of codesthat are us<strong>ed</strong> and extend<strong>ed</strong> by large groups. A code must have a large arsenalof modules at its disposal: different initial data sets, gauge conditions, horizonfinders, slicing conditions, waveform extraction, elliptic equation solvers, AMRsystems, boundary modules, different evolution modules, etc. Furthermore,these codes must run efficiently on the most advanc<strong>ed</strong> supercomputers available.Clearly, the development of such a sophisticat<strong>ed</strong> code is beyond any single personor group. In fact, it is beyond the capability of a single community! Differentresearch communities, from computer science, physics and astrophysics, mustwork together to develop such a code.As an example of such a project, the ‘Cactus’ code has been develop<strong>ed</strong>by a large international collaboration [133, 134]. This code is an outgrowthof the last decade of 3D numerical relativity development primarily atNCSA/Potsdam/Wash U, and builds heavily on the experience gain<strong>ed</strong> indeveloping previous generation codes [34, 45, 148]. Cactus has a verymodular structure, allowing different physics, analysis, and computational sciencemodules to be plugg<strong>ed</strong> in. In fact, versions of essentially all the modules list<strong>ed</strong>above are already develop<strong>ed</strong> for the code. For example, several formulations


400 Numerical relativityof Einstein’s equations, including the ADM formalism and the Bona–Massóhyperbolic formulation, can be chosen as input parameters, as can differentgauge conditions, horizon finders, hydrodynamics evolvers, etc. Cactus was alsodesign<strong>ed</strong> as a community code. After first developing and testing it within ourrather large community of collaborators, it is available with full documentation.By having an entire research community using and contributing to such a code,we hope to accelerate the maturation of numerical relativity. Information aboutthe code is available online, and can be access<strong>ed</strong> at http://www.cactuscode.orgAcknowl<strong>ed</strong>gmentsIt is a pleasure to acknowl<strong>ed</strong>ge many friends and colleagues who have contribut<strong>ed</strong>to the work describ<strong>ed</strong> in this article, some of which was deriv<strong>ed</strong> from papers wehave written together. I especially thank Wai-Mo Suen, who help<strong>ed</strong> write previousreviews on which this article is partly bas<strong>ed</strong>. The Cactus code was originallystart<strong>ed</strong> by Joan Massó and Paul Walker at AEI. Without the contributions frompeople at many institutions, the work describ<strong>ed</strong> here would not have beenpossible. This work has been support<strong>ed</strong> by AEI, NCSA, NSF Grant No PHY-96-00507, NASA HPCC/ESS Grand Challenge Applications Grant No NCCS5-153and NSF MRAC Allocation Grant No MCA93S025.18.9 Further readingHere are some references that we think will help fill in the details of many issueswe can only gloss over. These references are clearly not complete; it was justeasier for us to heavily bias this list towards work coming out of our own group,or closely associat<strong>ed</strong> groups! Our apologies to many others who have written finepapers on these subjects, but we have tri<strong>ed</strong> to give references that are current andrelevant to the most important topics in numerical relativity, if incomplete andbias<strong>ed</strong>.18.9.1 Overviews/formalisms of numerical relativityFor basic 3 + 1 formalism, see [10] and the PhD thesis of Cook [88]. Thisprovides the basics in a very clear, readable way. A somewhat more recent Yorkarticle describes many ‘miscellaneous’ topics, such as more modern initial dataand apparent horizon conditions, etc [20]. However, there are no details on morerecent reformulations of Einstein’s equations for numerical relativity, which arebecoming very important. This is a breaking research area, with new papers everymonth, but some that stand out for hyperbolicity are [32–34, 46, 52, 63, 70, 149,150, and references therein]. Even these are being overtaken by some recentdevelopments! Every month there is new excitement in some variations on all


