Você está na página 1de 472

Observational Tests of Cosmological Inflation

NATO ASI Series


Advanced Science Institutes Series

A Series presenting the results of activities sponsored by the NA TO Science Committee,


which aims at the dissemination of advanced scientific and technological knowledge,
with a view to strengthening links between scientific communities.

The Series is published by an international board of publishers in conjunction with the


NATO Scientific Affairs Division

A LHe Sciences Plenum Publishing Corporation


B Physics London and New York

C Mathematical Kluwer Academic Publishers


and Physical Sciences Dordrecht, Boston and London
D Behavioural and Social Sciences
E AppHed Sciences

F Computer and Systems Sciences Springer-Verlag


G Ecological Sciences Berlin, Heidelberg, New York, London,
H Cell Biology Paris and Tokyo
I Global Environmental Change

NATO-PCO-DATA BASE

The electronic index to the NATO ASI Series provides full bibliographical references
(with keywords and/or abstracts) to more than 30000 contributions from international
scientists published in all sections of the NATO ASI Series.
Access to the NATO-peO-DATA BASE is possible in two ways:
- via online FILE 128 (NATO-peO-DATA BASE) hosted by ESRIN,
Via Galileo Galilei, 1-00044 Frascati, Italy.
- via CD-ROM "NATO-PCO-DATA BASE" with user-friendly retrieval software in
English, French and German (©WTV GmbH and DATAWARE Technologies Inc.
1989).

The CD-ROM can be ordered through any member of the Board of Publishers or
through NATO-peO, Overijse, Belgium.

Series C: Mathematical and Physical Sciences - Vol. 348


Observational Tests
of Cosmological Inflation
edited by

T. Shanks
A. J. 8anday
R.S.Ellis
C. S. Frenk
and

A. W. Wolfendale
Physics Department,
University of Durham, UK

..

"
~

Kluwer Academic Publishers


Dordrecht / Boston / London

Published in cooperation with NATO Scientific Affairs Division


Proceedings of the NATO Advanced Research Workshop on
Observational Tests of Inflation,
Durham, UK
December 10-14, 1990

ISBN 0-7923-1431-X

Published by Kluwer Academic Publishers,


P.O. Box 17, 3300 AA Dordrecht, The Netherlands.

Kluwer Academic Publishers incorporates the publishing programmes of


D. Reidel, Martinus Nijhoff, Dr W. Junk and MTP Press.

Sold and distributed in the U.S.A. and Canada


by Kluwer Academic Publishers,
101 Philip Drive, Norwell, MA 02061, U.S.A.

In all other countries, sold and distributed


by Kluwer Academic Publishers Group,
P.O. Box 322,3300 AH Dordrecht, The Netherlands.

Printed on acid-free paper

All Rights Reserved


© 1991 Kluwer Academic Publishers
No part of the material protected by this copyright notice may be reproduced or
utilized in any form or by any means, electronic or mechanical, including photo-
copying, recording or by any information storage and retrieval system, without written
permission from the copyright owner.
PREFACE

This book represents the proceedings from the NATO sponsored Advanced
Research Workshop entitled "Observational Tests of Inflation" held at the University of
Durham, England on the 10th-14th December, 1990. In recent years, the cosmological
inflation model has drawn together the worlds of particle physics, theoretical cosmology
and observational astronomy. The aim of the workshop was to bring together experts in
all of these fields to discuss the current status of the inflation theory and its observational
predictions.

The simplest inflation model makes clear predictions which are testable by
astronomical observation. Foremost is the prediction that the cosmological density
parameter, no, should have a value negligibly different from the critical, Einstein-de
Sitter value of 00=1. The other main prediction is that the spectrum of primordial density
fluctuations should be Gaussian and take the Harrison-Zeldovich form.

The prediction that n o=l, in patticular, leads to several important consequences


for cosmology. Firstly, there is the apparent contradiction with the limits on baryon
density from Big Bang nucleosynthesis which has led to the common conjecture that
weakly interacting particles rather than baryons may form the dominant mass constituent
of the Universe. Secondly, with n o =l, the age of the Universe is uncomfortably short if
the Hubble constant and the ages of the oldest star clusters lie within their currently
believed limits. The workshop therefore took place at an exciting time for cosmology,
with a feeling abroad that either inflation or one of the standard foundations of modern
cosmology might have to be surrendered. Interest was heightened by many, new, ground
and space-based astronomical advances.

The first day of the workshop reviewed the current theoretical status of the
inflation predictions. In succeeding days, the workshop investigated the constraints on no
from the cosmological timescale test, from observations of galaxies at high redshift and
from the dynamics of galaxy clusters. The implications for the primordial spectrum of
density fluctuations from observations of the large-scale galaxy distribution and from
measurements of the isotropy of the microwave background were also discussed. The
far-reaching impact of inflation on cosmology can be judged by the breadth of topics
covered, all of which are reviewed by first rank research workers in this book. Overall,
the workshop's conclusion was that inflation remains viable, although relatively small
improvements in astronomical data will soon test the basic tenets of the theory. For the
time being, inflation is likely to remain the framework for theoretical work in cosmology.

v
vi

I should like to take this opportunity to thank fellow members of the International
Organising Committee, P.J.E. Peebles (Princeton), D.W. Sciama (Trieste), and A.W.
Wolfendale (Durham) and the Local Organising Committee, A.J. Banday, R.S. Ellis and
C.S. Frenk (Durham) for all their help in making the workshop possible. Thanks are also
due to the conference secretary, Margaret Norman, for her efficient administrative and
secretarial skills. We are also grateful to Carol Webster for her help in producing this
manuscript. We should also like to thank R.G. Bower, R.L. Guzman, S.J. Lorrimer, I.R.
Smail and N.R. Tanvir, for the assistance they gave while the workshop was in progress.
Finally we thank the NATO Science Committee for funding assistance.

T. Shanks (Durham)
24th May, 1991
TABLE OF CONTENTS

Preface......................................................................................................................... v

Conference Photograph......... ........... ......... .......... ... ..................... .... ..... ............ ... ........ xii

List of Participants...... .... ............ .................... .... ........ ............... ... .... .... ...... ..... .... .... .... xv

I. INFLATION THEORY

Fundamental Arguments for Inflation


A.H. Guth*..................................................................................................... 1

Predictions of Inflation
A.R. Liddle*................................................................................................... 23

Classicality of Density Perturbations in the Early Universe


R. Brandenberger, R. Laflamme & M. Mijic................................................. 39

The Influence of Non-Linear Density Fluctuations on the Microwave Sky


I. L Sanz & E. Martinez-Gonzalez ................................................................ 47

Quantum Cosmology and the Cosmological Constant


I. Moss............................................................................................................ 53

Lessons from Inflation and Cold Dark Matter


P.I.E. Peebles*...................................................................................... ........ 63

The Topology of Galaxy Clustering


P. Coles & M. Plionis.................................................................................... 75

Can Non-Gaussian Fluctuations for Structure Formation Arise from Inflation?


D.S. Salopek.............................................................................................. ..... 81

Non-Baryonic Dark Matter


S. Sarkar*.................................. .................................................................... 91

Are Galactic Halos Made of Brown Dwarfs or Black Holes?


B.I. Carr........................................................................................................ 103
viii

II. COSMOLOGICAL TIMESCALE TEST

Ages of Globular Clusters


P. Demarque*, c.P. Deliyannis & A. Sarajedini.......................................... 111

Globular Cluster Ages and Cosmology


A. Renzini*..................................................................................................... 131

The Local Distance Scale: How Reliable Is It?


M. W. Feast*................................................................................................... 147

Distances to Virgo and Beyond


M. Rowan-Robinson*.................................................................................... 161

The Luminosity-Line-Width Relations and the Value of Ho


M.l. Pierce..................................................................................................... 173

Observational Status of Ho
G.A. Tammann* ............................................................................................. 179

Calibrating Cepheid Sequences in Nearby Galaxies


N. Metcalfe & T. Shanks................................................................................ 187

New D-cr Results for Coma Ellipticals


l.R. Lucey, R. Guzman, D. Carter & R.I. Terlevich ..... ................................. 193

Novae and the Distance Scale


C.l. Pritchet................................................................................................... 199

A High Resolution, Ground Based Observation of a Virgo Galaxy


T. Shanks, N. Tanvir, P. Doe~ C. Dunlop, R, Myers, l. Major, M. Redfern,
N. Devaney & P. O'Kane............................................................................... 205

Globular Clusters as Extragalactic Distance Indicators


D.A. Hanes ..................................................................................................... 211

III. HIGH REDSHIFT TESTS OF Do


High-Redshift Tests of no
B. Guiderdoni*.............................................................................................. 217

Cosmology with Galaxies at High Redshifts


S.l. Lilly......................................................................................................... 233
ix

Prospects for Measuring the Deceleration Parameter


R.S. Ellis......................................................................................................... 243

Aligned Radio Galaxies


K. C. Chambers............................................................................................... 251

K Band Galaxy Counts and the Cosmological Geometry


LL Cowie...................................................................................................... 257

Selection Effects in Redshift Surveys


Y. Yoshii & M. Fukugita ................................................................................ 267

An Inflationary Alternative to the Big-Bang


F. Hoyle ......................................................................................................... 273

IV. GALAXY CLUSTERING, 00 AND THE PRIMORDIAL SPECTRUM

Dynamical Estimates of no
from Galaxy Clustering
S.D.M. White*................................................................................................ 279

ROSAT Observations of Clusters of Galaxies


H. Biihringer*, W. Voges, H. Ebeling, R.A. Schwarz, A.C. Edge, V.G.
Briel and J.P. Henry ...................................................................................... 293

A Deep ROSAT Observation at High Galactic Latitude


I. Georgantopoulos, T. Shanks, G. Stewart, K. Pounds, RJ. Boyle & R.
Griffiths.......................................................................................................... 309

Large Scale Structure and Inflation


J.P. Huchra*.................................................................................................. 315

The Structure of the Universe on Large Scales


A.G. Doroshkevich......................................................................................... 327

Testing the Zeldovich Spectrum


w.J. Sutherland.............................................................................................. 331

Q on the Scale of 3Mmls


D. Lynden-Bell*............................................................................................. 337

Tests of Inflation Using the QDOT Redshift Survey


C.S. Frenk. ..................................................................................................... 355
x

Testing Inflation with Peculiar Velocities


A. Dekel.. ........................................................................................................ 365

The Invisible Cosmological Constant


O. Lahav, P.B. Lilje, J.R. Primack & M.J. Rees............................................ 375

Support for Inflation from the Great Attractor


A. Heavens..................................................................................................... 379

The Angular Large Scale Structure


y. Hoffinan ..................................................................................................... 385

Is There Any Observational Evidence for Non- Gaussian Primordial Density


Fluctuations?
A.L Melott..................................................................................................... 389

V. MICROWAVE BACKGROUND ANISOTROPY

COBE DMR Observations of CMB Anisotropy


G.F. Smoot*................................................................................................... 395

Observations of Microwave Background Anisotropy at Tenerife and Cambridge


A.N. Lasenby, R.D. Davies, R.A. Watson, R. Rebolo, C. Gutierrez & J.E.
Beckman......................................................................................................... 413

Foreground Effects and the Search for Fluctuations in the CMB Radiation
A.J. Banday, M. Giler, B. Szabelska, J. Szabelski & A. W
Wolfendale ..................................................................................................... 419

Microwave Background Anisotropies and Large Scale Structure in the Universe


G. Efstathiou*................................................................................................ 425

Discovery of the Small Scale Sky Anisotropy at 2.7cm: Radio Sources or Relic
Emission?
Yu. N. Parijskij, B.L Erukhimov, M.G. Mingaliev, A.B. Berlin, N.N.
Bursov, N.A.Nizhelskij, M.N. Naugolnaja, v.N. Chernenkov, O. V.
Verkhodanov, A. V. Chepurnov & A.A.
Starobinsky.................................................................................................... 437

Balloon-Borne Observations of CMB Anisotropies at Intennediate Angular Scales,


at Sub-MM and MM Wavelengths
P. de Bernardis, S. Masi, B. Melchiorri & F. Melchiorri............................ 443
xi

VI. POSTER PAPERS

The Durham/UKST Galaxy Redshift Survey


A Broadbent, D. Hale-Sutton, T. Shanks, F.G. Watson, AP. Oates, R.
Fong, C.A Collins, H.T. MacGillivray, R. Niclwl & Q.A..
Parker............................................................................................................ 447

Time Evolution of Lensed Image Separations


T.l. Broadhurst & S. Oliver........................................................................... 449

Deep Galactic Surveys as Probes of the Large Scale Structure of the Universe
O.E. Buryak. M. Demia'nski & A.G. Doroshkevich. ..................................... 453

Intergalactic Absorption in the Spectra of High-Redshift QSOs


S. Cristiani & E. Giallongo........................................................................... 457

A Complete Quasar Sample at Intermediate Redshift


F. La Franca, S. Cristiani, C. Barbieri, R.G. Clowes & A Iovino ................ 461

Radio-Luminosity Dependence of the IR-Radio Alignment Effect in High-z Radio


Galaxies
l.S. Dunlop & l.A. Peacock........................................................................... 463

Density and Peculiar Velocity Fields in the Region of Dressler's Supergalactic


Plane Survey
M.l. Hudson ................................................................................................... 467

Scale Invariance Induced by Non-linear Growth of Density Fluctuations


F. Moutarde, l.-M Alimi, F.R. Bouchet & R. Pellat... ................................... 469

The Power Spectrum of Galaxy Clustering


l.A. Peacock................................................................................................... 471

Higher Moments of the IRAS Galaxy Distribution


C.A. Scharf .................................................................................................... 475

Collapse of a Protogalactic Cloud


S. Yoshioka. .................................................................................................... 477

INDEX OF AUTHORS............................................................................................. 479

SUBJECT INDEX ...................................................................................................... 481

* denotes invited speaker


NATO ADVANCED RESEARCH WORKSHOP

OBSERVATIONAL TESTS OF INFLATION


The University of Durham. December 10-14. 1990

f\
A
1. Richard Ellis 19. Tom Broadhurst 37. Alan Heavens 55. Stephen Lorrimer
2. Pierre Demarque 20. Andy Taylor 3S. Rudolf Treumann 56. Andrei Doroshkevich
3. Bruno Guiderdoni 21. Paul Saleh 39. Yuzuru Yoshii 57. Jose Luis San7.
4. George Smoot 22. Peter Coles 40. Dave Salopek 58. Mike Fitchett
5. Simon White 23. Chris Pritchet 41. Stefano Cristiani 59. Marat Mingaliev
6. Alan Guth 24. Mike Hudson 42. David Schade 60. Adrian Melott
7. Hans Bohringer 25. Mark Hindmarsh 43. Ken Chambers 61. Gerhard Borner
8. Donald Lynden-Bell 26. Bernard Carr 44. Richard Bowers 62. Ewan Stewart
9. Margaret Norman 27. Yannick Mellier 45. Satoshi Yoshioka 63. Andrew Liddle
10. Fred Hoyle 28. Yehuda Hoffmann 46. Nathan Roche 64. Stephen Hancock
11. Arnold Wolfendale 29. Francesco Melchiorri 47. Avishai Dekel 65. Alain Blanchard
12. Haydek'Sirousse-Zia 30. Ray Sharples 48. Ian Smail 66. Dominic Lefebvre
13. Kamilla Piotrkowska 31. Gordon Stewart 49. John Peacock 67. Yuri Parijskij
14. John Huchra 32. Nial Tanvir 50. Dick Fong 68. Fablen Moutarde
15. Mike Feast 33. Alvio Renzini 51. Ben Moore 69.
16. Tom Shanks 34. Ioannis Georgantopoulos 52. Robert Mann 70. Caleb Scharf
17. Tony Banday 35. Carlos Frenk 53. Nigel Metcalfe 71. Renee Kraan-Korteweg
IS. Mike Pierce 36. Ofer Lahav 54. Duncan Hale-Sutton 72. Simon Lilly
73. Rafael Guzman
List of Participants

Aragon, A.S., Physics Department, University of Durham, South Road, Durham DHI
3LE, U.K

Banday, A.J., Physics Department, University of Durham, South Road, Durham DHI
3LE, U.K.

Blanchard, A., DAEC, Observatoire de Meudon, 92195 Meudon Cedex, France.

Bohringer, H., Max Planck-Institut fur Extraterrestrische Physik, Giessenbachstrasse, D-


8046, Garching bei Munchen,West Germany.

Borner, G., Max Planck-Institut fur Physik & Astrophysik, Instiut fur Astrophysik, Karl-
Schwarzschild-Strasse 1, 8046 Garching bei Munchen, West Germany.

Bower, R.G., Physics Department, University of Durham, South Road, Durham DHI
3LE, U.K.

Broadhurst, T.J., Theoretical Astronomy Unit, School of Mathematical Sciences, Queen


Mary & Westfield, Mile End Road, LONDON EI 4NS, U.K

Buryak, 0., Inst. of Applied Maths, Miusskaya Sq. 4, MOSCOW 125047, USSR.

Cannon, R.D., Anglo-Australian Observatory, Epping Laboratory, PO Box296, Epping,


NSW 2121, AUSTRALIA.

Carr, B, J., Theoretical Astronomy Unit, School of Mathematical Sciences, Queen Mary
& Westfield, Mile End Road, LONDON EI 4NS, U.K.

Chambers, KC., Sterrewacht, Postbus 9513, 2300 RA LEIDEN, The NETHERLANDS.

Coles, P., Astronomy Centre, School of Mathematical and Physical Sciences, University
of Sussex, Falmer, Brighton, BNI9QH, UK

Colless, M.M., Institute of Astronomy, Madingley Road, Cambridge CB3 OHA, U.K.

Cowie, L.L., Institute for Astronomy, University of Hawaii, 2680 Woodlawn Drive,
Honolulu, HI 96822, USA.

Cristiani, S., Dipartimento di Astronomia, Universita di Padova, vicolo dell'Osservatorio


5,35122 Padova, ITALY.

Davies, R. D., Nuffield Radio Astronomy Labs., Jodrell bank, Macclesfield, Cheshire
SKI I 9DL, U.K

xv
xvi

Dekel, A., Racah Inst. of Physics, The Hebrew Univ. of Jerusalem, JERUSALEM 91904,
ISRAEL.

Demarque, P., Yale Univ. Observatory, 260 Whitney Avenue, PO Box 6666, NEW
HAVEN, CT06511, USA.

Doroshkevich, A.G., Inst. of Applied Maths, Miusskaya Sq. 4, MOSCOW 125047,


USSR.

Efstathiou, G., Dept of Astrophysics, Univ. of Oxford, South Parks Road, OXFORD OXI
3RQ.

Ellis, R.S., Physics Department, University of Durham, South Road, Durham DHI 3LE,
U.K.

Feast, M.W., South African Astronomical Observatory, PO Box 9, Observatory 7935,


Cape Town, South Africa.

Frenk, C.S., Physics Department, University of Durham, South Road, Durham DHI 3LE,
U.K.

Georgantopoulos, I., Physics Department, University of Durham, South Road, Durham


DHI 3LE, U.K.

Guiderdoni, B., Institut d'Astrophysique, 98 bis, Boulevard Arago, F-70154, PARIS,


France.

Guth, A.H., Centre for Theoretical Physics, Dept of Physics, Massachusetts Inst. of
Technology, Cambridge MA02139, USA.

Guzman, R., Physics Department, University of Durham, South Road, Durham DHI
3LE.

Hanes, D., Queen's University, Astronomy Group, Physics Department, Stirling Hall,
KINGSTON K7L 3N6, CANADA.

Heavens, A.F., Royal Observatory, Blackford Hill, EDINBURGH EH9 3HJ, U.K.

Hindmarsh, M., Physics Department, George's Place, University of Newcastle,


NEWCASTLE UPON TYNE, NEI 7RU, U.K.

Hoffman, Y., TECHNION - Israel Institute of Technology, Dept. of Physics, 32000


Haifa, ISRAEL.

Hoyle, F., 102 Admiral's Way, West Cliff Road, Boumemouth, DORSET, BH2 5HF,
U.K.
xvii

Huchra, J.P., Smithsonian Observatory, Centre for Astrophysics, 60 Garden Street,


CAMBRIDGE, MA02138, USA.

Hudson, M., Institute of Astronomy, Madingley Road, Cambridge CB3 OHA, U.K.

Kaiser, N., CITA, University of Toronto, McLennan Labs., 60 St. George Street,
Toronto, Ontario, M5S IAI, CANADA.

Kraan-Korteweg, R., Astron. Inst. of the Univ. of Basel, Venusstrasse 7, ch-4102,


Binningen, Switzerland.

Labav, 0., Institute of Astronomy, Madingley Road, CAMBRIDGE CB3 OHA, U.K.

Lasenby, A., MRAO, Cavendish Laboratory, University of Cambridge, Madingley Road,


CAMBRIDGE CB3 OHA, U.K.

Lefebvre, D., MRAO, Cavendish Laboratory, University of Cambridge, Madingley Road,


CAMBRIDGE CB3 OHA, U.K.

Liddle, A.R., Astronomy Centre, School of Mathematical and Physical Sciences,


University of Sussex, Falmer, Brighton, BNI9QH, U.K.

Lidsey, J., Room 201, Theoretical Astronomy Unit, School of Mathematical Sciences,
Queen Mary & Westfield, Mile End Road, LONDON El 4NS, U.K.

Lilly, SJ., Institute for Astronomy,University of Hawaii, 2680 Woodlawn Drive,


Honolulu, HI 96822, USA.

Lorrimer, S., Physics Department, University of Durham, South Road, Durham DH!
3LE, U.K.

Lucey, J.R., Physics Department, University of Durham, South Road, Durham DH! 3LE,
U.K.

Lynden-Bell, D., Institute of Astronomy, The Observatories, Madingley Road,


Cambridge CB3 OHA, U.K.

Mann, R., Royal Observatory, Blackford Hill, Edinburgh EH9 3HJ, U.K.