Further reading 401the above that seem to have very good stability properties: [48,51] are among themore recent ones.For the most recent overviews of numerical relativity especially relat<strong>ed</strong> toblack holes, see the review articles by the author and Wai-Mo Suen: [1, 27, 151–154].18.9.2 Numerical techniquesAn old but still very useful primer on numerical techniques for numericalrelativity can be found in a little article by Smarr in [155]. More moderntreatments for solving PDEs are available in Numerical Recipes [66]. Forhyperbolic systems, the one we learn<strong>ed</strong> everything from is [69].18.9.3 Gauge conditionsFor gauge conditions, we recommend the classics: York [10] and Smarr and York[79, 80] for standard maximal slicing and variational principle shift conditions(minimal distortion, etc). For more modern views on how to actually implementsuch conditions more effectively, including the ‘driver’ ideas, see [87] and thevery recent paper [51]. For work on so-call<strong>ed</strong> algebraic slicing conditions, see[33, 45], and for problems that can develop with such conditions, see [83, 84].For the most recent ideas on shift conditions, see [156]. There is still a lotof work to do here, especially on shift conditions: please publish some ideasyourselves! This is a crucial area of ne<strong>ed</strong><strong>ed</strong> research in numerical relativity thathas not receiv<strong>ed</strong> much attention, especially in 3D.18.9.4 Black hole initial dataThere are by now many black hole initial data sets. There are early references byMisner [157,158], and Brill and Lindquist [115] and then Bowen and York [159],but more recent ones cover the same older material sufficiently. Take a lookat [21,26] and references therein for the classic work. [160] looks at some physicsof initial data sets. For the very large family of distort<strong>ed</strong> black hole plus Brill wav<strong>ed</strong>ata sets, check out [161], or including rotation [128]. These were extend<strong>ed</strong> to 3Dand discuss<strong>ed</strong> briefly in [132]. More recently, important new ways of determininginitial data were develop<strong>ed</strong> by Brandt and Bru¨gmann [25]; see also [18, 162].There are many others!18.9.5 Black hole evolution18.9.5.1 Spherical and distort<strong>ed</strong> (axisymmetric) black holesThere have been extensive studies of numerical evolution of distort<strong>ed</strong> black holes.For that we would check out [73] for 1D (spherical, undistort<strong>ed</strong>), [59, 163, 164]for 2D, and finally [128, 129] for 2D rotating black holes.


402 Numerical relativity18.9.5.2 Colliding black holesFor colliding Misner black holes, the original work of Smarr and others isremarkable: see [165–167]. For a more advanc<strong>ed</strong> attempt see [168, 169]. Forboost<strong>ed</strong> black holes, see [170], and for unequal mass black hole collisions see[171]. For an alternative type of black hole collision, with particles forming blackholes, see [110] and references therein.18.9.5.3 3D black holesIn 3D, so far little has been publish<strong>ed</strong>. 3D Schwarzschild was evolv<strong>ed</strong> numericallyby [45, 60]. Distort<strong>ed</strong> 3D black holes were studi<strong>ed</strong> numerically, and compar<strong>ed</strong> toperturbation theory for the first time, in [131, 132].Colliding black holes in 3D were studi<strong>ed</strong> in [172], and the first full 3Dcollisions were perform<strong>ed</strong> in [2, 147].For a completely alternative approach to that consider<strong>ed</strong> here, characteristicevolution techniques have been us<strong>ed</strong> very successfully to evolve black holes in [8].18.9.6 Black hole excisionFor black hole excision, see [105] for an early 1D success (but the idea wasfloating around long before). Then [173] follow<strong>ed</strong> up with a more advanc<strong>ed</strong>treatment. Other successful 1D work includes [43, 109, 174]. [6, 7, 45, 60] givethe first successful, and increasingly complex, 3D attempts.18.9.7 Perturbation theory and waveform extractionFor work on the perturbative/numerical synergy approach, there are many papers.The original that really start<strong>ed</strong> it all, for Misner data, was [175], follow<strong>ed</strong> by[176], and with boost<strong>ed</strong> holes, [170]. An extraordinary paper that carries theMisner problem to second order is [177]. [178] gives a review. For the movetowards the Teukolsky formalism for rotating black holes, see [179–181], amongothers.For the first attempts to use perturbation theory for 3D distort<strong>ed</strong> black holes,showing the incr<strong>ed</strong>ible accuracy one can achieve in waveforms, see [132].18.9.8 Event and apparent horizonsFor a nice description of the apparent and event horizons, see [182]. Techniquesto find apparent horizons abound; the most recent variations can be found in[124, 125, 145, 183, 184].Two different event horizon finding methods are describ<strong>ed</strong> in [103,110]; thelatter method is us<strong>ed</strong> now by all groups we are aware of. More details of themethod can be found in [104].


References 403Physics that can be extract<strong>ed</strong> by studying numerically generat<strong>ed</strong> horizonshas been detail<strong>ed</strong> in [110, 111, 113, 164, 185, 186].18.9.9 Pure gravitational <strong>waves</strong>We have focus<strong>ed</strong> almost exclusively on black holes in these lectures, but we alsotouch<strong>ed</strong> on recent and very exciting work in pure gravitational wave evolutions.For the original classic (a ‘must read paper’), see [142]. Many followers of the‘Brill wave’ school are [86, 116, 143, 187], and most recently, for the first true,long-term 3D evolutions, see [50].18.9.10 Numerical codesFor Cactus, see [34, 133–135, 188]. Cactus, a full 3D numerical relativity code,is available as a community code this year. Please feel free to test it out andcontribute to it!References[1] Seidel E and Suen W-M 1999 J. Comput. Appl. Math.[2] Seidel E 1999 Proc. Yukawa Conf. (Kyoto)[3] Flanagan É É and Hughes S A 1998 Phys. Rev. D 57 4566 (gr-qc/9710129)[4] Flanagan É É and Hughes S A 1998 Phys. Rev. D 57 4535 (gr-qc/9701039)[5] For a description of the project, seehttp://www.aei-potsdam.mpg.de/research/astro/eu network/index.html[6] Daues G E 1996 PhD Thesis Washington University, St Louis, Missouri[7] Cook G B et al 1998 Phys. Rev. Lett. 80 2512[8] Gomez R et al 1998 Phys. Rev. Lett. 80 3915 (gr-qc/9801069)[9] Nakamura T and Oohara K 1999 Preprint gr-qc/9812054[10] York J 1979 Sources of <strong>Gravitational</strong> Radiation <strong>ed</strong> L Smarr (Cambridge:Cambridge University Press)[11] Bishop N et al 1998 On the Black Hole Trail <strong>ed</strong> B Iyer and B Bhawal (Dordrecht:Kluwer) (gr-qc/9801070)[12] Fri<strong>ed</strong>rich H 1981 Proc. R. Soc. A 375 169[13] Fri<strong>ed</strong>rich H 1981 Proc. R. Soc. A 378 401[14] Fri<strong>ed</strong>rich H 1996 Class. Quantum Grav. 13 1451[15] Hübner P 1996 Phys. Rev. D 53 701[16] Bernstein D and Holst M Private communication[17] Laguna P Private communication[18] Thornburg J 1998 Phys. Rev. D 59 104007[19] Miller M Private communication[20] York J 1989 Frontiers in Numerical Relativity <strong>ed</strong> C Evans, L Finn and D Hobill(Cambridge: Cambridge University Press) pp 89–109[21] Cook G B et al 1993 Phys. Rev. D 47 1471[22] Nakamura T and Oohara K 1989 Frontiers in Numerical Relativity <strong>ed</strong> C Evans,L Finn and D Hobill (Cambridge: Cambridge University Press) pp 254–80


404 Numerical relativity[23] Ashby S F, Manteuffel T A and Saylor P E 1990 SIAM J. Num. Anal. 27 1542[24] Cook G B 1989 Frontiers in Numerical Relativity <strong>ed</strong> C Evans, L Finn and D Hobill(Cambridge: Cambridge University Press) and references therein[25] Brandt S and Brügmann B 1997 Phys. Rev. Lett. 78 3606[26] Cook G B 1991 Phys. Rev. D 44 2983[27] Seidel E 1996 Relativity and Scientific Computing <strong>ed</strong> F Hehl (Berlin: Springer)[28] Brady P R, Creighton J D E and Thorne K S 1998 Phys. Rev. D 58 061501[29] Arnowitt R, Deser S and Misner C W 1962 Gravitation: An Introduction to CurrentResearch <strong>ed</strong> L Witten (New York: Wiley) pp 227–65[30] Choquet-Bruhat Y and Ruggeri T 1983 Commun. Math. Phys. 89 269[31] Bona C and Massó J 1992 Phys. Rev. Lett. 68 1097[32] Bona C, Massó J, Seidel E and Stela J 1995 Phys. Rev. Lett. 75 600[33] Bona C, Massó J, Seidel E and Stela J 1997 Phys. Rev. D 56 3405[34] Bona C, Massó J, Seidel E