Melchiorri, F., Dip Fisica, Univ. La Sapienza, P.zza Aldo Moro, ROMA, ITALY.

Mellier, Y., Observatoire de Toulouse, 14 ave. Edouard Belin, F31400, TOULOUSE,


FRANCE.

Melon, A.L., Physics & Astronomy, University of Kansas, Lawrence, KS 66045, USA.
xviii

Metcalfe, N., Physics Department, University of Durham, South Road, Durham DHI
3LE, U.K.

Mingaliev, M., Special Astrophysical Observatory, Nizhny Aekhys, Zelenchukskaya,


STAVROPOLOSKYU KRAJ, USSR.

Moore, B., Physics Department, University of Durham, South Road, Durham DHI 3LE,
U.K.

Moss, I., Physics Department, George's Place, University of Newcastle, NEWCASTLE


UPON TYNE, NEI 7RU, U.K.

Moutarde, F., DAEC, Observatoire de Meudon, 5 PI. J. Janssen, 92195 Meudon Cedex,
France.

Oliver, S., Theoretical Astronomy Unit, School of Mathematical Sciences, Queen Mary
& Westfield, Mile End Road, LONDON El 4NS, U.K.

Parijskij, Y., Special Astrophysical Observatory, Nizhny Aekhys, Zelenchukskaya,


STAVROPOLOSKYU KRAJ, USSR.

Peacock, J.A., Royal Observatory, Blackford Hill, EDINBURGH EH9 3HJ, U.K.

Peebles, P.J.E., Physics Dept., Jadwin Hall, Princeton University, P.O. Box 708,
PRINCETON NJ 08544, USA.

Penny, A.J., Rutherford Appleton Laboratory, Chilton, Didcot, OXII OQX, Berks.

Pierce, M.J., Dominion Astrophysical Observatory, 5071 W. Saanich Road, VICTORIA,


B.C. V8X 4M6, CANADA.

Piotrkowska, K., Dept of Astrophysics, Nuclear Physics Building, Univ. of Oxford, South
Parks Road, OXFORD OXI 3RQ, UK

Pritchet, C.J., Dominion Astrophysical Observatory, 5071 W. Saanich Road, VICTORIA,


B.C. V8X 4M6, CANADA.

Rees, M.J., Institute of Astronomy, University of Cambridge, Madingley


Road,Cambridge CB3 OHA, U.K.

Renzini, A., Osservatorio Astronomico di Bologna,Via Zamboni 33, 40126, BOLOGNA,


Italy.

Rowan-Robinson, M., Dept. of Maths & Astronomy, Queen Mary College, Mile End
Road, London El 4NS, U.K.
xix

Sarkar, S., Dept. of Theoretical Physics, University of Oxford, Keble Road, Oxford OXI
3NP, U.K.

Saich, P., Astronomy Centre, School of Mathematical and Physical Sciences, University
of Sussex, Falmer, Brighton, BNI9QH, U.K.

Salopek, D., NASAlFermilab Astrophysics Group, PO Box 500, MS-209, Batavia,


ILLINOIS 6051, USA.

Sanz, J.L., Dpto Fisica Moderna, Univ. de Cantabria, Av. de Los Castros s.n., 39005
SANTANDER, SPAIN.

Schade, D., Institute of Astronomy, Madingley Road, Cambridge CB3 OHA, U.K.

Scharf, C., Institute of Astronomy, Madingley Road, Cambridge CB3 OHA, U.K.

Sciama, D.W., International School of Advanced Studies, Strada Costiera II, 34014
TRIESTE, Italy.

Secco, L., Dipartimento di Astronomia, Universita di Padova, vicolo dell'Osservatorio 5,


35122 Padova, ITALY.

Shanks, T., Physics Department, University of Durham, South Road, Durham DHI 3LE,
U.K.

Sirousse-Zia, H., Institut Henri Poincare, Laboratoire de Physique Theorique, II, Rue
Pierre & Marie Curie, 75231 PARIS Cedex OS, FRANCE.

Smoot, G.F., Space Sciences Lab, Univ. of California, I Cyclotron Road,Berkeley


CA94720,USA.

Starobinskii, A., Landau Insitute for Theoretical Physics, Academy of Sciences of USSR,
GSP-I 117940, Kosygina Str., Moscow V-334, USSR.

Stewart, E., Dept. of Physics, Univ. of Lancaster, Lancaster, LAI 4YB, U.K.

Stewart, G.C., Dept. of Physics & Astronomy, Univ. of Leicester, University Road,
Leicester LEI 7RH, U.K.

Sutherland, W.J., Dept of Astrophysics, Nuclear Physics Building, Univ. of Oxford,


South Parks Road, OXFORD OXI 3RQ, U.K.

Tammann, G.A., Astron. Inst. of the Univ. of Basel, Venusstrasse 7, ch4102 Binningen,
Switzerland.

Tanvir, N.R., Physics Department, University of Durham, South Road, Durham DHI
3LE, U.K.
xx

Taylor, A.N., Theoretical Astronomy Unit, School of Mathematical Sciences, Queen


Mary & Westfield, Mile End Road, LONDON E1 4NS, U.K.

Treumann, R.A., Institut fur Extraterrestriche Physik, Max-Planck-Institut fur Physik


und Astrophysik, 8046 Garching b. Munchen, West Germany.

White, S.D.M., Institute of Astronomy, Madingley Road,Cambridge, CB3 OHA, U.K.

Wolfendale, A.W., Physics Department, University of Durham, South Road, Durham


DH1 3LE, U.K.

Yoshii, Y., National Astronomical Observatory, Mitaka, Tokyo 181, JAPAN.

Yoshioka, S., Department of Physics, Tokyo University of Mercantile Marine, Koto-ku,


Tokyo 135, JAPAN.
FUNDAMENTAL ARGUMENTS FOR INFLATION

ALAN H. GUTH
Center for Theoretical Physics,
Laboratory for Nuclear Science and Department of Physics,
Massachusetts Institute of Technology,
Cambridge, Massachusetts, 0£139
and
Harvard-Smithsonian Center for Astrophysics,
60 Garden Street,
Cambridge, Massachusetts 02138

ABSTRACT. The mechanism of inflation is described, and the fundamental arguments in favor
of inflation are summarized. It is claimed that the inflationary model provides a very plausible
explanation for (1) the large number of particles in the universe, (2) the Hubble expansion, (3) the
large-scale uniformity of the universe, (4) the nearness of the universe to a critical density, and (5)
the absence of magnetic monopoles.

1 IntrC'duction

I try to be modest about my own role in the development of inflation, and I think I am aided
in these efforts by the fact that I have a lot to be modest about. When I began working
on cosmology, essentially all the key ideas of inflation had already been discovered. On the
particle physics side, there had been much investigation of grand unified theories and other
spontaneously broken gauge theories; important properties such as phase transitions, false
vacua, and the decay of false vacua were already understood. On the cosmology side, the
shortcomings of the standard big bang theory had been studied, and the horizon and flatness
problems were both known. My role, then, was mainly just to pull these ideas together
into a coherent model. I see no cause, on the other hand, for any of us to be modest about
the inflationary universe theory itself. It is a dramatic development of modern physics,
providing for the first time a theory that accounts for the origin of essentially all of the
matter and energy in the universe. The model is certainly a major achievement in cosmology
if it is correct, and if it turns out to be wrong, it will be a disappointment to a large number
of people.
In this article I will try to explain the fundamental arguments for inflation. To put it
another way, I will try to explain why many people presently believe in the inflationary
model, even though the astronomers have not yet found enough matter to make no = 1.
Fig. 1 shows the first (and most naive) fundamental argument for inflation. The graph
shows the number of articles per year related to inflation, as tabulated from the SPIRES
database at the Stanford Linear Accelerator Center. It is mainly a particle physics database,
so some of the more astrophysical papers on the subject are probably not represented. In
any case, one sees that the inflationary model has stirred up a lot of interest.

T. Shanks et al. (eds.). Observational Tests of Cosmological Inflation. 1-21.


@ 1991 Kluwer Academic Publishers.
2

250
(j) 200
UJ
-.J
u 150
-
r-
0::: 100
« 50
a
80 81 82 83 84 85 86 87 88 89 90
YEAR
Figure 1: A histogram of articles concerning the inflationary universe model. The graph
shows all articles in the SPIRES database that either refer to any of the three basic in-
flationary universe papers ([Guth, 1981; Linde, 1982; or Albrecht & Steinhardt, 1982]), or
contain the string "inflation" in the title.
3

2 The Mechanism of Inflation

Before explaining the arguments for inflation, however, I will first explain how the infla-
tionary model works.
The mechanism of inflation depends on scalar fields, so I will begin by briefly summa-
rizing the role of scalar fields in particle physics. To begin with, the reader should recognize
that in the context of modern particle physics, all fundamental particles are described by
fields. The best known example is the photon. The classical equations describing the
electromagnetic field were written down in the 1860's, but then in the early 20th century
physicists learned that the underlying laws of nature are quantum and not classical. The
quantization of the electromagnetic field can be carried out in a very straightforward way.
For simplicity one can consider the fields inside a box, letting the size of the box approach
infinity at the end of the calculation. One can then think of the electromagnetic fields inside
the box as a mechanical system. The electromagnetic field is written as a sum of normal
modes, and the coefficients of the normal mode functions can be taken as the dynamical
degrees of freedom of the system. It turns out that each of these coefficients obeys the
equations of a harmonic oscillator, and there is no interaction between the coefficients. The
system is quantized by the same rules that one uses to quantize the hydrogen atom or
the harmonic oscillator, and the result is that each normal mode has evenly spaced energy
levels. In this case, we interpret each energy level above the ground state as the occupation
of the mode by a photon. Thus, the photon is interpreted as the quantized excitation of a
field.
In contemporary particle theory, all elementary particles are described in this way.
There is an electron field to correspond to an electron, a quark field to correspond to a
quark, a neutrino field to correspond to a neutrino, etc. Among the different types of fields,
the simplest is the scalar field- a field that has the same value to any Lorentz observer. The
quantized excitation of a scalar field is a spinless particle. Although spinless particles that
are regarded as elementary have yet to be observed, they are nonetheless a key ingredient
to a number of important theories. In particular, the Glashow-Weinberg-Salam model of
the electroweak interactions makes use of a scalar field, called the Higgs field, to cause a
symmetry in the theory to be spontaneously broken. (If this symmetry were not broken,
then electrons and neutrinos would both be massless, and would be indistinguishable.) This
Higgs field corresponds to a neutral spinless Higgs particle, which will hopefully be observed
at the SSC (Superconducting Super-Collider), if not before. Grand unified theories make
use of similar Higgs fields, but at a much higher mass scale, to spontaneously break the
grand unified symmetry which relates electrons, neutrinos, and quarks.
There is much current interest in superstring theories, which actually go beyond the
pattern of the field theories that were described above. In these theories the fundamental
object is not a field, but is really a string-like object, which has length but no width. These
theories are believed to behave as field theories, however, at energy scales well below the
Planck scale (Mp == 1/.../G = 1.02 X 1019 GeV, where G is Newton's constant, and I use
units for which n == c == 1). In fact, these field theories contain a large number of scalar
fields. Thus, the particle physics motivation for believing that scalar fields exist is quite
strong.
4

.,-- False Vacuum


Pr
( metastable)

<Pr <Pt
Scalar Field

Figure 2: The potential energy function for the scalar field 4>, appropriate for the new
inflationary universe model.

The electromagnetic field has a potential energy density given by

v = J.-(E2
811'
+ ffl) , (1)

but the potential energy density of the scalar field has a wider range of possibilities. It is
restricted by the criterion that the field theory be renormalizable (Le., the requirement that
the field theory leads ultimately to finite answers), but there are still several free parameters
involved in specifying the potential. The form of the scalar field potential appropriate for
the new inflationary universe model is shown in Fig. 2. The potential has a minimum at a
value of the scalar field 4> that is nonzero, a property that is generic for Higgs fields. The
potential also has a plateau centered at 4> = cPr, a feature that Higgs fields do not necessarily
have. This property is necessary, however, in order for the new inflationary scenario to be
possible.
A particle physicist defines the word "vacuum" to mean the state of lowest possible
energy density, so it is the state in which 4> lies at the minimum of the potential, labeled
4>t on the diagram. Notice that the energy density of this state is shown as zero, which is
equivalent to the statement that the cosmological constant is either zero or immeasurably
small. The explanation for this fact has been a long-standing mystery to particle physicists,
although now we have at least a possible solution [Coleman, 1988] in the context of a crude
understanding of quantum gravity. Note that even if the cosmological constant has a
value that is cosmologically significant, the effect on this diagram would be completely
imperceptible. For clarity, I will refer to the vacuum in this situation as the "true vacuum."
5

Before: False Vacuum


{
p = Pr P = -pr

_____
~_4~} {p=o True Vacuum
p=o
After:
dV

........ -----.. F

dW= pdV

Figure 3: A thought experiment to calculate the pressure of the false vacuum. As the piston
chamber filled with false vacuum is enlarged, the energy density remains constant and the
energy increases. The extra energy is supplied by the agent pulling on the piston, which
must pull against the negative pressure of the false vacuum.

The state in which the scalar field, in some region of space, is perched at the top of
the plateau of the potential energy diagram is called the "false vacuum." Note that the
energy density of this state is pc, and is therefore fixed by the laws of physics. The false
vacuum is of course not stable, since eventually the scalar field will no doubt evolve toward
the minimum of the potential. Nonetheless, if the plateau is broad and flat enough, it will
take a long time for the scalar field to roll off the hill. Thus the false vacuum is metastable,
and can be very long-lived, by the time scales of the early universe.
The crucial property of the false vacuum is that the energy density is positive, but on
short time scales it cannot be lowered. This explains the etymology of the name "false
vacuum." Here "false" is being used to mean temporary, and "vacuum" means the state of
lowest possible energy density.
It is now possible to construct a simple energy-conservation argument to determine the
6

pressure of the false vacuum. Imagine a chamber filled with false vacuum, as shown in
Fig. 3. Since the energy density of the vacuum is fixed at Pc, the energy inside the chamber
is given by U = pcV, where V is the volume of the chamber. Now suppose the piston
is pulled outward, increasing the volume by dV. Unlike any normal substance, the false
vacuum will maintain a constant energy density despite the increase in volume. The change
in energy is then dU = pcdV, which must be equal to the work done, dW = pdV, where p
is the pressure. Thus, the pressure of the false vacuum is given by

p = -pc· (2)

If one puts this relation between pressure and energy density into the general form of
the energy-momentum tensor, one finds

(3)
where gl''' = diag[-l, 1, 1, 1] is the Lorentz metric. Thus, the energy-momentum tensor
of the false vacuum is Lorentz-invariant, as one would expect, since the state is Lorentz-
invariant. That is, if a region of space has a scalar field with a constant value <Pc, to another
Lorentz observer it will also look like a region of space with a scalar field of value <Pc. One
could in fact have used this Lorentz-invariance as an alternative derivation of Eq. (2).
Finally, we are ready to discuss the gravitational effects of this very peculiar energy-
momentum tensor. Starting with the Einstein field equations

'Ill''' - !gl''''Il = 81rGTI''' - Agl''' , (4)


2
one sees that the effect of the energy-momentum tensor of Eq. (3) is precisely the same as
having a temporarily nonzero value of the cosmological constant A, related by

A = 81rGpc . (5)

To see the effect on a Robertson-Walker universe, note that the scale factor R( t) obeys the
equation of motion
.. 41r
R = -3"G(p + 3p)R. (6)
In the present universe, the pressure term is a small relativistic correction. If the universe
was ever dominated by false vacuum, however, then the pressure term has the opposite
sign, and overcomes the gravitational attraction caused by the energy density term:
.. 81r
R= -GpcR. (7)
3
The force of gravity actually becomes repulsive, and the expansion rate of the universe is
accelerated. The general solution to Eq. (7) is
(8)

where CI and C2 are arbitrary constants, and the exponential rate is given by

(9)
7

After some time the growing term will dominate, and the expansion becomes a pure expo-
nential.
While the new inflationary scenario proposed by Linde [1982] and Albrecht & Steinhardt
[1982] assumes a potential energy function of the form shown in Fig. 2, Linde [1983a, 1983b]
has shown that such a severe restriction is not necessary. In a model known as chaotic
inflation, Linde showed that inflation can work for a scalar field potential as simple as
V(4)) = >'4>4, provided that one makes some assumptions about the initial conditions. He
proposed that the scalar field begins in a chaotic state, so that there are some regions in
which the value of 4> is a few times larger than the Planck mass Mp. These regions must
exceed some minimal size, which is estimated to be several times H-l, where H denotes
the Hubble constant. Then 4> rolls down the hill of the potential energy diagram, and a
straightforward calculation indicates that there is an adequate amount of inflation. The
Hubble "constant" is not a constant in this case, but it is slowly varying, so the expansion
can be called "quasi-exponential".

3 The Inflationary Universe Scenario

The inflationary universe scenario begins with a patch of the universe somehow settling
into a false vacuum state. The mechanism by which this happens has no influence on the
later evolution. The following three possibilities have been discussed in the literature:

1. Supercooling from high temperatures. This was the earliest suggestion [Guth, 1981;
Linde, 1982; Albrecht & Steinhardt, 1982]. If we assume that the universe began very
hot, as is traditionally assumed in the standard big bang model, then as the universe
cooled it presumably went through a number of phase transitions. For many types
of scalar field potentials, supercooling into a false vacuum occurs naturally. This sce-
nario has the difficulty, however, that there is no known mechanism to achieve the
desired pre-inflationary thermal equilibrium state. Calculations [Starobinsky, 1982;
Guth & Pi, 1982; Hawking, 1982; Bardeen, Steinhardt, & Turner, 1983] show that
the scalar field must be very weakly coupled in order for quantum-induced density
perturbations to be sufficiently small, and consequently the scalar field would require
much more than the available time to relax to thermal equilibrium. It has been
shown, however, that true thermal equilibrium is not really necessary: a variety of
random configurations give results that are very similar to those of thermal equi-
librium [Albrecht, Brandenberger, & Matzner, 1985; Kung & Brandenberger, 1990;
Brandenberger, 1991; Goldwirth, 1991].

2. Tunneling from "nothing". These ideas are of course very speculative, since they
involve a theory of quantum gravity that does not actually exist. The basic idea,
however, seems very plausible. If geometry is to be described by quantum theory,
then the geometry of space can presumably undergo quantum transitions. One can
then imagine an initial state of absolute nothingness- the absence of matter, energy,
space, or time. The state of absolute nothingness can presumably undergo a quantum
transition to a small universe, which then forms the initial state for an inflationary
8

scenario. Variations of these ideas have been studied by Tryon [1973], Vilenkin [1982,
1985], Linde [1983c, 1984a, 1984b], and Hartle & Hawking [1983].

3. Random fluctuations in chaotic cosmology. In Linde's [1983a, 1983b] chaotic cosmol-


ogy, it is assumed that the scalar field ~ begins in a random state in which all possible
values of ~ occur. Inflation then takes place in those regions that have appropriate
values of ~, and these inflated regions dominate the universe at later times. In these
models it is not necessary for the scalar field potential energy function V( ~) to have
a plateau, but as in other models it must be very flat (i.e., weakly self-coupled) in
order to minimize the density perturbations that result from quantum fluctuations.

Regardless of which of the above mechanisms is assumed, one expects that the correlation
length of the scalar field just before inflation is of the order of the age of the universe at the
time. For typical grand unified theory parameters, this gives a correlation length of about
10- 24 cm.
The patch then expands exponentially due to the gravitational repulsion of the false
vacuum. In order to achieve the goals of inflation, we must assume that this exponential
expansion results in an expansion factor ~ 10 25 • For typical grand unified theory numbers,
this enormous expansion requires only about 10-32 sec of inflation. During this inflationary
period, the density of any particles that may have been present before inflation is diluted
so much that it becomes completely negligible. At the same time, any nonuniformities in
the metric of space are smoothed by the enormous expansion. The explanation for this
smoothness is identical to the reason why the surface of the earth appears to be flat, even
though the earth is actually round- any differentiable curve looks like a straight line if one
magnifies it enough and looks at only a small segment. The correlation length for the scalar
field is stretched by the expansion factor to become at least about 10 cm. If the duration of
inflation is more than the minimal value, which seems quite likely, then the final correlation
length could be many orders of magnitude larger. There appears to be no upper limit to
the amount of inflation that may have taken place.
The false vacuum is not stable, so it eventually decays. If the decay occurs by the usual
Coleman-Callan process [Coleman, 1977; Callan & Coleman, 1977] of bubble nucleation,
then the randomness of the bubble nucleation process would produce gross inhomogeneities
in the mass density [Hawking, Moss, & Stewart, 1982; Guth & Weinberg, 1983]. This
problem is avoided in the new inflationary scenario [Linde, 1982; Albrecht & Steinhardt,
1982] by introducing a scalar field potential with a flat plateau, as was shown in Fig. 1. This
leads to a "slow-rollover" phase transition, in which quantum fluctuations destabilize the
false vacuum, starting the scalar field to roll down the hill of the potential energy diagram.
These fluctuations are initially correlated only over a microscopic region, but the additional
inflation that takes place during the rolling can stretch such a region to be large enough to
easily encompass the observed universe.
When the phase transition takes place, the energy that has been stored in the false
vacuum is released in the form of new particles. These new particles rapidly come to
thermal equilibrium, resulting in a temperature with kT ~ 1014 GeV. At this point the
scenario rejoins the standard cosmological model.
9

The baryons are produced [see, for example, Kolb & Turner, 1983] by baryon noncon-
serving processes after inflation. Any baryons that may have been present before inflation
are simply diluted away by the enormous expansion factor. Thus, inflationary cosmology
requires an underlying particle theory, such as a grand unified theory, in which baryon
number is not conserved.
It is a dramatic feature of inflationary models that essentially all of the matter, energy,
and entropy of the observed universe is produced by the expansion and subsequent decay
of the false vacuum. (I used the qualifier "essentially" to acknowledge the fact that a
small patch of false vacuum is necessary to start inflation. For typical grand unified theory
parameters, with a mass scale of order 1014 GeV, the required volume of radius rv H-l has
a mass of order 10 kg.) For this reason I sometimes refer to the process of inflation as the
ultimate free lunch.

4 The Eternal Nature of Inflation

A fascinating feature of inflation, which in my opinion is also important in evaluating the


plausibility of inflation, is the fact that inflation is eternal- if inflation ever begins, then
it will never stop [Vilenkin, 1983; Starobinsky, 1982, 1986; Linde, 1986, 1987; Aryal &
Vilenkin, 1987].
To understand the endlessness of inflation, one first notices that the decay of the false
vacuum, like the decay of many other unstable systems, is an exponential process. S.-Y.
Pi and I [Guth & Pi, 1985] have verified the exponential decay law in a simplified but
exactly soluble model of a slow-rollover phase transition, in which the potential is taken
as V(</» = _!p.2</>2. For the case of chaotic inflation, on the other hand, one might think
that the scalar field would roll inexorably down the hill in the potential energy diagram,
completing the decay in a finite time. Linde [1986] has shown, however, that if the scalar
field starts at a sufficiently high value, then it can be sustained by quantum fluctuations,
with again an exponential decay law. As long as the false vacuum endures it drives an
exponential expansion, and for reasonable parameters the rate of expansion is much faster
than the rate of decay. Thus, even though the false vacuum is decaying, the total volume
of the false vacuum region actually increases with time.
As time goes on, pieces of the false vacuum region are constantly undergoing decay. As
each piece decays, it releases energy and thereby sets into motion a hot big bang universe.
Other regions of false vacuum, however, continue to exponentially expand, so the false
vacuum never disappears.
The infinity of universes produced in this way rapidly become causally disconnected, so
there is no way for us to verify, even in principle, that the other universes exist. Nonetheless,
I feel that the eternal character of inflation makes it a more plausible theory. In the
absence of this feature, there is some difficulty in deciding whether the initial conditions
required for inflation are sufficiently plausible. Since there is no established theory of
initial conditions, questions of this sort can easily lead to inconclusive answers. Given
the endlessness of inflation, however, the question becomes much less significant. Just as
most of us accept the claim that complicated DNA molecules originated through random
processes sometime during the history of the earth, we can also accept the claim that a
10

region of false vacuum originated through random processes sometime during the history
of spacetime. Just as primitive DNA molecules increased their abundance by replication,
one patch of false vacuum would inflate to produce an infinity of universes, one of which
could be the universe in which we live.

5 Evidence for Inflation

5.1 BIGNESS OF THE UNIVERSE

To most students of cosmology, one of the most startling features of the universe is its
incredible size-- the observable universe contains approximately 1090 particles. Since the
standard big bang model (without inflation) contains no mechanism to produce such a huge
amount of entropy, the model requires us to assume that essentially all of these particles
were here at the start.
The inflationary model, on the other hand, can actually explain where such a vast
number of particles can come from. Particles are produced by the expansion and decay of
the false vacuum. Since the expansion is exponential, it makes sense to write 1090 = e201 =
(e 69 )3. Thus, an exponential expansion of 69 e-foldings is sufficient to turn a single particle
into 1090 particles. Inflation therefore reduces the problem of explaining the 1090 particles
to the problem of explaining why there were more than 69 e-foldings of inflation. In fact it
is easy to construct underlying particle theories that will give far more than 69 e-foldings
of inflation. The suggestion is that even though the observed universe is incredibly large,
it is only an infinitesimal fraction of the entire universe.

5.2 HUBBLE EXPANSION

Although the standard cosmological model is called the big bang, the theory in fact contains
no description whatever of the "bang". It is really a theory of the aftermath of a bang,
describing how the matter expands and cools, coagulating to form galaxies and other visible
structures in the universe.
With inflation, however, we have for the first time a theory of the bang itself- the
outward thrust of the big bang can be attributed to the repulsive gravity of the false vacuum.
As we noticed in discussing Eq. (7), the false vacuum leads to a gravitational repulsion that
in turn leads to exponential expansion. This uniform expansion agrees precisely with the
pattern of expansion discovered by Hubble.

5.3 HOMOGENEITY/ISOTROPY

My third argument for inflation is known as the homogeneity or isotropy problem, and it
is also called the horizon problem. The problem is that the extreme uniformity that is
observed on very large scales, particularly in the cosmic background radiation, cannot be
explained without inflation.
Observationally, the effective temperature of the cosmic background radiation is known
to be isotropic to about one part in 103 , and even this anisotropy can be accounted for
11

by the assumption that the earth is moving through the background radiation. If one
removes the dipole component that can be attributed to the earth's motion, then the
residual anisotropy is known to be less than one part in 104 [Wilkinson, 1986]. The limits
on anisotropies continue to improve, and the latest results will be summarized later in this
conference by G. Smoot and F. Melchiorri. The extreme uniformity in the observations is
very difficult to understand in the context of the standard cosmological model, in which the
horizon distance (i.e., the distance that a light pulse could have traveled since the initial
singularity) is rather short. The existence of horizons in cosmology was first discussed by
Rindler [1956], and the horizon problem was discussed by Weinberg [1972] and by Misner,
Thorne, & Wheeler [1973].
Consider, for example, two microwave antennae pointing in opposite directions. Each
is receiving radiation that is believed to have been emitted (or last scattered) at the time
of hydrogen recombination, t r , about 105 years after the big bang, when the temperature
Tr was about 4000oK. (At earlier times the plasma that filled the universe was opaque to
this radiation.) At the time of emission, these two sources were separated from each other
by many horizon distances. To estimate how many, let us assume that n ~ 1, and that the
universe can be approximated as being matter-dominated during the relevant time period.
Then the Robertson-Walker scale factor is given by

R(t) = bt 2/ 3 , (10)

for some constant b. The scale factor specifies the ratio of a physical distance to a coordinate
distance. Since the physical speed of light in our units is one, the coordinate speed of light
is 1/ R(t). The coordinate distance between the source of the cosmic background radiation
and us is then given by

rcoord -
_l tr
to dt'
bt,2/3 = 3 b- 1( to1/3 - 1/3)
tr , (11)

where to denotes the present time. The coordinate value of the horizon distance at time tr
is given by
_t
lhor,coord - 10
r ~ _
bt12 / 3 - 3b tr
-11/3
. (12)

Thus, the number of horizon distances separating the two sources in opposite directions is
given by
N = 2rcoord
lhor,coord
=2 [(to)1/3
tr
-1] = 2 [(Tr)1/2
To
-1] ~ 75, (13)

where we used the fact that RT ~ const and To ~ 2.74°K. The problem is to understand
how two regions over 75 horizon distances apart came to be at the same temperature at the
same time. (In the above discussion we have assumed for simplicity that n = 1 and that
the universe was completely matter-dominated. More detailed calculations [Guth, 1983a],
however, show that the problem is if anything a little worse- the two regions were at least
90 horizon distances apart.)
The horizon problem is not an inconsistency in the standard model, but represents in-
stead a lack of explanatory power. If the universe is assumed to have begun homogeneously,
12

then it will continue to evolve homogeneously. The problem is that the very striking large-
scale homogeneity of the universe is not explained or predicted by the model, but instead
must simply be assumed.
In the inflationary model, on the other hand, the horizon distance is stretched dur-
ing inflation by the overall inflationary expansion factor, which we denote by Z. At all
subsequent times the entire observed universe is much smaller than the horizon distance.
Furthermore, the inflationary model does more than simply enlarge the horizon distance-
it actually provides a mechanism to create the observed large-scale homogeneity. In the
inflationary model the size of the observable universe at times before the GUT phase tran-
sition was smaller than it would have been in the standard scenario by a factor of Z. If
Z ~ 1025 , then the entire observable universe would have been within its horizon before
inflation; it would have become homogeneous at this time by normal thermal processes,
and then this very small homogeneous region would have been stretched by inflation to
become large enough to encompass the observed universe. The region would then remain
homogeneous as it continued to evolve.

5.4 FLATNESS

Inflation can also solve the "flatness" problem, which was first highlighted by Dicke &
Peebles [1979]. That is, inflation can explain why the mass density of the early universe
was so close to the critical value.
The critical mass density Pc is defined as that mass density which is just barely sufficient
to eventually halt the expansion of the universe. Today the crucial ratio n == pI Pc (where
p is the mass density of the universe) is known to lie in the range

n ~ 2.
0.1 ~ (14)
Despite the breadth of this range, the value of n at early times is highly constrained, since
n = 1 is an unstable equilibrium point of the standard model evolution. Thus, if n is every
e:r:actiy equal to one, it will remain exactly equal to one forever. If, on the other hand, n
is slightly greater than one in the early universe, then it will rapidly rise toward infinity. If
it is slightly less than one, it will rapidly fall toward zero. In particular, it can be shown
that n - 1 grows as
n _1
()(
{tt 2/ 3
(during radiation-dominated era)
(during matter-dominated era) .
(15)

At t = 1 sec, for example, which was the beginning of the processes of big bang nucleosyn-
thesis, n must have been equal to one to an accuracy of one part in 1015. If we extrapolate
further to t = 10-35 sec, the typical time scale for grand unified theory cosmology, we find
n
that must have been equal to one to an accuracy of one part in 1049 . Standard cosmology
provides no explanation for this fact- it is simply assumed as part of the initial conditions.
In the inflationary model, however, n is driven during the period of inflation very rapidly
toward one, as
n - 1 ()( e- 2Ht , (16)
where H is the Hubble constant during inflation. Thus, in an inflationary model one can
begin with a value of n far from unity, and inflation will drive n toward one with spectacular
13

swiftness. This mechanism is so effective, in fact, that it is expected to overshoot by a wide


margin. This leads to the cleanest prediction of inflation, which is that even today the value
of n should differ from one by no more than about one part in 104 • This discrepancy from
one is a quantum effect, comprising the long wavelength tail of the density fluctuations that
are possibly responsible for galaxy formation. The magnitude of these fluctuations is not
fixed by theory, since we do not presently know enough about the parameters of particle
physics models to calculate this number from first principles. Thus the magnitude must
be estimated from observation, using the scale-invariant spectrum predicted by inflation to
extrapolate from much shorter wavelengths.
So far I have described the flatness problem under the assumption that the cosmological
constant A is identically zero. If it is nonzero, then the role of inflation can be described
by specifying that it drives the universe to a state of geometric flatness, corresponding to

A
n + 3H2 = 1. (17)

It is useful to regard the quantity on the left-hand side of this equation as neff, with the
A/(3H2) term regarded as the vacuum energy contribution. Using this definition, inflation
always drives the universe to neff = 1.

5.5 ABSENCE OF MAGNETIC MONOPOLES

Finally, the inflationary model can cure the "magnetic monopole problem." In the context
of grand unified theories, cosmologies without inflation generally lead to huge excesses of
superheavy ('" 1016 GeV) 't Hooft-Polyakov ['t Hooft, 1974; Polyakov, 1974; for a review,
see Coleman, 1983] magnetic monopoles. These monopoles are produced at the grand
unified theory phase transition, when the GUT Higgs fields acquire their nonzero values.
The rapidity of the phase transition implies that the correlation length of the Higgs fields
is very short, and the fields therefore [Kibble, 1976] become tangled in a high density of
knots- these knots are the magnetic monopoles. For typical grand unified theories the
expected mass density of these magnetic monopoles would exceed [Zel'dovich & Khlopov,
1978; Preskill, 1979; Guth & Tye, 1980] the mass density of everything else by a factor
of about 1012 • A value this high would imply that the expansion rate of the universe
would slow to its present value in only 30,000 years, which is rather clearly excluded by
observation.
The monopole problem is easily solved in the context of an inflationary model by ar-
ranging for the Higgs field to acquire its nonzero expectation value either before or during
the inflationary era. The monopole density would then become negligible as it is diluted
by the inflation. Some monopoles would still have been produced by thermal fluctuations
after reheating, but this production is suppressed by a large Boltzmann factor- it would
be negligible in the minimal SU(5) model [Lazarides, Shafi, & Trower, 1982; Guth, 1983b],
and presumably in most other models as well.
14

6 Conclusion

In evaluating the plausibility of a theory, it is always reasonable to ask if there are other
theories that might provide alternative explanations of the same phenomena. In the case
of inflation, I think it is fair to say that there are no direct competitor theories. For the
most part, we are choosing between inflation and the possibility that these questions are
not ready to be answered. Nonetheless, I would like to comment on two alternative points
of view that I sometimes hear mentioned.
The first pertains to the flatness problem in particular. Some people have said that
perhaps we are simply living in a flat universe, and that the flatness should be accepted as
a fundamental law in its own right, not to be explained in terms of other laws. In response
to this suggestion there are two points that I like to emphasize. First, one must recognize
that the closed, open, and flat cases of the Robertson-Walker metric should not be viewed
as being of equal a priori probability. The Robertson-Walker metric is in fact well-defined
for any real value of the parameter k, and it is only by rescaling the coordinates and the
scale factor that one achieves the three standard cases of k = +1, -1 and O. Thus the three
cases really correspond to values of k that are positive, negative, or precisely zero. When
phrased in this manner, the special case k = 0 sounds like it should have a probability
of zero. Nonetheless, the laws of physics are not usually thought to be random, so this
probability estimate is not a convincing answer to someone who claims that k = 0 has the
status of a new law of physics. To such a person I would argue that the proposed law of
physics is peculiarly vague. The ratio

(18)

is a quantity that must be defined by averaging over some region of space, and in fact
we know that because of density fluctuations it will certainly not in general be equal to
one. The most that one can hope for is that it will tend to one as the volume used for
averaging approaches infinity, but such a statement is much more vague than any of the laws
of physics that we normally discuss. In particular, this statement could never be falsified
by measurements made in any finite volume, no matter how accurate those measurements
were.
Secondly, I would like to comment on the anthropic principle, which is sometimes offered
as an explanation of flatness and perhaps some other properties ofthe universe. This prin-
ciple holds roughly that some of the properties of the universe must be the way they are, or
else intelligent life would never have evolved to observe it. To me, however, anthropic argu-
ments are unsatisfactory for two reasons. First, they presume a knowledge, which I think
is lacking, of the minimal conditions necessary for intelligent life. Anthropic arguments are
therefore convenient to "explain" what is known, but I am not aware of any predictions.
Second, even if we had a complete understanding of life and its evolution, I would say that
anthropic arguments entail the risk of bypassing the important questions. For example, I
would guess that some properties of the water molecule are necessary to support life, so one
might say that these properties are "explained" by the anthropic principle. Yet we would
all agree that these properties are also dictated by the laws of quantum theory, and most
of us would agree that the quantum approach is the more productive.
15

Most of our fellow scientists have avoided the anthropic principle completely. I am quite
sure that I have never seen a book on anthropic chemistry, that I have never seen a book
on anthropic nuclear physics, and that Jackson's book (even the newer and thicker edition)
does not contain a chapter on the implications of the anthropic principle for electromagnetic
theory. If used in biology, the anthropic principle would allow scientists to sidestep all
questions about the evolution of life by arguing that otherwise we would not be here to
ask the question- but I doubt very much that this argument would be acceptable to most
biologists. If we are to pursue cosmology with the same standards and methods of argument
that are used in the other sciences, then the anthropic principle should be avoided whenever
possible.
To summarize, I think it is clear that the inflationary universe model explains many
of the most salient features of the observed universe. It explains (1) why the universe
contains a huge number of particles, (2) why the universe is undergoing Hubble expansion,
(3) why the cosmic background radiation is so isotropic, (4) why the early universe had a
mass density so extraordinarily close to the critical value, and (5) why the universe is not
filled with magnetic monopoles. Of these arguments, I find items (3) and (4) to be the
most persuasive, because of the precision involved. The isotropy of the cosmic background
radiation has an accuracy better than one part in 104 , and the value of n at t = 1 sec is
measured indirectly to an accuracy of 15 decimal places. Inflation also makes predictions:
it predicts the value of n, and it makes a slightly less rigid prediction for the shape of
the spectrum of the primordial density perturbations. It is still unclear whether these
predictions will be borne out by observation, but we can look forward to hearing the latest
results on these questions at this conference.

Acknowledgements. This work was supported in part by funds provided by the U. S. Department
of Energy (D.O.E.) under contract #DE-AC02-76ER03069.

R.eferences

Albrecht, A., Brandenberger, R., & Matzner, R., 1985. Phys. Rev. D, 32, 1280.

Albrecht, A. & Steinhardt, P. J., 1982. Phys. Rev. Lett., 48, 1220.

Aryal, M. & A. Vilenkin, 1987. Phys. Lett., 199B, 351.

Bardeen, J. M., Steinhardt, P. J., & Thrner, M. S., 1983. Phys. Rev. D, 28, 679.

Brandenberger, R. H., Feldman, H., & Kung, J., 1991. To be published in the Proceedings of Nobel
Symposium 79: The Birth and Early Evolution of Our Universe, Graftii.vallen, Sweden, June 11-16,
1990, (Physica Scripta).
16

Callan, C.G. & Coleman, S., 1977. Phys. Rev. D, 16, 1762.

Coleman, S., 1977. Phys. Rev. D, 15, 2929.

Coleman, S., 1983. in The Unity of Fundamental Interactions, p. 21, ed. A. Zichichi, Plenum Press.

Coleman, S., 1988. Nucl. Phys., B310, 643.

Dicke, R. H. & Peebles, P.J .E., 1979. in General Relativity: An Einstein Centenary Survey, p. 504,
ed. S.W. Hawking & W. Israel, Cambridge University Press.

Goldwirth, D. S., 1991. "On Inhomogeneous Initial Conditions for Inflation," to be published in
Phys. Rev. D.

Guth, A. H., 1981. Phys. Rev. D, 23, 347.

Guth, A. H., 1983a. in Asymptotic Realms of Physics: Essays in Honor of Francis E. Low, p. 199,
ed. A.H. Guth, K. Huang, & R.L. Jaffe, MIT Press.

Guth, A. H., 1983b. in Magnetic Monopoles, p. 81, ed. R.A. Carrigan, Jr. and W.P. Trower,
Plenum Press.

Guth, A. H. & Pi, S.-Y., 1982. Phys. Rev. Lett., 49,1110.

Guth, A. H. & Pi, S.-Y., 1985. Phys. Rev. D, 32, 1899.

Guth, A. H. & Tye, S.-H., 1980. Phys. Rev. Lett., 44, 631 [Erratum: Phys. Rev. Lett., 44, 963,
1980).

Guth, A. H., & Weinberg, E. J., 1983. Nucl. Phys., B212, 321.

Hartle, J. B. & Hawking, S. W., 1983. Phys. Rev. D, 28, 2960.

Hawking, S. W., 1982. Phys. Lett., U5B, 295.

Hawking, S. W., Moss, I. G., & Stewart, J. M., 1982. Phys. Rev. D, 26, 2681.

Kibble, T.W.B., 1976. J. Phys., A9, 1387.

Kolb, E. W. & Turner, M. S., 1983. Ann. Rev. Nucl. Part. Sci., 33, 645.
17

Kung, J. & Brandenberger, R. H., 1990. Phys. Rev. D, 42, 1008.

Lazarides, G., Shafi, Q., & Trower, W. P., 1982. Phys. Rev. Lett., 49, 1756.

Linde, A. D., 1982. Phys. Lett., 108B, 389.

Linde, A. D., 1983a. Zh. Eksp. Teor. Fiz., 38,149 [JETP Lett., 38,176].

Linde, A. D., 1983b. Phys. Lett., 129B, 177.

Linde, A. D., 1983c. in The Very Early Universe, p. 205, ed. G.W. Gibbons, S.W. Hawking, and
S.T.C. Siklos, Cambridge University Press.

Linde, A. D., 1984a. Nuovo Cimento Lett., 39, 401.

Linde, A. D., 1984b. Zh. Eksp. Teor. Fiz., 40, 496 [JETP Lett., 40, 1333].

Linde, A. D., 1986. Mod. Phys. Lett., AI, 81.

Linde, A. D., 1987. Int. J. of Mod. Phys., A2, 561.

Misner, C. W., Thorne, K. S., & Wheeler, J. A., 1973. Gravitation, Freeman, San Francisco, pp.
740 and 815.

Polyakov, A. M., 1974. Pis'ma Zh. Eksp. Teor. Fiz., 20,430 [JETP Lett., 20, 194].

Preskill, J.P., 1979. Phys. Rev. Lett., 43, 1365.

Rindler, W., 1956. Mon. Not. R. astr. Soc, 116,663.

Starobinsky, A. A., 1982. Phys. Lett., 117B, 175.

Starobinsky, A. A., 1986. in Field Theory, Quantum Gravity and Strings, ed. M.J. de Vega & N.
Sanchez, Lecture Notes in Physics, Springer Verlag, 246, 107.

't Hooft, G., 1974. Nuc\. Phys., B79, 276.

Tryon, E. P., 1973. Nature, 246, 396.

Vilenkin, A., 1982. Phys. Lett., 117B, 25.


18

Vilenkin, A., 1983. Phys. Rev. D, 27, 2848.

Vilenkin, A., 1985. Nucl. Phys., B252, 141.

Weinberg, S., 1972. Gravitation and Cosmology, Wiley, New York, pp. 525-526.

Wilkinson, D., 1986. in Inner Space I Outer Space: The Interface between Cosmology and Particle
Physics, p. 126, ed. E.W. Kolb, M.S. Turner, D. Lindley, K. Olive, & D. Seckel, Chicago University
Press.

Zel'dovich, Ya. B. & Khlopov, M. Y., 1978. Phys. Lett., 79B, 239.

DISCUSSION:

Borner: Are there possible observations that could disprove the inflationary idea?