and Walker P 1998 Preprint gr-qc/9804052[35] Alcubierre M et al in preparation[36] Frittelli S and Reula O 1994 Commun. Math. Phys. 166 221[37] Choquet-Bruhat Y and York J 1995 C. R. Acad. Sci., Paris 321 1089[38] Abrahams A, Anderson A, Choquet-Bruhat Y and York J 1995 Phys. Rev. Lett. 753377[39] Frittelli S and Reula O 1996 Phys. Rev. Lett. 76 4667[40] van Putten M H and Eardley D 1996 Phys. Rev. D 53 3056[41] Abrahams A, Anderson A, Choquet-Bruhat Y and York J 1997 Class. QuantumGrav. A 9[42] Bona C, Massó J and Stela J 1995 Phys. Rev. D 51 1639[43] Scheel M et al 1997 Phys. Rev. D 56 6320[44] Cook G and Scheel M Private communication[45] Anninos P et al 1995 Phys. Rev. D 52 2059[46] Reula O 1998 Living Reviews in Relativity 1[47] Fri<strong>ed</strong>rich H 1985 Commun. Math. Phys. 100 525[48] Baumgarte T W and Shapiro S L 1999 Phys. Rev. D 59 024007[49] Shibata M and Nakamura T 1995 Phys. Rev. D 52 5428[50] Alcubierre M et al 2000 Phys. Rev. D D61 041501 (gr-qc/9904013)[51] Arbona A, Bona C, Massó J and Stela J 1999 Phys. Rev. D 60 104014 (grqc/9902053)[52] Alcubierre M et al 1999 Preprint gr-qc/9908079[53] Nakamura T, Oohara K and Kojima Y 1987 Prog. Theor. Phys. Suppl. 90 1[54] Shibata M 1999 Prog. Theor. Phys. 101 1199 (gr-qc/9905058)[55] Baumgarte T W, Hughes S A and Shapiro S L 1999 Phys. Rev. D 59 024007[56] Shibata M 1999 Phys. Rev. D 60 104052 (gr-qc/9908027)[57] Shibata M and Uryu K 2000 Phys. Rev. D 61 064001 (gr-qc/9911058)[58] Shibata M, Baumgarte T W and Shapiro S L 2000 Phys. Rev. D 61 044012[59] Bernstein D et al 1994 Phys. Rev. D 50 5000[60] Brügmann B 1996 Phys. Rev. D 54 7361[61] New K C B, Watt K, Misner C W and Centrella J M 1998 Phys. Rev. D 58 064022(gr-qc/9801110)[62] Huq M Private communication[63] Alcubierre M et al 2000 Phys. Rev. D 62 044034[64] Teukolsky S 2000 Phys. Rev. D 61 087501


References 405[65] Gustafsson B, Kreiss H-O and Oliger J 1995 Time Dependent Problems andDifference Methods (New York: Wiley)[66] Press W H, Flannery B P, Teukolsky S A and Vetterling W T 1986 NumericalRecipes (Cambridge: Cambridge University Press)[67] Clune T, Massó J, Miller M and Walker P Technical Report National Center forSupercomputing Applications, unpublish<strong>ed</strong>, in preparation[68] Massó J 1992 PhD Thesis University of the Balearic Islands[69] Leveque R J 1992 Numerical Methods for Conservation Laws (Basel: Birkhauser)[70] Bona C 1996 Relativity and Scientific Computing <strong>ed</strong> F Hehl (Berlin: Springer)[71] Yee H C 1989 Computational Fluid Dynamics (NASA Ames Research Center, CA:von Karman Institute for Fluid Dynamics)[72] Choptuik M 1991 Phys. Rev. D 44 3124[73] Bernstein D, Hobill D and Smarr L 1989 Frontiers in Numerical Relativity <strong>ed</strong>C Evans, L Finn and D Hobill (Cambridge: Cambridge University Press) pp 57–73[74] Lee S 1993 PhD Thesis University of Illinois Urbana-Champaign[75] Ashby S et al 1995 Elsevier Sci.[76] Detweiler S 1987 Phys. Rev. D 35 1095[77] Brodbeck O, Frittelli S, Hübner P and Reula O A 1998 gr-qc/9809023[78] C W L et al 1998 Preprint Physics Department, Chinese University of Hong Kong[79] Smarr L and York J 1978 Phys. Rev. D 17 1945[80] Smarr L and York J 1978 Phys. Rev. D 17 2529[81] Piran T 1983 <strong>Gravitational</strong> Radiation <strong>ed</strong> N Deruelle and T Piran (Amsterdam:North-Holland)[82] Bona C and Massó J 1988 Phys. Rev. D 38 2419[83] Alcubierre M 1997 Phys. Rev. D 55 5981[84] Alcubierre M and Massó J 1998 Phys. Rev. D 57 4511[85] Tobias M 1997 PhD Thesis Washington University, Saint Louis, MO[86] Eppley K 1979 Sources of <strong>Gravitational</strong> Radiation <strong>ed</strong> L Smarr (Cambridge:Cambridge University Press) p 275[87] Balakrishna J et al 1996 Class. Quantum Grav. 13 L135[88] Cook G 1990 PhD Thesis University of North Carolina at Chapel Hill, Chapel Hill,North Carolina[89] Martí JM,Ibánez J M and Miralles J A 1991 Phys. Rev. D 43 3794[90] Font J, Ibánez J, Martí J and Marquina A 1994 Astron. Astrophys. 282 304[91] Banyuls F et al 1997 Ap. J. 476 221[92] Donat R, Font J A, Ibánez J M and Marquina A 1998 J. Comput. Phys. 146 58[93] For the NASA Neutron Star Grand Challenge Project, see, e.g.,http://wugrav.wustl.<strong>ed</strong>u/Relativ/nsgc.html.[94] Wang E and Swesty F D 1998 Proc. 18th Texas Symp. on Relativistic Astrophysics<strong>ed</strong> J F A Olinto and D Schramm (Singapore: World Scientific)[95] Font J A, Miller M, Suen W M and Tobias M 2000 Phys. Rev. D 61 044011 (grqc/9811015)[96] Roe P L 1981 J. Comput. Phys. 43 357[97] Donat R and Marquina A 1996 J. Comput. Phys. 125 42[98] Eulderink F and Mellema G 1994 Astron. Astrophys. 284 652[99] Shu C W and Osher S J 1989 J. Comput. Phys. 83 32[100] Abrahams A M et al 1998 Phys. Rev. Lett. 80 1812 (gr-qc/9709082)


406 Numerical relativity[101] Shapiro S L and Teukolsky S A 1986 Dynamical Spacetimes and NumericalRelativity <strong>ed</strong> J M Centrella (Cambridge: Cambridge University Press) pp 74–100[102] Thornburg J 1987 Class. Quantum Grav. 14 1119 (Unruh is cit<strong>ed</strong> here by Thornburgas originating AHBC)[103] Anninos P et al 1995 Phys. Rev. Lett. 74 630[104] Libson J et al 1996 Phys. Rev. D 53 4335[105] Seidel E and Suen W-M 1992 Phys. Rev. Lett. 69 1845[106] Alcubierre M and Schutz B 1994 J. Comput. Phys. 112 44[107] Anninos P et al 1995 Phys. Rev. D 51 5562[108] Scheel M A, Shapiro S L and Teukolsky S A 1995 Phys. Rev. D 51 4208[109] Marsa R and Choptuik M 1996 Phys. Rev. D 54 4929[110] Hughes S et al 1994 Phys. Rev. D 49 4004[111] Matzner R et al 1995 Science 270 941[112] Massó J, Seidel E, Suen W-M and Walker P 1999 Phys. Rev. D 59 064022 (grqc/9804059)[113] Shapiro S, Teukolsky S and Winicour J 1995 Phys. Rev. D 52 6982[114] Hawking S W and Ellis G F R 1973 The Large Scale Structure of Spacetime(Cambridge: Cambridge University Press)[115] Brill D and Lindquist R 1963 Phys. Rev. 131 471[116] Eppley K 1977 Phys. Rev. D 16 1609[117] Anninos P et al 1998 Phys. Rev. D 58 024003[118] Libson J, Massó J, Seidel E and Suen W-M 1996 The Seventh Marcel GrossmannMeeting: On Recent Developments in Theoretical and Experimental GeneralRelativity, Gravitation, and Relativistic Field Theories <strong>ed</strong> R T Jantzen,G M Keiser and R Ruffini (Singapore: World Scientific) p 631[119] Libson J 1994 Numerical Relativity Conference <strong>ed</strong> P Laguna (Urbana, IL: GrandChallenge Meeting, NCSA, Champaign)[120] Camarda K 1998 PhD Thesis University of Illinois at Urbana-Champaign, Urbana,Illinois[121] Camarda K and Seidel E 1999 Phys. Rev. D 59 064026 (gr-qc/9805099)[122] Baumgarte T W et al 1996 Phys. Rev. D 54 4849[123] Nakamura T, Kojima Y and Oohara K 1984 Phys. Lett. A 106 235[124] Gundlach C 1998 Phys. Rev. D 57 863 (gr-qc/9707050)[125] Thornburg J 1996 Phys. Rev. D 54 4899[126] Huq M Private