Guth: In general, I think ideas die when they are replaced by a better idea. It's very rare
that an idea is given up in exchange for absolutely nothing. In the case of inflation, I would
say that if observations could convincingly show that n + (A/3H2) is not equal to one, then
I think the theory would have to be given up. The density fluctuations, on the other hand,
are not so clear-cut. If the density fluctuations were shown to have a primordial spectrum
that is not Harrison-Zeldovich, then there are ways to modify the theory. I would still say,
however, that measurements of the density fluctuations are crucial. If inflation is correct,
they will help us to refine the theory. If inflation turns out not to be correct, then these
measurements may help to point the way to the new theory that will replace it.

Rowan-Robinson: Surely inflation cannot predict a universe like the one we observe
from arbitrary initial conditions? Specifically, if the inflation factor is Z, inflation predicts
uniformity and flatness only if the initial inhomogeneity and deviation from flatness are
< Z. In a sense, even though Z may be a large number, this still makes our universe a
member of a set of measure zero. (George Ellis' argument).

Guth: Yes, inflation does not work for arbitrary initial conditions. For example, one can
always start with a closed universe with all quantities of Planck scale, and such a universe
will collapse on the order of a Planck time, without ever reaching the GUT scale for inflation
to begin. However, I think it is the wrong question to ask whether inflation can work for
arbitrary initial conditions. Inflation is a theory of evolution, and I think it should be
judged by the same criteria that are used for other theories of evolution. In particular,
I have never heard anyone criticize Darwin's theory of the evolution on the grounds that
it doesn't work for arbitrary initial conditions. Inflation is not a complete theory of the
universe, but simply tries to describe the evolution of the universe from about 10-35 sec
19

to maybe 10-30 sec. So inflation depends on having some other theory of creation, which
will describe how the initial conditions for inflation could have been achieved. I think
inflation should be judged on whether it gives a plausible explanation of how the universe
evolved to where it is now, starting from initial conditions that seem reasonable- more
precisely, initial conditions that can hopefully be explained later by some theory of creation,
or perhaps some theory that merely describes the period immediately before inflation.
Your question seems to refer specifically to the idea that space may have a fractal
structure, with inhomogeneities persisting down to arbitrarily small length scales. We have
to keep in mind, however, that the universe is quantum mechanical. This means that to
have an excitation of wavelength A requires an energy of order lie/A. Thus it would require
an infinite energy density to have excitations down to arbitrarily small wavelengths, so I
think it's reasonable to assume that this is impossible. As long as the energy density is
finite, there must be some wavelength below which there are no excitations at all. There
are, however, also quantum zero point fluctuations, and the story about them is a bit more
complicated. These zero point fluctuations do persist down to arbitrarily small wavelength,
at least in the context of a quantum field theory, but the fluctuations are not arbitrary-
they are completely controlled by the theory. When one quantizes a scalar field in de Sitter
space, one finds that the zero point fluctuations of extremely short wavelength at one time
evolve to become the long wavelength zero point fluctuations at a later time. Furthermore,
this simple evolution is dictated by the fact that the zero point fluctuations are de Sitter
invariant. For the real theory, we in principle have to deal with the zero point fluctuations
of quantum gravity on very short wavelengths, and here we of course have no rellable theory.
Nonetheless, it seems reasonable to hope that these zero point fluctuations are also de Sitter
invariant, since the classical theory is, and then the process of inflation will stretch these
perturbations to become the usual de Sitter invariant zero point fluctuations on longer
wavelengths- which are completely tame.

Rowan-Robinson: You need an anthropic argument to explain why we do not find


ourselves in a region of false vacuum today.

Guth: Yes, I guess I do. But I guess that's the only application of the anthropic principle
that I am happy with. I would say that it avoids each of the two objections to anthropic
arguments that I mentioned in my talk. It does not rely on any knowledge of what it takes
to produce intelligent life, except for the rather safe assumption that life cannot form in
the inflating false vacuum. Furthermore the properties of the universe (its constituents, its
mass density, density fluctuations, everything) are still determined by the mechanism of
inflation, so this anthropic argument is not an attempt to bypass such an explanation.

Smoot: What level of isotropy and homogeneity is needed for inflation to begin? Then
can you estimate the probability of inflation starting by comparing the needed conditions
to the Planck scale?

Guth: It's hard to quantify, but one can solve for the evolution of a sphere of false
vacuum surrounded by true vacuum. If the sphere is larger than one horizon volume,
inflation will begin. So we assume that the necessary condition for inflation is approximate
homogeneity and isotropy on the scale of the horizon. If we assume that the universe
20

survives long enough to reach the energy scale of the false vacuum, then the horizon length
is the typical length scale, and inflation looks plausible. A detailed calculation depends,
however, on assumptions about the initial conditions, and different authors have reached
different conclusions.
There is the associated question of whether the universe ever reaches the false vacuum
energy scale. IT the patch of universe under discussion resembles an open universe, then it
will expand forever and there is no problem. IT it is closed, however, then some mechanism
is needed to make it plausible for the patch to survive long enough- one possibility is
a Planck-scale episode of inflation, which can be totally independent of the inflation to
occur later. (The most recent episode of inflation cannot, however, be at the Planck scale,
because too much gravitational radiation would be produced.) Another possibility is the
creation of the universe by quantum processes, with a quantum transition directly into the
false vacuum state.

Rees: The mass within the horizon of a "true-vacuum" Einstein-de Sitter region increases
linearly with time. So will our descendents eventually see the edge of our domain?

Guth: Yes, in principle our ancestors will at some time in the distant future be able to
see the edge of our domain. It is possible even that the domain is closed, in which case the
wall will hit them in the face. The time for this to happen, however, is proportional to the
cube of the overall expansion factor, and so it will not happen until the extremely distant
future.

Sarkar: You suggested that the entropy of the universe is large. How would you respond
to the remark by Roger Penrose that the observed entropy ('" 10 10 per baryon) is actually
completely negligible relative to the natural value of the gravitational entropy ('" 1040 per
baryon)? Does inflation shed any light on the very special initial conditions which seem to
be required for the gravitational field? In other words, how did the universe survive from
the Planck scale down to the energy scale where inflation occurred.

Guth: The comment of Penrose that you cite is quite old, and was made in the days when
people assumed that baryon number was exactly conserved. From that point of view the
baryon number of the visible universe, about lOBO, was assumed to be present from the
start. It was in this context that Penrose pointed out that the entropy of these baryons
would have been much higher if they were all put into one black hole. In the present context
in which baryons are believed to be produced by baryon nonconserving processes, I don't
see that Penrose's argument has any relevance. The first step is to explain the entropy of
the visible universe, '" 10 90 , and inflation can do this. The next step is to explain how each
10 10 units of entropy can produce one unit of baryon number, and grand unified theories
can presumably do this. And if inflation has spread the entropy uniformly through space,
then the baryons will also form uniformly, which is consistent with what we observe.

Rees: Although I'm perhaps less antipathetic to anthropic arguments than you are, I
think the strongest reason for regarding them as inadequate comes from the large-scale
anisotropy (tl.T IT < 10-4 ). There seems no purely anthropic reason why our existence
requires the universe to be so smooth on scales larger than superclusters.
21

Guth: Yes, I agree.

Hoyle: Are the partial differential equations connecting inflationary regions with those
of false vacuum actually solved? Or are the regions so big that the equations in different
regions are effectively chopped off from each other?
Guth: The equations that describe a spherically symmetric inflationary universe that
forms within a larger region of false vacuum are not difficult to write down and solve. And
since the de Sitter region expands so quickly that if forms causal horizons, these newly
forming universes will affect only a finite region of comoving coordinate space. It is then
easy to piece together solutions of this type to form the schematic fractal picture of the
decaying false vacuum that I showed in my talk. However, in a more realistic picture the
decay of the false vacuum would be random, the regions might not have spherical symmetry,
and the decaying regions of false vacuum would sometimes overlap. This situation could
get very complicated, and I would not be at all capable of solving these equations.

Starobinsky: Comment: First, I would like to remind the audience about two of my
papers on the Universe originating from the maximally symmetric (de Sitter) initial state
(this stage of exponential expansion was later called "inflation" by Prof. A. Guth). The
first paper (1979) contains a direct observational test of the existence of such a state in
the past - the search for a stochastic background of gravitational waves with the char-
acteristic flat spectrum generated during the de Sitter stage. The second paper (1980)
contains an internally self-consistent model of the Universe beginning from the de Sitter
stage and ending in the radiation - dominated stage. This model is purely geometrical,
it is based on a variant of the so-called higher-derivative theory of gravity. However, it is
remarkable that it was proved later that, from the gravitational point of view, this model is
completely mathematically equivalent to inflationary models based on the Einstein gravity
interacting with a minimally coupled scalar field with some specific potential (the latter is
actually intermediate between those used in "new" and "chaotic" models). Thus, from the
gravitational point of view, there exists no principal differences between these models.
Second, let me add a response to the question of Prof. Rees, too. It seems to me that
in case of many low-density universes separated by still inflating domains borders of the
latter ones are inaccessible for us (just as we can't hit a particle moving with a constant
acceleration even by means of photons, if it is sufficiently far from us initially). Moreover,
if constant energy density hypersurfaces are defined to be the hypersurfaces of constant
time or the synchronous system of reference is used (as is usually done in cosmology), these
inflating domains and other universes beyond them always lie in the past for us.
PREDICTIONS OF INFLATION

ANDREW R. LIDDLE
Astronomy Centre,
Division of Physics and Astronomy,
University of Sv.ssez,
Falmer, Brighton BNl 9QH, U. K.

ABSTRACT. Many models exist based upon the inflationary universe paradigm. The predictions
of these models are outlined, and it is emphasised that the differing models can have substantially
different implications. Consequently, inflation seems more likely to fall from favour through us being
restricted to unappealing models rather than being explicitly ruled out. Whether or not the universe
is at the critical density remains the most solid testable prediction of inflation.

1 Inflation and its Aims

After ten years, inflation (for a selection of reviews of differing emphasis, see [1,2,3,4])
remains a very attractive solution to a number of cosmological conundrums, and its influence
pervades, sometimes explicitly and sometimes implicitly, a considerable segment of modern
cosmology (for instance, it is often used as a justification for a scale invariant spectrum in
structure formation simulations [5]). One is therefore entitled to ask how much observational
evidence actually points in favour of the inflationary cosmology, and further one should of
course be very interested in knowing the types of evidence capable of ruling out the scenario.
This, in a nutshell, is the purpose of this conference.
As this article is very much aimed at providing an introduction and setting for topics
which will be examined in more detail in the remainder of this volume, I intend to keep
my presentation almost entirely non-mathematical, with pointers to the relevant parts of
the literature should the reader be interested in following up the details. After a brief
reminder of the aims of inflation, I will classify the different types of inflationary model
before proceeding with a discussion of the predictions (some fairly solid, others much less
so) that these models make. One point which will become clear during the article is that a
huge number of models have sprung from the original inflationary paradigm. These models
are capable of a whole range of predictions, and this means that inflation as a whole shows
considerable robustness in the face of observations. Hence the immediate moral is that it
is very hard to slay inflation with a single observation.
I'll begin by defining what I mean by inflation. Inflation proposes to solve a number
of cosmological problems by invoking a very rapid expansion of the universe in its earliest
stages [6,7]. The definition of inflation I will use is a period in the early universe where the
scale factor a obeys the condition a(t) > O. (For example, conventional exponential inflation
a '" eHt has a '" H 2 eHt .) Equivalently, inflation is an epoch in which the separation
between points that have been in causal contact may increase faster than the speed of light.
The hot big bang model is unable to answer several problems whose origin is outwith
23
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 23-38.
© 1991 Kluwer Academic Publishers.
24

the influence of the model. A selection of these are

• The Flatness Problem: Why is n (the ratio of the universe's density to the critical
density that makes the universe flat) so close to one, when in the big bang model
n = 1 is an unstable critical point?
• The Horizon Problem: Why are points at the opposite ends of the universe, which
have never experienced causal contact, in thermal equilibrium at the same tempera-
ture?

• Homogeneity and Isotropy: Why does the universe have these properties (on
sufficiently large scales that dynamics has been unable to move matter around signif-
icantly)?

• Relic Abundances: What stopped massive stable relics, such as monopoles, coming
to dominate the universe?

As discussed in the previous paper by Guth [8], a sufficiently long period of inflation, mul-
tiplying the scale factor by at least 1030, can provide an explanation for all these problems.

2 Inflationary Models

Before discussing the predictions that the inflationary model makes, it is necessary to detail
the mechanism by which inflation is envisaged to arise. For this, only one observation is
required; that if the universe is dominated by an energy density the solutions to Einstein's
equations are an exponential expansion of the scale factor, satisfying the condition for
inflation. Such solutions are well known from studies of space-times with a cosmological
constant [9]. However, much more significantly for our purposes, such an expansion arises if
the universe is dominated by the potential energy of some quantum field. This is normally
taken to be a scalar field (though some work has been done where inflation is driven by
a vector field [10]); ideally this would be a field associated with a symmetry breaking in
the early universe as such fields are ubiquitous in current particle theories. Recent trends
have however favoured the introduction of scalar fields for the sole purpose of providing a
satisfactory inflationary model. In any case, the potential energy of the scalar mimics a
cosmological constant and leads to inflation.
For the purposes of this article, the plethora of inflationary models can be slotted into
two categories. This is a rather lopsided classification, with nearly all models belonging in
the first category, but useful because the two categories lead to rather different predictions.

2.1 SLOW-ROLLING INFLATION

Members: New inflation [7], chaotic inflation [11], power law inflation [12,13]' R2 (and
other super-Einstein) inflation [14], double inflation [15] ...
In slow-rolling inflation, a scalar field is displaced from the minimum of its potential.
Under a wide range of circumstances [16], its kinetic energy becomes negligible and its
evolution is dominated by its potential energy, thus mimicking a cosmological constant and
25

providing inflation. The scalar field 'slow-rolls' to the minimum of its potential, bringing
inflation to an end - see figure 1.

V(u)

o u
Figure 1: Slow rolling down a potential.

This class includes almost all currently favoured examples; the variants are plentiful
due to the wide choice of possible potentials and the different identities of the scalar, which
for instance may be a general scalar introduced solely to provide inflation, or connected
with extra compact spatial dimensions, or a consequence of extra terms in the gravitational
action. The expansion of space may be approximately exponential, or a power law, or
something intermediate. A small subset of the options, where V( a) is the potential of the
scalar field a, are
• One scalar field, V( a) = tm2a2: Standard chaotic inflation scenario, with a massive
scalar field. Apparently experimentally viable for m < 10 14 GeV.

• One scalar field, V(a) = VoC/A<T: Exact power law inflation solutions a ex: t 2 //A 2 •
Satisfy criterion for inflation provided 2/ p,2 > 1. Such exponential potentials are
common in higher dimensional theories.

• Two or more scalars with appropriate potentials: Double (multiple) inflation. More
than one inflationary epoch; the epochs may show quantitative differences.
Clearly many varieties of the slow-rolling model exist and can be made to work. How-
ever, such models commonly suffer the plague of requiring fine-tuning [17) (though see [18]
for a case where the fine- tuning arises from the particle physics). This is because in order for
sufficient inflation to occur, the potentials used for these models have to be very flat, which
also introduces problems because the density perturbations arising from inflationary models
(see later) with such flat potentials are generally unacceptably large, and one must be care-
ful to keep their amplitude low. That the potential should be flat requires the introduction
of a small dimensionless parameter, as quantified in [17). In some particle physics models,
especially those where the fundamental parameter in the action is dimensionless, this can
appear very unnatural. However, it is much less unnatural for dimensionful parameters; for
26

instance in the basic chaotic inflation scenario above the small parameter required is just
m2 /mJ,t, forcing m < 1014 GeV, but no-one feels disturbed by the possible existence of such
a mass in the way they would when confronted with a dimensionless coupling of 10- 10 . The
naturalness or otherwise connected to the fine-tuning parameter is still a disputed issue at
this time.

2.2 INFLATION WITH A META-STABLE VACUUM

Members: Old inflation [6], extended inflation [19], double field inflation [20].
In this rather under-represented version of inflationary model, inflation occurs via a
first order phase transition. As the universe cools, the scalar field becomes trapped in
a false vacuum state which has non-zero potential energy, as in figure 2, giving rise to
inflation as before. Inflation is brought to an end by quantum tunnelling of the scalar to
the true vacuum, via nucleation of true vacuum bubbles within the sea of false vacuum.
Once quantum effects cause these bubbles to nucleate, they expand at close to the speed
of light and collide with each other. All the original false vacuum energy is swept up into
the walls. These collisions of the bubbles release the energy stored in the bubble walls and
the post-inflation universe thermalises.

V(a)

o
o a
Figure 2: Trapped in a meta-stable vacuum.

This model is very much in the spirit of Guth's original (now dubbed 'old') inflationary
scenario. Unfortunately, it fails to work when the space expands exponentially, because
although the nucleation of bubbles, like a decay process, is also exponential it is the ex-
ponential expansion of the false vacuum regions which dominates; consequently the phase
transition cannot complete and give rise to the universe around us [21]. This situation has
been salvaged recently in the guise of 'extended' inflation [22], where the first order transi-
tion is implemented in a gravitational theory extended beyond Einstein's by the addition
of extra terms (a pursuit very popular with particle physicists, particularly in the quest for
the quantum theory of gravitation). In such a theory, the solutions for the expansion of
space tend to be slower than exponential (often a rapid power law). In such circumstances,
27

the exponential nucleation of true vacuum bubbles always dominates eventually, and the
transition completes and can give rise to a conventional universe. These models can usually
be regarded as having a time-varying gravitational 'constant' or similar.
It remains early days for extended inflation, and I include it here essentially in a section
of its own merely to highlight the rather different mechanism because it can give rise to
substantially different predictions than the slow-rolling models; it still remains one model
amongst many. The hopes and fears of the extended inflation model at present are
Hopes Inflation may be implemented as part of a Grand Unified Theory,
rather than by the introduction of a scalar for the sake of it.
Small enough density perturbations without fine-tuning the inflaton
field [23J.
Naturally consistent with cosmic strings; indeed, the string and in-
flat on fields can be one and the same [24J.
The density perturbations may have an interesting form [23J.
Fears May require a fine-tuned gravitational sector [17J.
May not solve the monopole/domain wall problem if it arises in the
GUT context [24J.

Presumably time will tell whether the hopes or the fears win out; at this moment the
fear of requiring a fine-tuned gravitational sector is probably the most prominent.
Double field inflation [20J is essentially a variant on this theme where an extra scalar
is added by hand (rather than as part of an extended gravitational theory) in order to
precipitate the successful completion of the first order transition.

3 The Inevitable Predictions of Inflation

Inflation was devised with the goal of solving a particular set of cosmological problems,
and hence clearly predicts a universe in which these problems are solved. Certainly such a
universe fitted observations at the time inflation was devised, but one is free to ask whether
or not such a universe is consistent with more recent observations. So let us examine these
points once more, with a view to determining what can rule out inflation .

• n =1 (± '" 10- 4 ) : A quick glance at the contents of this volume will rapidly con-
vince the reader that this is an extremely important prediction. While observations
may only suggest, taking an extremely pessimistic viewpoint, that 0.01 < n < 10,
inflation is much more stringent, and in the process of solving the other cosmological
problems adequately n is forced very close to 1 [25). (Note that I am including any
contribution from a cosmological constant in n.) As we shall shortly see, quantum ef-
fects during inflation can also give rise to density perturbations which are constrained
by current observations to be no larger than say 10- 4 . These are superimposed on the
'classical' value of 1, which would be obtained to a much higher degree of accuracy
even than this if density perturbations were not generated. Because n is so close to
1, we predict ...
28

• Dark Matter: If we assume (as seems eminently reasonable) that the standard model
for primordial nucleosynthesis [3,26] is correct, then we are forced to accept that the
baryonic content of the universe can provide no more than 0.1 of the critical density.
The actual visible baryonic matter provides even less than this, and quite possibly
less than the lower limit from nucleosynthesis, prompting suggestions for some of the
baryonic matter to be dark [27]. For our purposes here, a much more significant
prediction is that at least 90% of the matter in the universe is both non-baryonic
and dark. The discovery of dark matter would be a significant bonus for inflation;
one might hope that the identity of dark matter may be discovered via terrestrial
detectors, say for axions or WIMPs [28]. It is also possible that dark matter may
reside in compact objects, perhaps detectable via gravitational microlensing [29]. At
worst, it could reside in the form of a cosmological constant and be undetectable on
short scales [30].
• Homogeneity and Isotropy: The limits on anisotropy from the microwave back-
ground experiments become stronger and stronger; there seems no doubt that but
for small superimposed density perturbations the universe is isotropic and homoge-
neous at horizon crossing (the homogeneity being more dramatically broken later by
structure formation on sub-horizon scales).
• Lack of Relics: The vast amount of inflation should reduce the abundance of stable
relics by at least a factor of 10 90 . Hence, even if the underlying theory has magnetic
monopole solutions, one would not expect to find any. Should a significant abundance
of high energy relics be found then one would have to suspect that a less drastic
solution to the relic abundance problem should be found. The simplest versions of
slow-rolling inflation are also inconsistent with the cosmic string galaxy formation
scenario [31,32], as strings form before or early during inflation and are diluted away.
(This is much more a problem for strings than inflation, as it deprives the string
scenario of a solution to the horizon problem etc.) Such an inconsistency does not
however arise in extended inflation [24].
The above represents the implications of a fairly standardised inflation scenario - the sort
of result one would get by averaging over the currently viable models. However, the escape
clauses are many and varied, probably to the extent of the number of models. For instance,
an extended inflation scenario which has just enough inflation to solve the various problems
may have inflated by sufficiently little that n is not driven as close to 1 as in slow-rolling
models [25], which require more inflation to avoid the density perturbation constraint.
This would leave some probability for n < 0.5, say. Equally, any inflationary model with
sufficiently carefully chosen initial conditions could lead to a smaller n [33]. But in either
case it seems a very unnatural choice must be made, such as why there should be just
enough inflation to solve say the horizon problem in this epoch, but not in any subsequent
ones. But such models may avoid the requirement of dark matter, particularly of the non-
baryonic kind (should this be considered a good thing). There do exist slow-rolling inflation
models which are consistent with cosmic strings [32,34], and it is even possible to tune those
models to allow an 'interesting' monopole density near the Parker limit [35]. However, the
extent to which particular models are to be believed is entirely at the discretion of the
researcher.
29

4 The Inferred Consequences of Inflation

Beyond those implications listed in the preceding section, inflation does of course have extra
consequences not connected with the solution of cosmological problems. These thus provide
additional opportunities for experimental tests of inflation.