communication, unpublish<strong>ed</strong>[127] Moncrief V 1974 Ann. Phys. 88 323[128] Brandt S and Seidel E 1996 Phys. Rev. D 54 1403[129] Brandt S and Seidel E 1995 Phys. Rev. D 52 856[130] Brandt S and Seidel E 1995 Phys. Rev. D 52 870[131] Allen G, Camarda K and Seidel E 1998 gr-qc/9806014[132] Allen G, Camarda K and Seidel E 1998 gr-qc/9806036[133] Allen G, Goodale T and Seidel E 1999 7th Symposium on the Frontiers of MassivelyParallel Computation-Frontiers 99 (New York: IEEE)[134] Allen G et al 1999 IEEE Comput. 32[135] http://www.cactuscode.org[136] Bonazzola S, Gourgoulhon E and Marck J-A 1998 Preprint astro-ph/9803086[137] Choptuik M 1989 Frontiers in Numerical Relativity <strong>ed</strong> C Evans, L Finn and D Hobill(Cambridge: Cambridge University Press)


References 407[138] Berger M and Oliger J 1984 J. Comput. Phys. 53 484[139] Choptuik M 1993 Phys. Rev. Lett. 70 9[140] Wild L and Schutz B in preparation[141] Papadopoulos P, Seidel E and Wild L 1998 Phys. Rev. D 58 084002 (gr-qc/9802069)[142] Brill D S 1959 Ann. Phys. 7 466[143] Holz D, Miller W, Wakano M and Wheeler J 1993 Directions in General Relativity:Proc. 1993 Int. Symp., Maryland; Papers in Honor of Dieter Brill <strong>ed</strong>BHuandT Jacobson (Cambridge: Cambridge University Press)[144] Brandt S, Camarda K and Seidel E 2000 Proc. 8th M Grossmann Meeting: OnRecent Developments in Theoretical and Experimental Relativity <strong>ed</strong> T Piran(Singapore: World Scientific)[145] Alcubierre M et al 2000 Class. Quantum Grav. 17 2159[146] Camarda K and Seidel E 1998 Phys. Rev. D 57 R3204 (gr-qc/9709075)[147] Brügmann B 1999 Int. J. Mod. Phys. D 8 85[148] Anninos P et al 1997 Phys. Rev. D 56 842[149] Anderson A and York J 1999 Preprint gr-qc/9901021[150] Alcubierre M, Brügmann B, Miller M and Suen W-M 1999 Phys. Rev. D 60 064017(gr-qc/9903030)[151] Seidel E 1997 Gravitation and Cosmology <strong>ed</strong> S Dhurandhar and T Padmanabhan(Dordrecht: Kluwer) pp 125–44[152] Seidel E 1996 Helv. Phys. Acta 69 454[153] Seidel E 1998 Proc. GR15 <strong>ed</strong> N Dadich[154] Seidel E 1998 Bad Honnef Meeting on Black Holes <strong>ed</strong> F Hehl (Berlin: Springer)[155] Smarr L 1979 Sources of <strong>Gravitational</strong> Radiation <strong>ed</strong> L Smarr (Cambridge:Cambridge University Press) p 139[156] Thorne K 1999 Phys. Rev. D (gr-qc/9808024)[157] Misner C W 1963 Ann. Phys. 24 102[158] Misner C 1960 Phys. Rev. 118 1110[159] Bowen J and York J W 1980 Phys. Rev. D 21 2047[160] Cook G B 1994 Phys. Rev. D 50 5025[161] Bernstein D, Hobill D, Seidel E and Smarr L 1994 Phys. Rev. D 50 3760[162] Matzner R A, Huq M F and Shoemaker D 1999 Phys. Rev. D 59 024015[163] Anninos P et al 1997 Computational Astrophysics: Gas Dynamics and ParticleMethods <strong>ed</strong> W Benz et al (New York: Springer) in press[164] Anninos P et al 1993 IEEE Comput. Graphics Appl. 13 12[165] Smarr L, Čadež A, DeWitt B and Eppley K 1976 Phys. Rev. D 14 2443[166] Smarr L 1977 Ann. N. Y. Acad. Sci. 302 569[167] Smarr L 1979 Sources of <strong>Gravitational</strong> Radiation <strong>ed</strong> L Smarr (Cambridge:Cambridge University Press) p 245[168] Anninos P et al Technical Report No 24, National Center for SupercomputingApplications, unpublish<strong>ed</strong>[169] Anninos P et al 1995 Phys. Rev. D 52 2044[170] Baker J et al 1997 Phys. Rev. D 55 829[171] Anninos P and Brandt S 1998 Phys. Rev. Lett. 81 508[172] Anninos P, Massó J, Seidel E and Suen W-M 1996 Physics World 9 43[173] Anninos P et al 1996 The Seventh Marcel Grossmann Meeting: On RecentDevelopments in Theoretical and Experimental General Relativity, Gravitation,and Relativistic Field Theories <strong>ed</strong> R T Jantzen, G M Keiser and R Ruffini


408 Numerical relativity(Singapore: World Scientific) p 637[174] Scheel M A et al 1998 Phys. Rev. D 58 044020[175] Price R H and Pullin J 1994 Phys. Rev. Lett. 72 3297[176] Anninos P et al 1995 Phys. Rev. D 52 4462[177] Gleiser R J, Nicasio C O, Price R H and Pullin J 1996 Phys. Rev. Lett. 77 4483[178] Pullin J 1998 Proc. GR15 <strong>ed</strong> N Dadhich and J Narlikar (Puna: Inter Univ. Centrefor Astr. and Astrop.) p 87[179] Campanelli M and Lousto C O 1998 Phys. Rev. D 58 024015[180] Campanelli M, Krivan W and Lousto C O 1998 Phys. Rev. D 58 024016[181] Campanelli M et al 1998 Phys. Rev. D 58 084019[182] Hawking S W 1973 Black Holes <strong>ed</strong> C DeWitt and B S DeWitt (New York: Gordonand Breach) pp 1–55[183] Anninos P et al 1998 Phys. Rev. D 58 24003[184] Shibata M 1997 Phys. Rev. D 55 2002[185] Anninos P et al 1994 Phys. Rev. D 50 3801[186] Anninos P et al 1995 Austral. J. Phys. 48 1027[187] Miyama S M 1981 Prog. Theor. Phys. 65 894[188] Seidel E 1999 Forschung und wissenschaftliches Rechnen <strong>ed</strong> T P and P Wittenburg(München: Max-Planck-Gesselschaft)


Indexaction, 50Einstein–Hilbert, 247affine Kac–Moody algebra, 7AIGO, 26ALLEGRO, 99anholonomy, 246antennaresonant, 4approximationfar-field, 58short-wave, 15, 52weak-field, 15AURIGA, 38, 99, 197back reaction evasion, 37beamsplitter, 5big bang, 48binarypulsar PSR 1913+16, 3, 44stars, 44, 71system, 153chirping, 44, 73coalescing, 45, 171, 174inspiralling, 338black hole, 24, 43, 45central charge, 261CFS instability, 75chirpmass, 72, 165signal, 167collapsing star, 153conformal factor, 7conformal invariance, 320constraints, 250in relativity, 365, 373continuous signals, 32, 38cosmic strings, 214cosmological constant, 153cosmological perturbation theory,297cross-correlations, 185overlap r<strong>ed</strong>uction function, 191signal-to-noise ratio, 186curvature produc<strong>ed</strong> by <strong>waves</strong>, 55dark matter, 49decay time, 155detectorbar, 4, 25, 34, 152cryogenic, 25quantum limit noise, 37sensor noise, 36thermal noise, 35beam, 20cryogenic bar, 5hollow sphere, 153, 168, 170, 173icosah<strong>ed</strong>ral, 26, 38interaction with <strong>waves</strong>, 19interferometer, 28, 183quantum limit noise, 172, 174resonant bar, 196resonant mass, 4, 34, 91, 152sensitivity, 99sphere, 196spheres, 6spherical, 26, 38, 152, 165dimensional r<strong>ed</strong>uction, 247drag-free control, 39409


410 IndexEhlers Lagrangian, 254Einstein equations, 245including matter, 51lineariz<strong>ed</strong> theory, 16vacuum, 51elasticity equation, 169Ernst equation, 255, 262EXPLORER, 99, 197Fermi–Walker transport, 8fieldgravitomagnetic, 3axion, 250current-quadrupole, 60, 65, 66dilaton, 7, 250duality symmetry, 284mass-octupole, 60, 65mass-quadrupole, 60, 62, 72scalar, 153forc<strong>ed</strong>riving, 156tidal, 18frameJordan–Fierz, 160local free-falling, 153frequencybandwidth, 5monopole resonant, 159gamma-ray bursts, 45, 46gaugecondition, 161, 247conformal, 7, 249Lorentz–Hilbert, 16preserving choice, 253transformation, 16, 52transverse-traceless, 4triangular, 254TT, 17gauge