4.1 GENERATION OF THE PRIMORDIAL DENSITY PERTURBATION SPECTRUM

This is by far the most important prediction of this section. Because inflation implies that
scales which are large (horizon sized or greater) today were once in causal contact (when
those scales were microscopic), it allows a causal production of large scale density pertur-
bations to take place. The standard mechanism is via the freezing of quantum fluctuations
on scales as they move outside the horizon during inflation. A given inflationary model
leads to a prediction for the amplitude and shape of the spectrum [36J.
The standard assessment of density perturbations from inflation is that

• Inflation predicts a Harrison-Zel'dovich spectrum - constant amplitude at horizon-


crossing.

• The spectrum amplitude depends on the parameters of the model.

• The perturbation statistics are Gaussian.


Figure 3 is a very important figure which demonstrates how inflation gives rise to density
perturbations causally. The Hubble radius H-I is the scale over which causal interactions
(within a given epoch) can act. During conventional matter or radiation dominated evo-
lution, this simply grows at the speed of light, ex t, but during inflation it is a constant.
A indicates the physical size of a comoving scale, which simply grows proportional to the
scale factor; that is, ex t l / 2 , t 2 / 3 during radiation and matter domination, but exponentially
ex e Ht during inflation. For the sake of this diagram, we can imagine A to be say the scale of
galaxy clusters. Note that since inflation increases A by a factor 10 30 at least, this diagram
is not drawn to scale!
Observe on the right hand side the time when the illustrated scale comes within the
horizon. If preceding that there were only matter or radiation dominated evolution, that
scale would never be within the horizon at an earlier time. However, if an era of inflation
occurred, then (looking backwards in time), our scale decreases exponentially rapidly while
the Hubble radius stays constant; sufficiently early in inflation we find our scale inside the
horizon and hence the possibility of causal generation of density perturbations exists at that
time. While the scales of cosmological interest today are within the horizon, they are mi-
croscopic and quantum effects are important, leading naturally to fluctuations which break
homogeneity. As the scale crosses outside the horizon, fluctuations existing at that time
suddenly find themselves frozen in, and persevere outside the horizon until later crossing
back within the horizon during the radiation or matter dominated era following inflation.
(This discussion is somewhat heuristic; for a more precise analysis see [1J.)
[A quick caveat is probably in order; it is certainly possible to generate useful pertur-
bations on the horizon scale causally even without inflation, provided one has a system in
which the correlation length grows at the speed of light. An example (indeed the only one
30

I ... - inflation - ~ I

o
o Time
Figure 3: The evolution of a wavelength A and the horizon size H-l during conventional and
inflationary expansion demonstates when causal interactions can act on a given comoving
scale.

I can think of) is the use of topological defects such as cosmic strings [37] or textures [38]
to provide density inhomogeneities. The existence of horizons implies that the phases of
the symmetry breaking fields featuring these defects must be uncorrelated above the hori-
zon scale; hence as the correlation length increases at the speed of light new uncorrelated
regions are constantly entering the horizon. Of course, these models may be ruled out in
various ways, which may only leave inflation, which has much greater flexibility, as a viable
mechanism.]
In the simplest inflationary models, the predicted spectrum is scale invariant; this is
essentially because the exponentially expanding space-time possesses a time translation
symmetry which ensures that all scales crossing outside the horizon feature the same physics.
In slow-rolling models, if one is not careful, the predicted amplitude of the perturbations
is far too large, of order 1 rather than 10- 4 • Being careful in this context normally means
the introduction of a small (fine-tuning) parameter to suppress the amplitude. Of course,
there is also no reason why the density perturbations should not be suppressed to well
below useful levels. In extended inflation, excessive density perturbations are probably not
a problem, though there the spectrum generically departs from the scale-invariant form [23]
(see later).
The primordial spectrum is presumably experimentally testable, at least once one be-
comes better acquainted with the nature of any dark matter. Hence one could hope to test
this inflationary prediction. Unfortunately, once more the plethora of inflationary models
comes into play; many of these feature significantly different predictions for the primordial
spectrum. More often than not, the prediction is for large scale power, which may actually
31

be useful for cold dark matter models. A selection of these are as follows

• Power Law Models: The spectrum ~ Ihor"" ).jJ2/(2_jJ2) [13] has more large scale
power (remember a "" t 2 /jJ2, so p,2 < 2. We tend to a scale invariant result as p, -> 0,
a limit in which the power law becomes exponential.

• Double Inflation: We may have different amplitudes [15] on different scales, because
different scales may belong to different inflationary epochs; the spectrum may have
the form of a step. This could for instance provide extra large scale power, though
the position of the step would have to be appropriately tuned.

• Extended Inflation - Quantum Fluctuations: Extended inflation generates den-


sity perturbations just as does the slow-rolling variant. However, these are not pre-
dicted to be scale-invariant, due to the variation of the gravitational constant. They
normally pick up extra power on large scales [23], with a fairly weak dependence on
scale. ~
p
Ihor= 1O-4).M
1 / 5 is a typical example (). measured in Megaparsecs); note also
pc
the substantially reduced short-scale power.

• Extended Inflation - Bubble Distribution: An aspect often mentioned (see e.g.


[23,39)) but seldom quantified is that the collisions from the largest colliding bubbles
may lead to residual density fluctuations. Some preliminary steps towards investigat-
ing these have been taken [40], but a proper assessment of the microwave background
constraints on such perturbations is required. Under rather special circumstances
these perturbations can be isocurvature [41]. They are of course in addition to the
quantum fluctuations, though they can be removed by suppressing large bubbles.

• Designer Potentials: One can consider inverting the problem; choose a desired
perturbation spectrum and ask if we can choose an inflaton potential that gives it,
either by analytic or numerical construction [42]. It seems that as often as not we
can, with features in the potential giving rise to features in the spectrum. What
naturalness conditions apply to the choice of potentials is far from clear, though
aesthetically one would like to see as simple a potential as possible.

• Non-Gaussian Perturbations: There seems no reason why inflation models with


suitable non-linearities in the basic physics should not be able to produce non-gaussian
statistics [43]. Whether or not these are actually desirable remains an open question.

• Alternative Sources: Inflationary perturbations can easily be suppressed to a level


where they will be too small to influence structure formation. In such a case another
mechanism may be the relevant one, such as cosmic strings, textures or something
completely new.

4.2 GRAVITATIONAL WAVE PRODUCTION

The quantum fluctuations responsible for generating density perturbations can also excite
transverse-traceless gravitational wave modes. This gives a stochastic gravitational wave
background, which can persist until today as gravitational waves are so weakly coupled as to
32

be effectively collisionless [44]. Unfortunately, as with density perturbations, the amplitude


of this spectrum is undetermined, though it should be flat over a wide range of wavelengths.
Its amplitude is constrained by the microwave background quadropole anisotropy, though
the amplitude could be well below this limit. This constraint forces down the amplitude
across the entire wavelength range, putting it far below the range accessible to detection
by laser interferometers.
The hot big bang model predicts a relic thermal spectrum of gravitons at O.9K. As this
would be generated before inflation, inflation does make the prediction that this spectrum
is absent. However, it is nowhere near detect ability in any case.

0 \ 1 UGO-1
\ I

LO
\1
I ~ UGO-2 cST/T
..--..
N
~

~ I
0
..... V Bubble collisions
a LO
'-'"
Cl .....
0 Inflation
.....J
Maximum quantum
0
C\I fluctuations
I
LO
C\I
I

0 10 20 30
Log (A/em)
Figure 4: The predicted gravitational wave spectrum, measured in fractions of critical
density per octave, and the proposed observational capacity of first and second generation
laser interferometers. The inflation prediction can be scaled across its range by any factor
which satisfies the microwave quadropole bound.

It has been suggested [45] that extended inflation models have a further and much
more significant source of gravitational waves in the undoubtably violent process of bubble
collisions. One would expect this to lead to gravitational waves with a wavelength around
the horizon size at the time of the bubble collisions. Once formed, the gravitational waves
are merely stretched by the expansion of the universe; by a fortuitous coincidence the
horizon scale after a phase transition in a Grand Unified theory is conform ally stretched to
a size of order kilometers today - exactly the size of proposed laser interferometer detectors.
Hence these detectors are ideally suited to detecting such waves. The amplitude has been
estimated in [45], and they suggest that it may be within the range of a second generation
laser interferometer. However, their estimate is very optimistic and the true result could
be several orders of magnitude lower, as well as displaced to larger or smaller wavelengths.
33

4.3 THE COSMOLOGICAL CONSTANT

The cosmological constant A doesn't really count as a prediction of inflation, but inflation
does provide a useful reference point in considering how reasonable such ideas, which are
gaining renewed attention, are. After all, the mechanism for inflation relies on a temporary
cosmological constant term. Can we expect such a term today?
There are two obvious sources for a residual cosmological constant after inflation. The
first is that the minimum of the potential of the inflaton need not necessarily be at zero
energy. While particle physics normally only worries about differences in potential, general
relativity is sensitive to the overall scale. The second source is that even if the minimum has
potential energy zero, it will still have a zero-point energy, just like a quantum mechanical
harmonic oscillator. Naively, one would expect these contributions to be very large [46];
indeed, so large that one would predict fiA > 1 (more accurately, the cosmological constant
would have had such an important influence in the early stages of the universe that the
universe could never have reached its present state). Why this is not so is a very serious
problem, and I think it's fair to say that as yet there is no compelling reason to think that
anyone has a clue how it will be solved.
Of course, there may be an overall selection rule, something higher than inflation, which
determines that fi tot = 1. But in that case if one is to advocate a cosmological constant
as dark matter, an idea that seems to be gaining in popularity, one must still explain why
Pbaryon and PA are of the same order of magnitude today when Pbaryon dilutes as the cube
of the scale factor while Ph remains constant. If A is taken to comprise say 0.8 of the
critical density today then it will be totally dominant in a few Hubble times. To have the
densities of the same order today requires a fine-tuning which can be every bit as puzzling
as those inflation sets out to solve. In that sense, while a cosmological constant may not be
in contradiction with inflation itself, it would certainly appear to contradict the philosophy
of inflation.

5 Conclusions

The above leads to the conclusion that the most solid test of inflation is the value of fi.
While fi f:. 1 is certainly possible with inflation, this for now appears very unnatural and
such an observation would be a vicious blow to the theory. (Of course, it is possible to
argue that given a value of fi less than 1, the inflationary model leading to such a universe
still requires much less fine-tuning than the hot big bang model does. However, one would
have to feel that that is not the whole story.) Determinations of fi are possible from several
angles, all represented in this conference. They include

• Distance Scales: An estimate of the Hubble constant combined with a lower bound
on the age of the universe (say from globular clusters) gives an upper bound on fi.

• Dark Matter Searches: Can give estimates of the density in dark matter.

• Structure Formation: The galaxy distribution on various scales gives information


about the density on these scales, though non-linear effects on short scales and dif-
34

ficulties related to selection and evolution effects in observing large scales can make
this tricky.
Inflation makes no definite prediction on density perturbations; the best that one can hope
for is that investigations of the spectrum will constrain the viable inflationary models. A
similar situation arises with the gravitational wave spectrum, though perhaps extended
inflation has a clear signal.
The attentive reader will have noticed by now that I have discussed ways in which
inflation may be ruled out. What has not appeared is a positive signal which would vindicate
inflation. After all, the prediction that n = 1 for instance is not the sole property of the
inflationary paradigm. The rather disappointing conclusion of this is that perhaps the best
inflation can hope for is to frequently fail to be ruled out.
Ultimately, then, inflation probably lives or dies based upon what researchers are willing
to accept as reasonable; I certainly would be very unhappy with some of the more exotic
models which I have mentioned. For now the number of viable models still seems to be
increasing with time; hopefully observations shall soon provide more stringent restrictions
and the number of viable models will start decreasing. Once that happens, we will probably
be in a much better position to assess the desirability of inflation.

Acknowledgements. Those who attended this conference may remember that I only discovered I
was giving this talk the night before, and that it was written in a panicked rush. The written version
has been compiled at a much more leisurely pace, though I decided to keep the general layout of
the original. I would like to thank the Organising Committee, and in particular Tom Shanks and
Richard Ellis, for their encouragement, and also for organising an extremely enjoyable Workshop. I
would also like to thank John Barrow for helpful discussions.

References

[1] S. K. Blau and A. H. Guth in 900 Years of Gravitation, eds S. W. Hawking and W. 1.
Israel, Cambridge University Press (1987)
[2] A. D. Linde in 900 Years of Gravitation, eds S. W. Hawking and W. 1. Israel, Cambridge
University Press (1987)
[3] E. W. Kolb and M. S. Turner, The Early Universe, Addison Wesley (1990)
[4] K. A. Olive, Phys Rep 190 (1990) 307

[5] G. Efstathiou in Physics of the Early Universe, eds J. A. Peacock, A. F. Heavens and
A. T. Davies, Scottish Universities Summer Schools Publications (1990)

[6] A. H. Guth, Phys Rev D 23 (1981) 347


[7] A. Albrecht and P. Steinhardt, Phys Rev Lett 48 (1982) 1220
A. D. Linde, Phys Lett B108 (1982) 389
35

[8] A. H. Guth, in this volume

[9] S. W. Hawking and G. F. R. Ellis, The Large Scale Structure of Space-Time, Cambridge
University Press (1973)

[10] L. Ford, Phys Rev D 40 (1989) 967

[11] A. D. Linde, Phys Lett B129 (1983) 177

[12] L. F Abbott and M. B. Wise, Nucl Phys B224 (1984) 541


J. D. Barrow, Phys Lett B187 (1987) 12
A. R. Liddle, Phys Lett B220 (1989) 502

[13] F. Lucchin and S. Matarrese, Phys Rev D 32 (1985) 1316

[14] A. A. Starobinsky in Quantum Gravity, eds M. A. Markov and P. West, Plenum (1983)
B. Whitt, Phys Lett B145 (1984) 176
M. Mijic, M. Morris and W.-M. Suen, Phys Rev D 34 (1986) 2934
J. D. Barrow and A. C. Ottewill, J Phys A16 (1983) 2757

[15] J. Silk and M. S. Turner, Phys Rev D 35 (1987) 419


H. M. Hodges, Phys Rev Lett 64 (1990) 7
A. A. Starobinsky, in this volume

[16] H. Feldman and R. H. Brandenburger, Phys Lett B227 (1989) 359


J. H. Kung and R. H. Brandenburger, Phys Rev D 42 (1990) 1008

[17] F. C. Adams, K. Freese and A. H. Guth, "Constraints on the Scalar Field Potential in
Inflationary Models", MIT preprint (1990)

[18] K. Freese, J. A. Frieman and A. V. Olinto, Phys Rev Lett 65 (1990) 3233

[19] D. La and P. J. Steinhardt, Phys Rev Lett 62 (1989) 376


E. J. Weinberg, Phys Rev D 40 (1989) 3950
D. La, P. J. Steinhardt and E. W. Bertschinger, Phys Lett B231 (1989) 231

[20] F. C. Adams and K. Freese, "Double Field Inflation", MIT preprint (1990)

[21] A. H. Guth and E. J. Weinberg, Nucl Phys B212 (1983) 321

[22] E. W. Kolb, "First-order Inflation", to appear in The Birth and Early Evolution of the
Universe, Nobel symposium #79

[23] E. W. Kolb, D. S. Salopek and M. S. Turner, "Origin of Density Fluctuations in


Extended Inflation", Fermilab preprint (1990)

[24] E. J. Copeland, E. W. Kolb and A. R. Liddle, Phys Rev D 42 (1990) 2911

[25] P. J. Steinhardt, Nature 345 (1990) 47


36

[26] A. M. Boesgaard and G. Steigman, Ann Rev Astron Astro 23 (1985) 319
K. A. Olive, D. N. Schramm, G. Steigman and T. P. Walker, Phys Lett B236 (1990)
454

[27] B. J. Carr, in this volume

[28] P. F. Smith and J. D. Lewin, Phys Rep 187 (1990) 203

[29] B. Paczynski, Ap J 304 (1986) 1

[30] O. Lahav, in this volume

[31] M. D. Pollock, Phys Lett B185 (1987) 34

[32] E. T. Vishniac, K. A. Olive and D. Seckel, Nucl Phys B289 (1987) 717

[33] G. F. R. Ellis, Class Quant Grav 5 (1988) 891


M. S. Madsen and G. F. R. Ellis, Mon Not Roy astr Soc 234 (1988) 67

[34] J. Yokoyama, Phys Lett B212 (1988) 273


J. Yokoyama, Phys Rev Lett 63 (1989) 712
[35] J. Yokoyama, Phys Lett B231 (1989) 49

[36] A. H. Guth and S.-Y. Pi, Phys Rev Lett 49 (1982) 1110

[37] A. Vilenkin, Phys Rep 121 (1985) 263

[38] N. Turok, Phys Rev Lett 63 (1989) 2625

[39] D. La and P. J. Steinhardt, Phys Lett B220 (1989) 375

[40] D. La, "True Vacuum Bubbles and the Origin of Voids", University of California
preprint (1990)
[41] J. D. Barrow, E. J. Copeland, E. W. Kolb and A. R. Liddle, to appear, Phys Rev D,
February 1990

[42] D. S. Salopek, J. R. Bond and J. M. Bardeen, Phys Rev D 40 (1989) 1753


H. M. Hodges and G. R. Blumenthal, Phys Rev D 42 (1990) 3329

[43] D. S. Salopek, in this volume

[44] V. A. Rubakov, M. V. Sazhin and A. V. Veryaskin, Phys Lett B115 (1982) 189
R. Fabbri and M. Pollock, Phys Lett B125 (1983) 445
L. F. Abbott and D. D. Harari, Nucl Phys B264 (1986) 487
B. Allen, Phys Rev D 37 (1988) 2078

[45] M. S. Turner and F. Wilczek, "Relic Gravitational Waves and Extended Inflation",
Fermilab preprint (1990)
[46] T. Padmanabhan, Int J Mod Phys A4 (1989) 4735
37

DISCUSSION:

Salopek: You mentioned naturalness of models. I would like to point out that there are
currently about 9 different models that give the correct level of density fluctuations even if
the potential ¥ is steep, A 2: 10-3 . See for example, Phys. Rev. D4 175 D. S. Salopek, J.
R. Bond, J. M. Bardeen, who employ a -~R"': coupling to gravity in chaotic inflation.

Liddle: I don't doubt that many models exist predicting the correct level of density per-
turbations, some with and some without obvious fine-tuning. Indeed, the thrust of the talk
was that it is far too easy for inflation to produce likely-looking density perturbations for
such considerations to significantly constrain model building as such.
One would have to assess whether or not the incorporation of extra non-minimal terms
was 'natural' or not.

Borner: Wouldn't an obvious 'natural' choice for an extended inflation potential be the
effective potential energy VeJJ of a scalar field if derived from a local field theory? Then,
however, the double-lump shape is a result of perturbation theory. The true VeJJ must be
convex. It seems difficult then to achieve an extended period of inflation.

Liddle: Though my knowledge of the theory of effective potentials is somewhat vague,


it is not clear to me that the normal results hold in the presence of a metastable vacuum
state at finite temperature. The tunnelling at the end of extended inflation is doubtless a
non-perturbative process.
In any event, one can readily see that the potential cannot adopt a symmetry break-
ing configuration (i.e. possess degenerate minima) without losing its convexivity. There
may well, however, be worries connected to the existence of massless propagators where
V"(u) = o.
Ellis: What scope is there to increase the flexibility of ± 10- 4 on the n= 1 prediction of
inflationary theory?

Liddle: In the context of slow-rolling inflation, essentially none. In order to achieve the
aims of inflation, and in particular to ensure sufficiently small density perturbations, n is
driven extremely close to 1. Assuming 1030 factors of expansion, one would expect In -11 :::;
(10- 30 )2 in the absence of density perturbations. The ±1O- 4 I mentioned arises from the
superposition of fluctuations on the n = 1 solution; such fluctuations are as large as one
can allow.
In extended inflation, if just enough inflation occurs to solve the horizon problem it
may be possible that n is not driven so close to 1; there seems a limited parameter range
even with n < 0.5. However, I would consider such a scenario to be unnaturally fine-tuned
although not excluded.

Sanz: Is there any definitive answer as regards the global topology of the universe that
emerges from the inflationary epoch?
38

Guth: The geometry of an inflationary universe is very complicated, as pieces of the false
vacuum decay to form normal universes, while other pieces go on inflating. We know, for
example, that an isolated sphere of false vacuum will form a new universe by splitting
off completely from the original space. From the point of view of the original space, it
will look like a black hole. The structure that develops when the false vacuum decays
randomly, however, is really too complicated for anyone to calculate or describe. It can be
said, however, that the really large scale global structure is not changed by inflation. For
example, if one starts from a closed Robert son-Walker universe and allows it to undergo
inflation, it will remain a closed universe. The closed universe would become so huge,
however, that its spatial curvature would be imperceptible, and it might also be sprinkled
with black holes.
CLASSICALITY OF DENSITY PERTURBATIONS IN THE EARLY UNIVERSE

Robert Brandenberger
Physics Department, Brown University
Providence, RI, 02912, USA
Raymond Laflamme
D.A.M. T.P., University of Cambridge
Cambridge, U.K. CB39EW
Milan Mijic
Scuola Internazionale Superiore di Studi A vanzati
Strada Costiera II, 34014 Trieste, Italy

Abstract. Quantum fluctuations in the de Sitter phase of an inflationary Universe


may produce the classical primordial density perturbations required to seed galaxies.
Implicit in this theory is a prescription for extracting classical quantities from a
quantum theory. We briefly review the original prescriptions for obtaining classical
density perturbations. We then solve a simple toy model and show that the original
prescriptions can be justified by applying decoherence arguments similar to those
used in quantum cosmology.

There has been a lot of progress recently in understanding the early era of
the Universe. The inflationary theory, in particular has received much attention.
It might not only explain the large scale homogeneity and isotropy[l] but also
elucidate the origin of very small inhomogeneities responsible for the formation
of structure in the Universe[2]. Inflation predicts inhomogeneities generated by
quantum fluctuations of scalar fields evolving in the Universe[3-6]. In the simplest
inflationary Universe models the perturbations have a scale invariant Harrison-
Zel'dovich spectrum, well fitted to describe the distribution of galaxies and clusters
of galaxies. A crucial assumption in this derivation is the assertion that these
fluctuations become classical as their wavelength becomes larger than the Hubble
radius (for a review see Ref.[7]).
In the context of quantum cosmology, there has recently been a lot of work
on the conditions under which a quantum mechanical wave function corresponds
to classical evolution[8]. In order that the wave function can be interpreted as a
statistical ensemble of classical trajectories, off-diagonal terms in the density matrix
must be highly suppressed[9J. This implies that interference terms can be neglected.
Interference is associated with the coherence of the system, i.e., the coherence in
the state between different points in configuration space[lO]. A measure of this
is the coherence length which gives the configuration space distance over which
off-diagonal terms are correlated.
An isolated system described by a Schrodinger equation cannot lose its coher-
ence; a pure state always remains pure. However, if it is coarse grained, it may

39
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 39-45.
© 1991 Kluwer Academic Publishers.
40

evolve from a pure to a mixed state. One way to realize coarse graining is to let
the system interact with an environment[10]. The environment consists of all fields
whose evolution we are not interested in. The state of the system is obtained by
tracing over all possible states of the environment. Now, even if the state describing
system plus environment is pure, the state of the system alone will in general be
mixed. Its evolution is no longer given by a Schrodinger equation, but an equation
of Liouville type.
We use units in which 1i = kB = C = 1. The background space-time is described
by an expanding Robertson-Walker (RW) metric with scale factor aCt). Greek
indices run from 0 to 3, roman indices over the spatial indices only.
A crucial question is how to model the environment. Any realistic model will be
very complicated and hard to analyze. However, the basic physics should emerge
from the simplest models. Hence, we choose a model which can be solved exactly:
the system is a real massless scalar field fl> 1, (the infiaton), the environment is
taken to be a second massless real scalar field fl>2 interacting with fl>l through their
gradients. We consider the fields in the de Sitter phase of an expanding Universe
with scale factor aCt) = exp(Ht), where H is the Hubble constant.
The action of system and environment is

(1)

where 9 is the determinant of the background metric with line element given by

(2)
c is a constant measuring the strength of interaction between system and environ-
ment. We shall normalize the conformal time T] such that 17 ranges between -00
and 0 and a = -(H17)-1 with H- 1 being the Hubble radius.
Our Lagrangian is quadratic in the derivatives of the fields and can hence be
diagonalized using fields fl>+ and fl>_ for which the interaction term disappears.
The coherences in the quantum state between fl>+ and fl>_ are purely given by the
initial conditions. For example, in (4) we choose these coherences to vanish. In this
case, a pure state gives rise to a pure state reduced density matrix when summing
over one of the fields. Decoherence of one field cannot occur by summing over the
other one. We, however, suppose that the environment and the infiaton do not form
the diagonal basis. This assumption is reasonable since any inflaton field (whose
reduced density matrix we want) will interact with gravitational perturbations (part
of the environment). As we will show, in this basis decoherence is possible.
We can expand the fields in harmonics in a box of fixed comoving volume (physi-
cal volume a 3) and investigate a particular wavenumber k = (k; +k~+k;)1/2. There
is no coupling between modes with different k. From now on we consider a single
wavelength and drop the index k for convenience.
The evolution of system and environment is described by the Schrodinger equa-
tion
ioi!! = Hi!!
aT]
41

(3)

where 7ri are the momenta conjugate to fi. As initial state we assume a vacuum
state with respect to the fields f+ = Jf+C b;!2
and f- = VI=C which b:;p
diagonalize the Hamiltonian. We take the wave function to be

(4)

with
(5)

which fixes the state to be the Euclidean vacuum[GJ. It has been shown[ll J that the
states picked out by different linear combinations of the fundamental mode functions
exponentially approach the Euclidean (or Bunch-Davies[12]) vacuum. Once we sum
over the second field, we obtain the reduced density matrix for h:

P(h,11j 11) = Jd12'It*(h, 12j 11) 'It (ii 12 j 11) = C exp [-(Afr + A* R + Bfd!)]
(6)
where

(7)

When Ik111 < 1 this density matrix obeys the approximate Liouville equation:

(8)

where
(9)

is the reduced Hamiltonian of the system. >. = c2 kj H21}3 is the diffusion strength,
also called the localization rate by Joos and Zeh[lOJ. It describes the overall effect
of the scattering from the environment. It produces an irreversible damping of the
interference terms in the density matrix and effectively multiplies the density matrix
by an exponential term which vanishes as its arguments are far from the diagonal.
This is exactly the effect required to decohere the system.
42

We define a dimensionless measure of the coherence /'i, by dividing the coherence


lenth lc = (A+AL B )1/2 by the dispersion a = (A+A.~B)1/2 to obtain

/'i,= (10)

/'i, goes to zero as the system decoheres. This measure of de coherence is invariant

under field redefinition and is related to the linear entropy. We see that if no
interaction is present (i.e. in the diagonal basis) the coherence length approaches
a constant value. Adding even a small interaction will reduce it to zero. The
coherence length starts decreasing exponentially when the wavelength crosses the
Hubble radius. At this point, q,2 particles are created which act as a bath for <Pl.
Finally, if we calculate the expectation of a mode and its conjugate momentum
we obtain
(11)

in agreement with the results of Ref. 6. Since decoherence starts at Hubble crossing,
the quantities in (11) evaluated for kTJ ~ -1 give the classical initial density per-
turbations. When evolved to the present time, we obtain[3-7] a scale-free spectrum
of perturbations.
In conclusion, we have investigated the qnantum to classical transition of quan-
tum fluctuations of a scalar field in the inflationary stage of the early Universe. '\file
have suggested that there exists an environment which will decohere these fluctu-
ations and make them behave classically. 'Ve have supported this suggestion by
presenting a simple model where indeed 'such behavior occurs. The decoherence
seems to be of general nature and sets in when the wavelength of a mode crosses
the Hubble radius, thus lending support to the usual assumptions made in calcu-
lating the spectrum of small perturbations in inflationary Universe models. Details
can be found in ref.[13].
Acknowlegments. We would like to thank A. Guth and A. Starobinsky for useful
conversations. R.B. is supported in part by the U.S. Department of Energy under
grant DE- AC02-76ER03130 Tasks K & A, and by the Alfred P. Sloan Foundation.
R.L. is indebted to Peterhouse College, Camhridge, for support. The work of M.M.
is supported by the Ministero Italiano per La Ricerca Scientifica.
References.
1. A. Starobinsky, Phys. Lett. 91B, 9D (lD80); A.Guth, Phys. Rev. D23, 347
(1981); K.Sato, Mon. Not. R. astron. Soc. 195,467 (1981); A.D.Linde, Phys.
Lett. 108B, 389 (1982); A.Albrecht & P ..J.Steinhardt, Phys. Rev. Lett. 48,
1220 (1982).
2. W. Press, Physica Scripta 21, 702 (1980); V. Lukash, JETP 52, 807 (1980);
G. Chibisov & V. Mukhanov, JETP Lett. 33, 532 (1981); V. Mukhanov & G.
Chibisov, JETP 56, 258 (1982).
3. S.W.Hawking, Phys. Lett. 115B, 295 (ID82); A.H.Guth & S.Y.Pi, Phys. Rev.
Lett. 49, 1110 (1982).
43

4. A.A.Starobinsky, Phys. Lett. 1I7B, 175 (1982).


5. J. Bardeen, P. Steinhardt & M. Turner, Phys. Rev. D28, 679 (1983).
6. J. Bardeen, unpublished (1984); R. Brandenberger, Nucl. Phys. B245, 328
(1984).
7. R.Brandenberger, Rev. Mod. Phys. 57, 1 (1985).
8. E. Joos, Phys. Lett. 1I6A, 6 (1986); H. Zeh, Phys. Lett. 1I6A, 9 (1986);
S. Wada, Nucl. Phys. B276, 729 (1986); J. Halliwell, Phys. Rev. D36, 3626
(1987); S. Wada, Mod. Phys. Lett. A3, 645 (1988); ibid. A3, 929 (1988); J.
Hartle, 'Predictions in Quantum Cosmology' in 'Gravitation and Astrophysics',
proceedings of the Cargese Advanced Summer Institute 1986.
9. C. Kiefer, Class. & Q1Lant. Grav. 4, 1369 (1987); .J. Halliwell, Phys. Rev. D39,
2912 (1989); T. Padmanabhan, Phys. Rev. D39, 2924 (1989); R. Laflamme,
in Proceedings of the Third Canadian Conference On General Relativity and
Astrophysics, eds. A. Coley, F. Cooperstock and S. Tupper, Singapore: World
Scientific Publishing (1990).
10. W. Zurek, Phys. Rev. D24, 1516 (1981); E. Joos & H. Zeh, Z. Phys. B59, 223
(1985).
11. C. Hill, 'Functional Schrodinger approa.ch to quantum field theory in de Sitter
space and inflation', Fermilab preprint PUB-85-37-T (1985).
12. T. Bunch & P. Davies, Proc. R. Soc. London A360, 117 (1978).
13. R.Brandenberger, R.Laflamme & M.Mijic, Mod. Phys. Lett. A5, 2311 (1990).
44

DISCUSSION:

Starobinsky: I don't want to deny the important role of the decoherence process in many
problems of quantum mechanics, however, for the case of adiabatic perturbations and grav-
itational waves generated during an inflationary stage in the Early Universe, it seems to be
irrelevant. The transition from quantum to classical (but stochastic) behaviour for these
perturbations results purely due to the expansion of the Universe that gives rise to differ-
ent time dependence of quasi-isotropic (growing) and decaying modes for scales exceeding
the cosmological horizon. No explicit account of interactions of these perturbations with
other fields is needed to get classical behaviour. Actually, if one simply neglects the decay-
ing mode and keeps the quasi-isotropic mode only (that is usually done in calculations),
then all perturbations appear to commute with themselves and, thus, may be treated as
c-number random functions. This is clear from the so-called stochastic approach to inflation
developed by me in 1982-1984. On your figures, this corresponds to ellipses degenerating
into straight lines - which are the diagrams for classical stochastic quantities with zero
average. Neither does the process described above depend on the form of the scalar field
potential as far as Im2e//1 == If~ I ~ H2.
Guth: I would agree with Starobinsky, but would like to be a little more specific. It is
true that interactions with an environment can randomize quantum phases, and I'm sure
that it is important to understand that. Nonetheless, I believe that it is not necessary, at
least in this case, to invoke an external environment. So-Young Pi and I wrote a paper
on this subject, and we found it very useful to consider, for a warm-up, the problem of
a quantum mechanical particle evolving in an upside-down harmonic oscillator potential.
One can solve exactly the problem of a wave function that is given initially as a Gaussian,
peaked at the center of the potential. The wave function spreads ad infinitum, so at first
sight it looks very nonclassical. Nonetheless it is possible to write down a classical phase
space probability density function, and to show that expectation values computed with this
probability function agree with those computed in the exact quantum theory, up to terms
that falloff exponentially in time. The operator for which one calculates the expectation
value must belong to a restricted class, but the class includes any power of x multiplied
by any power of p. The probability distribution corresponds to a highly squeezed state-
the x-space distribution is a broadly spread Gaussian, but for each x the momentum
distribution is concentrated in a b-function. In the density matrix language, I think that
the phases disappear because the operators in the restricted class are not sensitive to the
phases at late times. The method can be generalized to a free field theory by treating each
mode amplitude as an upside-down harmonic oscillator; we do not know, however, how to
generalize this method beyond the free field approximation.

Laflamme: I am happy you asked the question because your paper with So-Young Pi was
one of the motivations for our work. There you introduced a classical phase space probability
which was rather ad hoc, in fact it does not obey quantum mechanical properties, for
example the superposition principle (although your initial starting position was a quantum
mechanical wave function). I believe that the nearest analog to the phase space density
45

probability is the Wigner function. It is an object constructed from the wave function and
should converge to the phase space density probability function in the classical limit.
In the case of the harmonic potential, Gaussian wave functions are very special because
they remain Gaussian during evolution. They correspond to the so-called coherent and
squeezed states. They are the only pure states with positive Wigner function and could
naively be interpreted as classical phase probability distribution functions. In fact they are
not: a linear combination of them will strongly interfere, a phenomenon which certainly
does not belong to the realm of classical physics. Secondly the phases do not disappear as
you claim the coherence measure in eq.(lO) is exactly a measure of the phase, and in the
non-interacting case the phase is roughly constant on the scale of the object (defined by its
variance). Only some types of coarse graining (and interaction with an environment is a
special case) will destroy the coherence of the state.
For non-harmonic potential the situation is different, the work of Nieto & Gutschik show
that the coherent/squeezed states do not exist, Gaussian states do not remain Gaussian
under evolution. The pure state Wigner function will not remain positive definite and it will
be impossible to interpret it as a probability distribution. In the WKB approximation, for
example, the Wigner function is wildly oscillatory. The interaction with the environment
would render it positive and allow its interpretation as a phase space probability distribu-
tion. Therefore I would believe that the approach described here could be generalised to
non-linear fields.
Salopek: In realistic inflation models, are nonlinear interactions of a single scalar field
with gravity sufficient to decohere the long wavelength field?
Laflamme: It is the hope behind this paper. We study a very simple environment, however
in reality it could consist of gravitons and any other fields around. There are however
technical difficulties: the interaction is very non-linear and long wavelength modes of the
scalar field would interact with all momentum modes of the gravitational field. But in order
to make sense of the decoherence, a cut-off is needed. One of the most important problems
is to understand this mechanism. It is possible that in the example of cosmology, the
Hubble radius (l/H) might provide a cut-off, only modes longer than (l/H) are excited (in
the Euclidean vacuum) and might contribute to decoherence. It would be rather interesting
if this would work.
THE INFLUENCE OF NON-LINEAR DENSITY FLUCTUATIONS ON THE MI-
CROWAVE SKY

J .L. Sans " Be E. Martines-Gonl8.lel ""


"Departamento de Fisica Moderna,
Universidad de Cantabria,
Avda. Los Castros s.n.,
39005 Santander,
Spain
Be
"Instituto de Estudios Avanzados en Fisica Moderna y Biologia Molecular,
CSIC-Universidad de Cantabria,
Facultad de Ciencias,
Avda. Los Castros s.n.,
39005 Santander,
Spain.

ABSTRACT. Secondary anisotropies are interesting because they can erase/enhance the primary
ones at different scales. We emphasize on the integrated gravitational effect due to the presence
of non-linear density fluctuations as a possible competitive effect with other linear and non-
linear contributions at least on certain angular scales. In particular, we comment on the level of
t:.T/T fluctuations due to cosmic voids, great attractors and second order perturbations related
to theories of galaxy formation.

1. Introduction

The standard picture assumes that the microwave photons are freely propagated (except
for gravitational effects derived from linear theory) from the cosmic photosphere z,. ~ 103 , with
thickness t:.z ~ 10 2 , to the observer. In this context, primary anisotropies arise due to photon
fluctuations at recombination, the fall-in of the last scatters into the gravitational potential
(Doppler effect) and gravitational fluctuations at recombination (the Sachs-Wolfe effect). For an
n = 1 universe, linear density perturbations for vanishing pressure associated to the growing
mode lead to

(I)
However, secondary anisotropies can arise in the microwave sky due to different physical pro-
cesses: i) the gravitational effect of the growing non-linear density fluctuations on the photons
travelling to the observer, ii) hot gas inhomogeneously distributed between recombination and
the present time also generate anisotropies related to non-linear flows (the Vishniac effect) and
iii) Dust present in the early universe produced in the process of formation of bound structures
47
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 47-52.
@ 1991 Kluwer Academic Publishers.
48

also generate anisotropies specially in the Wien region. These are typical examples of the possi-
ble generation of anisotropies during the "dark ages" of the universe which distort the primary
anisotropies at some angular scales. Our aim is to comment on recent calculations about the
first effect related to non-linear density fluctuations.

2. The integrated gravitational effect

In a recent paper (Martinez-GonzaIez, Sanz and Silk 1990) we have calculated a secondary
anisotropy generated by the linear gravitational potential q, associated to non-linear density
fluctuations. For an 0 = 1 universe with vanishing pressure

aT)
(T 1GB =2
[0 dt aq,_
r at (t, i) (2)

This integrated effect, except for the factor 2 that comes from general relativity, can be un-
derstood in terms of Newtonian mechanics; it represents an extra factor needed to take into
account the work performed by the photon, propagating from recombination to the present time,
against the non-static gravitational potential q,.

3. The influence of voids

A void of galaxies in the thin-shell approximation is represented by an empty spherical


region surrounded by a thin-shell where the matter that would have been inside the void is
distributed, and the external universe is Einstein-de Sitter. The radius of the shell expands
according to the law Rex fY with "y ~ 0.13 either for collisionless gas with adiabatic compression
or collisional gas with rapid cooling. This compensated model has been recently considered in
the literature to represent voids like Bootes (Thompson and Vishniac 1987; Scaramella 1989;
Martinez-Gonzalez, Sanz and Silk 1990; Martinez-Gonzalez and Sanl 1990). The integrated
gravitational effect for such a void gives a maximum fluctuation at the center with respect to
the edge of ll.T /T(O) ~ -3 x 10- 1 , being the angular scale of the coldspot of ~ 15°. On the
other hand, if we assume the existence of: i) voids like Bootes at higher redshifts (z > I), the
aT/T(O) slightly increases but the angular scale is drastically reduced to ~ 1°, ii) a supervoid
(i.e. an empty region with radius ~ 100h-1Mpc and placed at a distance of ~ 150h- 1 Mpc) will
give a maximum fluctuation a/T(O) ~ -10- 5 with an angular scale for the coldspot of ~ 50°.
More recently, van Kampen and Martinez-Gonzalez (1990a) have simulated isolated voids (as
-2/T density fluctuations) in an unbiased CDM 0 = 1 scenario and calculated the integrated
effect for 642 photons propagating through the evolving N-body distribution; the result for
aT/T is ~ 10- 1 for a void of radius 5h- 1 Mpc placed at z ~ 0.6. Finally, the statistical effect
of many voids randomly distributed has been calculated by Thompson and Vishniac (1987),
they use voids with the same size in the thin-shell approximation starting at some redshift and
the aT/T value is ~ 10- 6 for the most favorable case of a radius of ~ 30h- 1 Mpc and a filling
factor of 1.
Summing up, considering all of these partial results one can conjecture that for a 0 = I
universe the real statistical distribution of voids in the universe (unless it iz quite different from
the one really observed) will produce a level of fluctuation aT/T that iz ~ 10- 6 , i.e. below the
expectations of present observations.
49

4. The influence of great attractors

A strong concentration of galaxies like the Great Attractor GA (Lynden-Bell et al. 1988),
with a mass of ~ S x 1016 M e at a distance of ~ 40h- 1Mpc, or the Shapley concentration SC
(Scaramella et al. 1989, Raychaudhury 1989), with a mass of ~ 3 x 1017Me at a distance of
~ 140h- 1Mpc, can be represented by a Swiss-cheese model (i.e. a homogeneous sphere of matter
is replaced by a concentric homogeneous internal sphere of matter leaving a shell without any
matter). An estimation of the t:..T/T profile based on time delay was done by Rees and Sciama
(1968), further studies using the framework of general relativity were done by Dyer (1976) and
Nottale (1984). We have recently studied the integrated effect in a clear way through equation
(2) (Martinez-Gonzalez, Sanz and Silk 1990) and applied it to both concentrations of matter
mentioned above (Marinez-Gonzalez and San. 1990). The result is a maximum fluctuation at
the center t:..T/T ~ -3 x 10- 6 for the GA with physical characteristics taken from Dressler
(1988), and t:..T/T ~ -3 x 10- 5 for the SC, being the angular scale of the coldspot of ~ 30· and
20· respectively. Moreover, if we take the GA and place it at a high redshift the effect decreases
and the angular scale is drastically reduced. More recently, van Kampen and Martinez-Gonzalez
(1991b) have simulated isolated lumps (from 2a' to Sa' peaks) in an unbiased CDM n = 1
universe to represent clusters and GA-like structures. Following Bertschinger et al. (1990), such
GA was now defined as an object having a peak amplitud of 6 = 1.2 when averaged over
a radius of 14h- 1 Mpc (this would corresponde to a S.Sa' peak in a CDM scenario). They
have estimated the integrated effect for 64 2 photons propagating through the evolving N-body
distribution representing such a big lump and the result is a fluctuation t:..T/T of a few times
10- 6 , almost independent of redshift. Finally, the statistical effect of many lumps, represented
by Swiss-cheese models randomly distributed, has been considered by Dyer and Ip (1988) finding
that the level of fluctuation is (t:..1'/T)rm. $ 10- 6 (except for a couple of models where the
mass of the lumps is 3 x 10 19 Me and the present number density 3.2 x 10- 3Mpc-3, respectively,
=
and Ho 70kms- 1 Mpc-l).
As a conclussion, considering all of these partial results one can conjecture that, for a n 1=
universe, the real statistical distribution of GA-like structures in the universe will produce a
level of fluctuation that is below the level ~ 10- 5 , whereas for SC-like structures the level could
be above this value, i.e. maybe large-scale observations with the appropiate angular resolution
could reveal in the near future the presence of such very large-scale features.