conditionsin relativity, 373GEO 600, 26, 33, 195geodesic equation, 18geon, 3Geroch group, 7, 259gravitationalantenna, 315collapse, 43, 165, 171, 173geon, 3radiation, 15self-energy, 163stress-energy tensor, 54gravitational wave, 15action, 52burst, 27, 32, 37energy, 55continuous, 27cosmological relic, 7in lineariz<strong>ed</strong> theory, 15in short-wave approximation, 52Jordan–Brans–Dicke, 169, 171observable, 26polarization states, 159scalar, 6, 152, 159stochastic, 7graviton spectrum, 304, 308groupsymmetry, 7gyroscope, 8, 268, 271, 273, 276precession, 271, 276Helmholtz equation, 162horizons, 292event, 386Hubble constant, 56, 73inflation, 223, 289interferometer, 22, 25, 28, 104Fabry–Perot, 28gravity gradient noise, 31Michelson, 28shot noise, 30, 105thermal noise, 30, 106uncertainty principle noise, 31vibrational noise, 31Isaacson tensor, 54JGWO, 26


Index 411Kaluza–Klein, 7, 153, 247, 251Killing vector, 7, 245, 247Kramer–Neugebauertransformation, 256LagrangianEhlers, 7Einstein–Hilbert, 7Matzner–Misner, 7Lamé coefficients, 154Lie group, 252LIGO, 5, 26, 34, 184LISA, 5, 26, 38, 115, 185antenna, 116loop corrections, 304masstest, 5Matzner–Misner Lagrangian, 252,255metricFerrari–Ibanez, 266Minkowski, 16trace-revers<strong>ed</strong> perturbation, 16minisuperspace, 295modesmonopole, 159, 169normal, 155, 168quadrupole, 169Rossby, 75scalar, 168spheroidal, 155, 157toroidal, 155, 156vibrational, 158Multiple moments, 339, 350–352NAUTILUS, 38, 99, 197neutron star, 45, 46, 79accreting, 46Newman–Penrose parameter, 158NIOBE, 99noiseground, 6narrowband, 167photon shot, 5seismic, 5, 6wideband, 167nonlinear σ -model, 7, 252nucleosynthesis, 310O(d, d) symmetry, 327optical theorem, 170Optimal signal filtering, 340, 341,344, 354perturbationof background metric, 16, 50velocity, 75phase transitions, 235photodetector, 5polarizationcircular, 18collinear, 8non-collinear, 8tensor, 17, 26Post-Newtonianapproximation, 339, 346expansion, 339, 347, 350, 351power recycling, 30pre-big bang, string cosmology, 281primordial magnetic fields, 296pulsar, 47timing, 22pulsars, 46quality factor, 4, 155r-modes, 4, 43, 46, 73, 75damping time, 76linear evolution, 76nonlinear evolution, 77radiationcurrent-quadrupole, 4, 75mass-quadrupole, 4, 69ranging to spaceraft, 21relic axions, 296relic gravitons, 295retard<strong>ed</strong> integral, 58


412 Indexscalarspherical harmonics, 162seismic isolationSuper-Attenuator, 103seismic noiseSuper-Attenuator, 108anti-spring, 109blade spring, 108control, 111Invert<strong>ed</strong> Pendulum, 110self-dual cosmology, 286σ model, 257signal recycling, 32signal-to-noise ratio, 165, 167, 174spin connection, 246SQUID, 4stochastic background, 27, 32, 38,91, 98astrophysical sources of, 237cosmological, 49, 55observational bounds about, 207r-modes, 48, 79stringcosmology, 8effective action, 317frame, 290pre-big bang cosmology, 283string cosmology, 229superinflation, 290supernovae, 43, 79supersymmetry, 288TAMA, 26, 33, 195theoryJordan–Brans–Dicke, 152, 153,159Kaluza–Klein, 153lineariz<strong>ed</strong>, 15post-Newtonian, 15scalar–tensor, 169thermal noiseLow Frequency Facility, 103, 111Fabry–Perot, 111R and D SA, 112variational principle, 50vielbein, 246vierbein, 7VIRGO, 5, 26, 34, 103, 184waveplane, 17vector, 17waveformbinary inspiral, 8wavelengthr<strong>ed</strong>uc<strong>ed</strong>, 5Weyl invariance, 249

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!