5. 2nd order perturbations and the integrated effect

Let us consider an Einstein-de Sitter universe as background, perturbation theory up to the


2nd order for vanishing pressure gives the following value for the density fluctuation (Peebles
1980)

62 S 2 1- - 1 ..
-=6+a(-6
a 7 +-V6·V..I.+-..I.··..I.,")
6 'I' 126'1',1,'1' t:..,p = 66 (3)
where all the quantities in the r.h.s. of both equations are taken at the present time (i.e.
6 = (1 + z,. )6.. 6r is the density fluctuation at recombination), the scale factor aCt) is normalized
to the present time ao = 1 and we take units such that c = 87/"G = 1 and the homon distance
at the present time dHO = 3'0 = 1. Therefore, the integrated effect -as given by equation (2)-
yields to (Martinel-Gonlalel, Sanl and Silk 1991)

( t:..T)
T IGE
= 2/° da1l'(z(a»
r
(4)
50

aif=6- -
8 (62 ) =-6 - - 1 ..
30 2 +V6·V4J+-4Jij4J,·J (5)
8a a 7 21 '
where n is the direction of observation. Preliminary calculations with the 62-term seems to
indicate that the integrated effect dominates over the linear contributions (Sachs-Wolfe and
Doppler terms) at intermediate angular seales of ~ 10 if one assumes a Gaussian correlation for
the matter density. We are currently checking this calculation, the next step will be to apply
this effect to CDM and HDM scenarios.
Finally, we would like to emphasise the possible relevance of this effect because it could be
competitive with other linear and non-linear effects at least on certain angular seales.

References

Dressler, A., 1988, Ap. 1.329, 519.


Dyer, C.C., 1976, Mon. Not. R. ast. Soc., 175, 429.
Dyer, C.C. " Ip, P.S.S., 1988, Mon. Not. R. ast. Soc., 235, 895.
van Kampen, E. " Martinez-Gonzalez, 1991a, lInd Rencontres de Blois: Pbysical Cosmology,
Pergamon Press, in press.
van Kampen, E. " Martinez-Gonzalez, 1991b, submitted to Mon. Not. R. ast. Soc ..
Lynden-Bell, D., Faber, S.M., Burstein, D., Davies, R.L., Dressler, A., Terlevich, R. " Wegner,
G., 1988, Ap. I., 326, 19.
Martinel-Gonzalel, E. " Sanl, J.L., 1990, Mon. Not. R. ast. Soc., 247, 473.
Martinel-Gonlalel, E., Sanl, J.L. " Silk, J., 1990, Ap. 1. Lett., 355, L5.
Martinez-Gonzalez, E., Sanz, J.L. " Silk, J., 1991, in preparation.
Nottale, L., 1984, Mon. Not. R. ast. Soc., 206, 713.
Peebles, P.J.E., 1980, The Large Seale Structure of the Universe, Princeton: Princeton Univ.
Press.
Raychaudhury, S., 1989, Nature, 342, 251.
Rees, M.J. " Seiama, D.W., 1968, Nature, 517, 611.
Scaramella, R., Baiesi-Pillastrini, G., Chincarini, G., Vettolani, G. " Zamorani, G., 1989,
Nature, 338, 562.
Thompson, K.L. " Vishniae, E.T., 1987, Ap. 1.,313, 517.
51

DISCUSSION:

Smoot: It would help to know where to look for effects because it is very difficult to
measure a large part of the sky with high sensitivity (= long observation times). H there
were good coordinates we could choose these areas to observe. We were planning now to
look at some convenient blank sky and a few selected X-ray clusters to look for the Sunyaev-
Zel'dovich effect. I note that the thin shell (compensated void) gives a pattern on the sky
in ¥- as collapsing domain walls.
The Great Attractor and Shapley-Ames concentration are unfortunately near the Galac-
tic plane and thus have a large foreground contamination. Other candidates nearer the
Galactic poles would be most welcome.
Sanz: Following your comment, it would be very interesting to look for regions devoid of
galaxies with scale 75 + 100h-1 Mpc, because in this case ¥-
~ -(0.4 + 1) X 10- 5 •

Rees: The effect you discussed is smaller, by a factor of order the ratio of cluster size to
Hubble radius, than the Sachs-Wolfe perturbation that would be produced by the precursor
of a shnilar cluster straddling the last scattering surface. Therefore, unless the Shapley
Concentration is unique, there should be some rare fluctuations of order 10- 4 due to the
Sachs-Wolfe effect.
It is nevertheless perhaps easier for observers to detect a small ¥-
if they know where
to look. So it is worth looking towards conspicuous superclusters, in the hope of detecting
the effect you've measured even if it is of order 10-6 • One would, however, need multi-
frequency measurements in order to separate out the expected effect due to discrete sources,
the Sunyaev-Zel'dovich effect, etc.

Sanz: I agree with you except for the angular scales involved. In the case of a Great
Attractor (like the Lynden-Bell et al. concentration) or supergreat attractor (like the
Shapley concentration) near our cosmic photosphere, the angular scales are ~ 10- arcmin,
therefore the fluctuations on ¥-would be smeared out by the beam width used in current
observations with a large sky coverage (Relict, COBE, Lubin et al.).
Starobinsky: My comment is similar to that of Prof. Lynden-Bell. It is worthwhile to
remind the audience that this integrated effect was first considered by M. Rees and D.
Sciama in 1968 - more than 20 years ago. Typically, it gives a contribution less than 3.10- 6
to the total ¥- for the CDM mode with bp ~ 2.
Sanz: I agree that the first estimation of the influence on an isolated lump (represented
by a Swiss-Cheese model) on the CMB is due to Rees and Sciama (1968). Except for
the qualitative approach of that paper, other authors have followed a straightforward but
lengthy calculation based on general relativity. I want to remark that our gravitational
potential approach represents a generalisation that applies to any non-linear density field
producing a linear potential. In particular, we have also applied it to other non-compensated
models representing an isolated structure. My final remark concerns the possibility to
calculate in a clear way what is the level of fluctuation associated to this "integrated effect"
52

for some models of galaxy formation (BM, CDM, HDM, etc.). We are currently involved
in such calculations to see the possible contribution on¥ as compared to other effects.
A crude estimate seems to indicate that the integrated effect could be the dominant one
on arcmin scales, but one needs to work out explicitly any particular model to obtain the
concrete ¥.
QUANTUM COSMOLOGY AND THE COSMOLOGICAL CONSTANT

IAN MOSS
Physics Department
University of Newcastle upon Tyne, U.K.

It has long been an axiom of mine that the little things are infinitely the most important.
Sir Arthur Conan Doyle (A Case ofIdentity)

ABSTRACT. The origin of our universe and the small scale structure of spacetime lie deep
in the realm of quantum theory, yet they seem to relate to the large scale features of the
universe. The value of the cosmological constant is probably zero, but a small value related
to a light particle mass is possible.

1. Introduction
The inflationary story of the beginning of our universe tells how a period of rapid expansion
can lead to the formation of matter and structure such as we see today. This story does not
tell about the period before inflation, how likely is inflation and how the universe began.
Quantum cosmology attempts to answer these questions using the theory of quantum me-
chanics. The resulting ideas relate to observations of the present day universe in a number
of ways, in particular to the density parameter n, the cosmological constant A and the
amplitude of density perturbations 6:
• n= 1 from the quantum origin of the universe;
• A = 0 from small scale flutuations in space;
• 6 = canst from quantum density fluctuations.
In this article the main concern is with the cosmological constant, and how small scale
fluctuations can affect it. The cosmological constant measures the extent to which empty
space generates curvature. In the standard model of elementary particles the vacuum is
active, can carry energy and undergo phase transtions. The cosmological limits on A imply
that an unexplained cancellation has taken place in the present day phase of the universe,
leaving a value which is very small compared to the values expected from particle physics.
This indicates the presence of a new symmetry principal or a new dynamical effect. The
view presented here is that the cosmological constant is itself a quantum observable whose
expectation value has been driven towards zero by a mechanism proposed by Hawking
(1984) and Coleman (1988).
The small-scale structure of spacetime is still uncharted territory and variations on the
route to A = 0 are possible. In particular, vacuum fluctuations playa larger role in fixing
the value of A than was expected and a small, but non-zero, value of A may be related to
the lightest particle mass. Another new twist to A comes from looking at negative values,
with the prediction that negative cosmological constants are quantised, with values related
to the electron charge e by A = -3e 2/GN2.

53
T. Shanks et al. (em.). Observational Tests of Cosmological Inflation. 53-61.
Ie 1991 Kluwer Academic Publishers.
54

2. Quantum tunnelling

The quantum story begins with a quantum cosmological principle. One possibility is to sup-
pose that the universe appears from nothing by a process of quantum tunnelling (Vilenkin
1982). This picture has an advantage that it can be analysed in some detail. Although
fundamentally a quantum process, the tunnelling of the universe has an underlying classical
interpretation. The universe begins with a Euclidean stage, a state where time and space
are locally indestinguishable. The metric of spacetime in a Euclidean stage is positive in
all directions. (This signature will be denoted by (4,0).)
The Euclidean stage of the universe merges smoothly with a de Sitter, or inflationary,
stage with the familiar distinction of time and space. This inflationary stage leads on to
the standard hot big bang model of the early universe. The scale of spatial curvature in
this model need never approach Planck values ( the Planck length'" 1O-33 cm and energy
1019GeV) where we loose confidence in General relativity or quantum mechanics.
The quantum tunneling picture represents one choice of initial conditions for the infla-
tionary universe. There are alternative proposals. The original inflationary model of Guth
(1981) began with a pre-inflationary hot big bang, and breaks down at the Planck energy
or it is singular. There are also chaotic inflationary models where the pre-inflationary era
is dominated by coherent matter fields. There have been some attempts to quantify the
likelyhood of predictions such as n = 1 in these models (Belinsky et al. 1985), but they
ultimately depend upon unknown conditions at the Planck energy.
2.1 BARRIER PENETRATION
Quantum tunnelling phenomena are widespread in the physical world, being responsible
for many chemical and nuclear reactions. In order to see what is involved, let us consider the
penetration of a potential barrier by a one dimensional particle. The quantum mechanical
amplitude to go from point x to x', (x', t' I x, t), can be defined by a sum over histories.
These are particle paths, weighted by amplitudes eiS , where S is the action.
If a wave packet '¢ is prepared with a mean energy less than the barrier height, then the
classical motion of this packet is reflected from the barrier at a point x where the velocity
vanishes. However, in quantum theory there is an amplitude for penetration of the barrier.
It is well known that this can be calculated semi-classically by using an imaginary time
coordinate for the tunnelling solution, but what does this really mean? The answer can
be seen by introducing an intrinsic time parameter r, related to t by a lapse function N,
dt = N(r)dr. The sum over histories is really a sum over paths (x(r),N(r». The variables
x and N can be moved into the complex plane using Cauchy's theorem, like any other
contour integral. The tunnelling amplitude is dominated by a complex saddle point, and
the action has real and imaginary parts S = SL + if. Therefore the amplitude becomes

The term A is a Gaussian pre-factor. The next term is the tunnelling rate, independent
of x, and the final term represents the motion beyond the barrier. The tunnelling path is
complex and does not correspond to a classically allowed particle motion. On the other
hand, the path satisfies the equations of motion and has a classical interpretation. The
x coordinate can be taken to be real, and N real outside the barrier, complex inside.
The section outside the barrier represents a classical path and the one inside represents
55

tunnelling through the barrier. These paths can join smoothly only at points where x/N
vanishes. This happens at the bounce points if.
2.2 QUANTUM GRAVITY
The barrier penetration example is nothing more than ordinary quantum mechanics
but a similar set-up can be constructed for quantum gravity. If the length scales in our
problem are always much larger than the Planck length then it would not be absurd to use
the Einstein action for gravity. The amplitudes would then be for spatial geometry g and
matter fields ~,
(g',~' I g,~)
There is no time in gravitational amplitudes appart from intrinsic time. A sum over histories
is a sum over spacetime geometry and matter field configurations.
The creation of the universe by quantum tunnelling is realised by placing "nothing" in
the initial slot. The sum over histories is then a sum over manifolds without an 'initial'
boundary. This amplitude is called the Hartle-Hawking wave function (Hartle and Hawking
1983). The square ofthe modulus of this wave function can be taken to represent the relative
probability of different geometries and matter contents of the universe.
A simple example can be constructed for the de Sitter universe, with a cosmological
constant A but no matter. Take the spatial geometry g to be a sphere of radius a. The
classical paths for a move in the potential shown in fig. 1. They bounce off the barrier at
ii, where ii 2 = 3/A. The corresponding geometry is de Sitter space, with a bounce at the
minimal surface.

Figure 1. The potential for a and the mixed signature tunnelling solution
The quantum amplitudes tunnel under the barrier, where they can be described using
an imaginary lapse function N (Moss 1986). This results in a Euclidean metric for a < ii,
where the solutions are isometric to sections of four dimensional spheres. The amplitude to
tunnel from nothing is therefore dominated by the complex solution made from two parts,
a hemisphere and part of a de Sitter spacetime, as shown in fig. 1.
The action of this tunnelling solution is not just a function of the geometry because it
depends also on the sign of the lapse function N and the signature of the spatial volume.
There are therefore four possible contributions to the amplitude from this one solution.
Two of these represent a reversal of the arrow of time. The remaining two give alternative
amplitudes, one favored by Vilenkin and one by Hawking. The latter has
56

This amplitude is weighted by the Gaussian correction term A and the exponential of the
Euclidean action I = -1rii 2 /G. The final term represents the classical de Sitter universe,
which can be said to begin at a = ii, where real time commences.
This example shows how a quantum amplitude can be approximated by complex solu-
tions to the field equations. The approximation will hold good if the Gaussian term A can
be calculated and the higher order terms can be shown to be small. In general, the things
to look out for are,
1. Solutions of the Einstein equations Gab = 81rG Tab;
2. Complex metric components;
or 2a. Real, mixed signature metric components, with surfaces of vanishing extrinsic cur-
vature K ab = 0 between different signatures;
3. Sections with real metrics 9 and matter fields 4>;
4. An action S = SL + iI, where SL(9,4» satisfies the Hamilton-Jacobi equations (for a
real universe), and I weights the different saddle points.
These solutions will be called complex instantons. The possibilty (2a) is included be-
cause, in many cases, complex cordinate transformations can be used to make the metric
components real, like in the de Sitter example above.
2.3 THE DENSITY PARAMETER
Inflation is often said to drive the density parameter of the universe towards the value
n = 1. In order to quantify the likelyhood of this prediction it is necessary to have a
quantitative model of the initial conditions of the universe. The quantum cosmological
approach makes this possible.
Inflationary models of the early universe are driven by the potential energy V( 4» of a
scalar field 4>, which acts in the same way as a cosmologicaJ constant,

A = 81rGV(4)).

The cosmological term rapidly drives the universe towards homogeneity and spatial flatness.
This subsection is concerned with flatness, therefore homogeneity will be assumed.
There are many complex instantons which have inflationary stages, beginning with a
signature (4,0) and evolving to have a signature (3,1). They can be parameterised by the
value (fi of 4> at the bounce point. Beyond a value (fil, both a and 4> are almost real. The
square of the wave function for these instantons is shown in fig. 2 (for a quadratic potential).
The solid line represents the exponential term,

originally found by Hawking and Moss (1982), for the nucleation rate of the universe.
The dashed lines include the Gaussian term A (Moss and Poletti 1990), which becomes
increasingly important as the energy density reaches the Planck scale. The two curves are
for models with different particle contents (Poletti 1990). The lower curve, characteristic
of supersymmetric theories, is preferable for its decreasing amplitude at large values of the
energy density.
57

_-------- --
,:'

...

Figure 2. The probability distribution of inflationary models

Values of n today between 0.2 and 1.0 correspond to values of 4> in the shaded strip in
the diagram. The width of this strip is a fraction of order fJ of the range shown, where fJ is
the scale free amplitude of density perturbations generated during inflation. The value of
fJ is limited by the level of cosmic microwave background anisotropy to be less than 10- 4 •

3. Wormholes and the cosmological constant

It has alwready been remarked that in quantum theory the vacuum has dynamical prop-
erties which can affect the cosmological constant. The simplest of these are due to the
presence of virtual particles, or vacuum fluctuations. Less well understood are the effects
of quantum fluctuations in the spacetime continuum itself.
Attempts have been made to describe small scale fluctuations in spacetime using quan-
tum wormholes (Hawking 1988, Coleman 1988). These are solutions of the Einstein equa-
tions which join large scale regions by much smaller scale throats,

Figure 3. A wormhole solution

For most types of matter, classical wormhole solutions are forbidden by arguments similar
to the singularity theorems. However, they fit naturally into the class of complex instantons
which play an important part in quantum theory. A complex instanton may involve many
wormholes linking various large scale instantons, as shown in fig,' 4.
58

Figure 4. A complex instanton with wormholes


Amplitudes with many wormholes can be simplified by replacing the wormholes with
pointlike interactions. These can be absorbed by changing the constants of the theory.
This has a profound effect, because it means that none of the fundamental quantities of
physics are fixed ab initio, being in reallity quantum averages. Amplitudes have to be
averaged over wormhole parameters a and disconnected field components.
It is possible to attatch four dimensional spheres to any complex instanton by using
wormholes. The result is to weight the amplitude with factors of
Ae- I = Ae3 ,../G(a)A(a)
for each sphere. Averaging over the a will drive
min
GA -+ a G(a)A(a)

From this, Coleman deduces that A = 0, but notice that there is an assumption here that
A(a) can be zero, but not G(a).
As a matter of fact, Dowker and Laflamme (1990) have shown recently that the worm-
holes contribute directly to terms in the action involving the square of the Weyl tensor.
Contributions to A come indirectly from virtual particle pairs. These effects are tabulated
for various constants of a typical model in table 1.
TABLE 1. Fundamental constants and their quantum corrections
constant symbol 81r 2 X correction
cosmological A(a) 1rGm4Iog~ m 2/ p,2 ~
Newton G(a) (41r /3)(1 - 6()G 2m 2 Iog m 2/ p,2
Weyl-Weyl (1/4801Iog( m2 / p,2)
Rkci-Ricci 19(al
,(a (1/288)(1-6() log(m 2/p,2)
Euler 6(a _(1r2 /45)log(m2 /p,2)
mass 3mAlog(m2/p,2)
charge
self
";~~l
A(a
0
18A 2 1og( m2 / p,2)
curvature (a) A(l - 6() log( m 2/ p,2)

The behaviour of A is closely linked to the effects which determine the particle masses.
The masses which remain near to the unification scales will combine to reduce A to the
59

smallest possible value. Other masses are fixed by effects which are not well understood.
There is a possibility of a residual A of around Gm4log( m 2 /1-'2), where m is the mass of
the lightest particle. A mass of the order 1O-oeV would correspond to a A of roughly
(50 Km 8- 1 Mpc-1 )-2, contributing the same curvature as a critical matter density. Such
small particle masses are no harder to explain than the mass of the electron, which is also
very small compared to the scales of unification of the fundamental forces.
The quantum effects which drive A towards zero can only work if A is positive. If A is
negative, then the amplitudes will be dominated by a different set of complex instantons. It
is possible to extend the wormhole arguments to this case by using metrics with a negative
signature (0,4).
There is a set of instantons for the case of negative A which has particularly interesting
properties (Mellor and Moss 1989). This set share some of the properties of magnetic
monopoles. Like magnetic monopoles, the photon becomes massive unless the electron
charge e is quantised (Moss 1990) and satisfies ea = N, where a is the Anti-de Sitter space
radius and N is a natural number. In terms of A,

A small, negative cosmological constant can be consistent with observation, but N would
have to be larger than 1060. If A is positive, then there is a similar set of instantons, but the
quantisation condition becomes eP = N, where P is the monopole charge. This condition
is the same as the one discovered by Dirac (1931).

4. References
Belinski, V.A., Grishchuk, L.P., Khalatnikov, I.M. and Zeldovich, Ya.B. (1985), Phys.
Letts. 155B 232.
Coleman, S. (1988), Nucl. Phys. B307867.
Dirac, P.A.M. (1931), Proc. Roy. Soc. (London) Al33 10.
Dowker, F. and Laflamme, R. (1990), Cambridge Preprint.
Guth, A. (1981), Phys. Rev. D23347.
Hartle, J.B. and Hawking, S.W. (1983), Phys. Rev. D28 2960.
Hawking, S.W. and Moss, I.G. (1982), Phys. Letts. 110B 35.
Hawking, S.W. (1984), Phys. Letts. B134403.
Hawking, S.W. (1988), Phys. Rev. D37904.
Mellor, F.A. and Moss, I.G. (1989), Class. Quantum Grav. 6 1379.
Moss, I.G. (1986),in 'Proceedings of the fourth Marcel Grossmann meeting', R. Ruffini
(ed.), Elsevier Science Publishers B.V.
Moss, I.G. and Poletti, S.J. (1990), Nucl. Phys. B341 155.
Moss, I. G. (1990), 'Gravitational instantons and quantisation of the cosmological constant',
Newcastle Preprint.
Poletti, S.J. (1990), Phys. Letts. 249249
Vilenkin, A. (1982), Phys. Letts. 117B 25.
60

DISCUSSION:

O. LAHAV Does your work suggest that reducing the vacuum energy (Le. the cosmological
constant) by 120 orders of magnitude to the present critical density is as natural as
reducing it to a value of zero?
I.G. MOSS I tried to show that a very small, but nonzero, value ofthe cosmological constant
is not unnatural when related to the existence of a light particle. The origin of particle
masses remains a mystery, but I think that relating two unknowns is progress of a sort.
DEMARQUE Does your theory imply possible variations in Newton's constant G?
I.G. MOSS These wormhole models predict fixed values for all ofthe fundamental constants
of nature. This is because the parameters which where left behind by the wormholes are
themselves constants. However, some physical processes can lead to time dependent
contributions to the cosmological constant and to Newton's constant. The inflationary
universe depends on the fact that the vacuum energy of a high temperature phase of
matter contributes to A. Given appropriate couplings, the high temperature phase can
also feel a different value of G.
M. REES H A is non-zero at the present time, would you expect it to be constant, or to
be declining as the universe expands?
I.G. MOSS The simple answer is that A would be constant, for the reasons given in answer
to the previous question. The more complicated answer is that this depends on how the
theory is renormalised. In quantum theory there is a parameter called the renormalisation
scale, which I called p.. This enters in a subtle way because particle physicists like to
work with theories in which p. can be absorbed into other parameters. How this is
done depends to some extent on how these parameters are defined. H measurements are
performed at different energies, the values of masses and charges may change. A similar
variation in A may occur.
G. BORNER What are the arguments that show that the symmetric instantons are really
the most important contribution to the functional integral?
I.G. MOSS To my knowledge there is no proof of this assertion and it would be an interesting
project to find the answer. The corresponding problem in scalar field theory was solved
by Coleman in his work on the fate of the false vacuum, where he demonstrated that
spherically symmetric solutions have the least action. For the gravitational case, the four
dimensional sphere has the most negative action of the known solutions to pure gravity
with a cosmolgical constant.
F. HOYLE Refering to your imaginary action: when I tried this in a more usual quantum
situation I found the sum of the probabilities over possible states of the system was not
conserved. Can you get around this difficulty in a cosmological situation?
61

I.G. MOSS I don't believe that anyone has yet come up with a convincing analogue in
quantum cosmology to the conservation of probability. The first step would be a universal
definition of time! In the semi-c1assicallimit the situation improves somewhat. The real
part of the action can be used to define a real time and the usual Schrodinger equation
can be recovered. There would be some very small corrections due to the imaginary
parts of the action, and these would lead to some loss of unitarity. This may be related
to what Stephen Hawking calls a 'loss of quantum coherence', but I will have to think
about this some more.
LESSONS FROM INFLATION AND COLD DARK MATTER

P. J. E. PEEBLES
IMtitute For Advanced Study and
Princeton Univer6ity
Princeton NJ 08544.

ABSTRACT. The progress of research in cosmology has been greatly stimulated by the flow of
ideas from the inflation scenario and its simplest implementation in physical cosmology, the cold
dark matter theory. I present my list of the major lessons we have learned from the long running
debate over the observational viability of this theory.

1 Introduction

Imagine explaining the highlights of cosmology since 1970 to a colleague who has just wak-
ened up from a twenty years nap. We could present a wonderful list of such observational
advances as the precision measurement of the cosmic background radiation spectrum; the
firming up of the standard model for the origin of the light elements; the extension of the
radio galaxy I1ubble diagram into the infrared and to redshifts well above unity; the discov-
eries of the great attractor and great wall. Our cosmologist might be struck by the extent
to which these advances were foreshadowed by pre-1970 cosmology. For example, Shapley's
group was surveying superclusters in the 1930s. The small scatter in the K-band Hubble
diagram for high redshift radio galaxies is startling, but we do have a precedent in the
small scatter in the Hubble diagram for first ranked cluster members. Direct observations
of cosmic evolution of galaxies is new (the Butcher-OemIer effect dates to the late 1970s),
but most of us in 1970 believed that the universe is evolving and so we had to believe that
galaxies evolve too.
The obvious lesson from the observational side would hardly prepare our cosmologist for
what has happened in theoretical cosmology. We could present a host of remarkable new
concepts, and the life histories of perhaps a half dozen major ideas for the origin of galaxies
and the large~scale structure of the universe. Each of these theories for structure formation
has tended to fade from public attention, sometimes because of observational problems,
sometimes because of the press of still newer concepts. Although this flourishing of ideas is
new, their life cycles might not seem so strange to our cosmologist, who would have a fresher
memory of the debate over the relative merits of the steady state and Friedmann-Lemaitre
cosmologies. The cycles are just faster now.
The debate over the steady state theory left lessons oflasting value to cosmology, and I

63
T. Shanks et al. (eds.), Observational Tests o/Cosmologicallnflation, 63-73.
© 1991 Kluwer Academic Publishers.
64

think we could explain to our recently awakend colleague why the same is true of the new
ideas in cosmology, whatever their fates. A general lesson is that cosmology by ansatz is a
chancy business. It can succeed; Einstein started cosmology with his assumption that the
universe is homogeneous, an observationally doubtful proposition at the time. However,
the success rate overall has been considerably lower than in particle physics. I suspect
this is because cosmology and extragalactic astronomy are a good deal more complicated,
in the sense that a much greater variety of things can and do happen. We have no real
alternative to theoretical cosmology by bold hypothesis, because there is no way we will
be able to deduce from the observations what happened at high redshifts. But we will do
well to remind ourselves and our colleagues in other fields that a low success rate is not
unexpected.

2 Lessons from the Cold Dark Matter Theory

The brightest of the shooting stars has been the cold dark matter (CDM) theory. It was
devised as a particularly simple case (Peebles 1982); it was improved with the introduction
ofthe biasing concept (Davis et al. 1985), and in the minds of many it became the standard
model within inflation because it fits so well with the simplest versions of that scenario.
CDM never was a standard model in the particle physicist's sense; it was and is a bold
trial that is being explored by methods that may prove to be too simple even for this
computationally convenient theory. That was carefully explained by those working on the
theory, but the message was not fully appreciated in the broader community. The result
is that CDM for a long time was overvalued, and now in reaction might be undervalued.
It is too soon to decide whether some recognizable variant of CDM might prove to be on
the right track. However, the theory certainly has highlighted some fundamental issues
that will have to be addressed in any viable cosmology. Here is my selection of the most
pressing.

2.1 THE OBSERVATIONAL PROBLEMS WITH BIASING

A central issue for cosmology is, if the mean mass density is equal to the critical Einstein-de
Sitter value, where is the mass? The two simple possibilities are that it is concentrated
with the galaxies, or it is in the voids between the galaxy concentrations. My impression
is that both are problematic in light of what is known about galaxies.
In standard CDM the seeds for galaxies were everywhere, because the seeds are Gaussian
fluctuations in the primeval mass distribution. The biasing idea is that the seeds in voids
germinated too late to turn into recognizable galaxies. The problem with this is that it
surely would follow that those galaxies that did form in the voids are irregular, or dwarfs,
or otherwise bear the stigmata of an exceedingly unhealthy youth. But all the evidence I
have heard indicates that dwarfs cluster with giants (eg. Binggeli 1989).
Our immediate neighborhood is a good example (Peebles 1989a). In a cube 800 km s-I
on a side, with us at the center, some 200 galaxies are known, of which perhaps a dozen
are large galaxies. So many dwarfs are catalogued because at this shallow distance galaxies
are visible well down into the luminosity function. Most of the galaxies are concentrated
65

in a sheet", 200 km 5- 1 wide. Above and below there are some distinctly empty regions.
Since the cube contains 200 seen galaxies, the biasing picture would say there are literally
hundreds of failed seeds in the voids in the cube. Given such odds, the few that made it
to galaxies in the voids really ought to look peculiar. But one of the more isolated of the
very nearby galaxies, NGC 6946, looks like a normal if somewhat gas rich spiral (Sandage
and Bedke 1988). The box does contain unusual objects. The "dark galaxy" of Carignan
and Freeman (1988) is just the thing for a nearly failed seed. But it is in the plane of the
Local Supercluster, with most of the other galaxies.
To summarize, the biasing picture would say that in our neighborhood there are hun-
dreds of failed seeds of galaxies, yet the candidates for seeds that nearly failed are found
not in the voids but in the concentrations of seeds that did produce galaxies. One might
conclude that the suppression of germination in voids was extremely efficient. But then
why do we see the occasional normal-looking galaxy in the voids?

2.2 HAS GRAVITY EMPTIED THE VOIDS? A TEST IN OUR NEIGHBORHOOD

Within CDM an alternative to biasing is the idea that gravity pulled the seeds out of the
voids. Of course gravity would have pulled the cold dark matter with the seeds. Our local
galaxy neighborhood, the Local Group and its immediate neighbors, again gives us a useful
test. My conclusion, from the model to reviewed here for the formation of the Local Group,
is that the idea make sense, but only if the mean mass density is '" 10% of the critical
Einstein-de Sitter value.
A numerical solution for the motions of the galaxies in our neigborhood requires a
computation of orbits, which is easy, and initial conditions consistent with the idea that
the material started out moving with the general expansion of the universe, which can
be a little more difficult to deal with. The approach I shall review here assumes we can
assign an effective central position to the mass concentrated around each galaxy. We want
the gravitational effect of this mass on the motions of the central positions of neighboring
mass concentrations to be the same as if there were a point mass at the central position.
That is, we are replacing the continuous mass distribution with a set of point masses m; at
positions z;(t). The initial condition is that the point mass positions at high redshift are
such that the masses move with the general expansion of the universe. A solution consistent
with this condition gives a sensible approximation to the early universe, which is nearly
homogeneous, and also to recent epochs, when the mass within the Local Group presumably
is concentrated around the Milky Way and M3l. But the best justification for this model
is that the numerical solutions successfully predict the redshifts of dwarf members of the
Local Group, as will now be discussed.
The trick for finding orbits consistent with the boundary condition goes as follows. Let
us neglect the expansion of the universe for a moment. Then the action for our set of
particles interacting through the potential V is

(1)

Under the infinitesimal change of orbit z; ---+ z; + 6z;(t) for particle i, the action changes
66

by the amount

5S = 1
o
to
dtb· [ -m·--'
' 'dt 2
av] + [m·5z·-'
tflz· - -
aZj "dt
dz.]t o
0 •
(2)

One usually fixes the orbit at the end points, so 5zj = 0 at t = 0 and t = to. Then the
last term vanishes, and 5S = 0 for orbits that satisfy the equations of motion, which is the
action principle. But note that the action principle also applies if 5zj is free at t = 0 and
the initial velocity vanishes, dzi/ dt = 0 at t = o. This latter condition reminds us of the
gravitational instability picture, in which structures grow by gravity out of small departures
from homogeneous expansion. The generalization to the expanding universe case is easy
(Peebles 1990).
In the expanding universe case one uses position coordinates that are expanding with
the general expansion of the universe. The initial condition on peculiar velocities here
implies that the positions of the mass centers at high redshifts are such as to make the
relative gravitational accelerations of the particles agree with a homogeneous universe. IT
they did not, the peculiar velocities would diverge at redshift z --> 00.
In the numerical application of the action method the orbits are modeled as functions of
time with free parameters, Zj(t, a), so the action can be numerically evaluated as a function
of the parameters a. Then one can ask the computer to seek a stationary point of S by
walking down the gradient of as/aa, or by inverting a matrix. The check that the method
really works is that one can use the numerical action solution to set initial positions and
velocities at high redshift for a conventional N-body integration forward in time that does
bring the particles back close to where they started (eg. Peebles 1990, Fig. 3).
The point of this exercise is not to compute orbits - conventional methods do a much
better job - but to find initial conditions consistent with the gravitational instability
picture. I have not seen any other practical way to do it in nontrivial arrangements of the
matter.
The solutions shown in Figure 1 apply to eight members ofthe Local Group, along with
the nearby Maffei and Sculptor groups modeled as point masses each with the mass of the
Local Group. The mass ratios of Local Group members are fixed ahead of time, and the
mass scale is adjusted to give the observed velocity of approach of M3l.
The action solution is not unique; there can be many parameter choices that make the
action an extremum. Trials suggest only NGC 6822 has solutions with seriously different
orbits, in one of which it is falling toward us (in expanding coordinates) for the first time,
while in the other it has passed us and now is nearly at rest. In the former solution the
velocity of NGC 6822 is much too large, while the latter is not far from the observations.
Figure 2 shows observed and predicted velocities of the dwarf members of the Local
Group, with the passing orbit solution for NGC 6822. This figure shows how the predictions
depend on the choice of Hubble's constant for given density parameter, n = 0.1, with a
cosmological constant added to make the universe cosmologically flat. The Einstein-de
Sitter model with n = 1 gives quite similar results, because once the expansion timescale
is fixed the motions within the Local Group are not much affected by the cosmology. I
have not checked but expect that the same is true of a low density cosmology with no
cosmological constant. (The insensitivity to the cosmology is not obvious from the orbits
67

~ (-)~
. . <:::, r
/""

~..."
\scu'Pt

Figure 1: Orbits for nearby galaxies. In the left solution n = 0.1, in the right n = 1.
The orbits in expanding coordinates commence at high redshift at the names. The present
positions are the boxes.

shown in Fig. 1 because the maps show motions relative to the center of mass of the three
groups.)
The observed and predicted velocities in Figure 2 are quite similar. It will be recalled
that the mass scale is set to get the right velocity for M31j this galaxy of course is not
shown in the test in Figure 2. The only other adjustments are the choice of one of the two
orbits for NCC 6822, and the choice of Hubble's constant, which we see is in a reasonable
range. I conclude we have a reasonable case for this solution as a useful approximation to
the way the Local Group really formed. The solution has several interesting features.
The galaxy masses derived from the model yield a mass to light ratio of about 100 solar
units. If this ratio applied to all galaxies it would translate into a mean mass density about
10% of critical. This is just one of the familiar set of results from dynamics on scales less
than about 10 megaparsecs that so consistently indicate n '" 0.1. It is worth noting that
in the Local Group the crossing time is larger than the Hubble time, so it is difficult to see
how dynamical relaxation processes could have ejected dark matter from the Local Group.
Rather, the indication is that the Local Group was assembled out of material with a low
mass per galaxy relative to the Einstein-de Sitter model. This conclusion is hardly new, of
coursej it was the motivation for the biasing scenario. But we saw that the biasing scenario
does not make much sense when applied to our neighborhood.
Another aspect of the numerical action solution is that if n = 1 (the right hand side
of Fig. 1) the two nearby groups are moving away from us (from initial positions at the
names to present positions at the boxes). That is because the local mass density is fixed by
adjustment to the motion of M31 relative to us. In the high density n = 1 case this makes
the local density lower than the cosmological mean. Since this would say we are in a hole,
68

11613 WLM N6822 N300 Seo A N3I09

200 x
/

0
X
0

~ 100
{;
{;

E
.>t! 0

~ {;
0
{;

~ 0
X
Jl
I -100
0
{;

8/
/
-200 -100 0 100 200

Observed Velocity km 5·'

Figure 2: Predicted and observed velocities of Local Group members for {} = 0.1 and
Hubble constant H = 50 (triangles) 75 (squares) and 100 km s-1 Mpc- 1.

the solution says we are expanding. Contrary to the idea that gravity emptied the voids,
in this solution we are expanding into them! If {} '" 0.1 the cosmological mean is less than
the local density, which makes a lot more sense, for we are after all in a local concentration
of galaxies. The collapse factor seems to be about right to empty the local voids.
To repeat the question I began with, where is the mass? Maybe the matter is hot, so
it resists clustering on scales less than a few tens of megaparsecs. Maybe there were no
seeds in the voids. Maybe the mass is in a field that cannot cluster, after the fashion of a
cosmological constant. Maybe the mass is hidden in some other artful way. Or maybe the
universe is open.

2.3 LARGE-SCALE STRUCTURE

There are two issues here. The first is whether CDM can account for the rms fluctuations
in the mass distribution on scales larger than about 30 Mpc. The second is whether CDM
or any theory in which large-scale structure forms by hierarchical growth out of Gaussian
primeval mass density fluctuations can account for the tendency of large-scale structures
to have a sheet-like character.
The latter effect was predicted in the adiabatic baryon and adiabatic hot dark matter
pancake pictures (Zel'dovich et al. 1982; Melott, 1983) and the explosion picture (Ostriker
and Cowie 1981). The pancake picture as it stands is problematic because it predicts
that galaxies form after the collapse to proto clusters, while we certainly seem to be in an
old galaxy in a Local Supercluster that is only now collapsing. However, it does show us
how a sheet-like pattern can form. If the velocity field initially has a coherence length,
69

continuity tells us that when orbits start to cross the intersections define a surface, which
is the pancake.
In CDM, and any hierarchical gravitational instability picture, a first generation of
pancakes may have formed at high redshift and then drained away into the first bound
clumps of matter. Might a new generation of pancakes form out of the motion of these
clumps? The argument just given has to be reconsidered, because the velocity field no
longer has a coherence length. This would not matter if the next substantial contribution
to the mass fluctuations were in a bump iI). the power spectrum on a scale substantially
larger than the first generation of pancakes. What is the condition on the power spectrum?
If its logarithmic derivative is steeper than -1 (that is, if the spectrum is modeled as
15kl2 oc kn, where 5k is the Fourier transform ofthe mass distribution, and n < -1), then in
linear perturbation theory the relative velocity increases with increasing smoothing scale.
It has been suggested that in this case the large-scale velocity field might be expected
to produce a second generation of pancakes. But surely what is relevant is the closing
time, which is the ratio of the smoothing length to the relative velocity on the scale of the
smoothing length. The closing time only increases with increasing smoothing scale, which
I would have guessed is necessary for a second generation of pancakes, if n < -3.
By this argument I would not have expected to see second generation pancakes in
the CDM theory. But contrary to my expectation, numerical experiments in CDM show
very striking large-scale pancakes (Park 1990). Here is an interesting challenge: find the
explanation for the pancake effect in CDM, or find what would have to be a very subtle
glitch in the numerical models.
The second issue has to do with the large-scale fluctuations in the mass distribution.
This has been thoroughly discussed (eg. Bertschinger and Juszkiewicz 1988), and recent
work continues to indicate that the mass fluctuations on large scale are greater than pre-
dicted in the standard CDM theory (Saunders et al. 1991). A possible way out within
CDM is to lower the mean mass density, keeping the universe cosmologically flat by adding
a cosmological constant (Efstathiou, Sutherland, and Maddox 1990). This has the immedi-
ate benefit of eliminating the biasing conundrum. In CDM the length scale of the spectrum
of primeval mass density fluctuations scales with the density parameter and Hubble's con-
stant, H, as 1/(OH2). Thus lowering the density parameter pushes the spectrum to larger
scales, which increases the mass fluctuations on large scales, which is another Good Thing.
However, this is not the end of the challenges for CDM. Let us consider next the problem
of assembling the mass in a great cluster of galaxies.

2.4 RICH CLUSTERS OF GALAXIES

Clusters of galaxies occupy a scale intermediate between the large-scale structures just
discussed and the mass concentrations in galaxies. Analyses of cluster formation by analytic
(Peebles, Daly, and Juszkiewicz 1989) and numerical (Frenk et al. 1990) methods both
indicate that in conventional CDM cluster masses are strongly limited. A conclusion from
the numerical work is that in CDM clusters with line of sight velocity dispersions greater
than 1000 km s-1 are exceedingly rare. Since there are many candidate clusters with galaxy
velocity dispersions apparently well above this limit (Teague, Carter, and Gray 1990), and
galaxy velocity dispersions generally agree with what is inferred from X-ray temperatures
70

and from gravitational lensing, the analytic side has high hopes of collecting a bottle of fine
scotch whisky.
The problem for CDM is not relieved by lowering the density parameter. Since biasing
is removed, the mass fluctuation amplitude is increased, which helps. However, the cluster
mass is fixed, so the length belonging to the original unclustered mass now in the cluster
scales as 0- 1 / 3 • The CDM spectrum on these scales is well approximated as a power law,
lc5kl2 ()( len, with n = -0.85. Thus increasing the radius from which the mass is collected
lowers the rms mass fluctuation on the cluster mass scale by the factor

c5M/M ()( 0-(3+n)/6 . (3)

Lowering the mass density to 0 = 0.2 lowers c5 M / M by the factor 0.56, which about cancels
the gain from the removal of biasing.
The issue is of more general interest because in the gravitational instability picture
the masses of the great clusters are assembled out of length scales comparable to those
traditionally used in normalizing the spectrum of primeval mass fluctuations. Thus one
expects to come fairly close to predicting the observed abundance of clusters whatever the
details of the theory. A closer comparison of theory and observation of clusters at high
redshift as well as low, under a reasonable range of adjustment of 0 and n, might yield an
interesting test of the assumption that structure grew out of Gaussian fluctuations in the
mass distribution.

2.5 GALAXIES AT HIGH REDSmFTS

In the conventional CDM theory, galaxies grew by gravitational instability out of scale-
invariant adiabatic Gaussian density fluctuations. This has several advantages. It is simple.
It agrees with the simplest implementations of the inflation scenario. And after decoupling
the mass fluctuation spectrum on small scales has bent over to lc5kl2 '" 1e-3. This means
density fluctuations short ward of the break in the spectrum go nonlinear all at about the
same epoch. This is a great convenience for numerical studies of the evolution of the
mass distribution into the nonlinear regime, because the non-linear fluctuations appear at
a modest redshift. But it has the observational consequence that galaxies are assembled
commencing at modest redshifts. There are differing opinions on the degree of difficulty this
causes in accounting for the observations of high redshift galaxies (Peebles 1989b, White
1989) and quasars (Efstathiou and Rees 1988, Turner 1991).
To our cosmologist from 1970 the fascinating point here surely would be that we have
observations of objects from a time when the density of the universe was some two orders
of magnitude greater than it is now. The absorption line spectra of these objects show
that a well developed intergalactic medium was in place then. It would seem sensible to
take seriously the straightforward interpretation, that protogalaxies existed then, as mass
concentrations. If so, CDM underpredicts the mass fluctuations on small scales.
71

3 Conclusions

The progress of cosmology over the past two decades presents us with a fascinating mix.
The impressive observational advances have made the hot big bang cosmology considerably
more credible, and are offering tantalizing hints to how we might improve the details of
the theory. However, there is relatively little that is revolutionary here, that might have
seemed foreign twenty years ago. That is quite a contrast to theoretical cosmology, which
is full of revolutionary ideas. Will this schizophrenic pattern persist? The detection of a
dark matter particle, or a cosmic string, or the like, would have a wonderfully healing effect.
Failing that, we theorists need not become the drudges of the observers, but we might start
paying them more attention.

Acknowledgements. This work was supported in part at Princeton University by the US National
Science Foundation, and at the Princeton Institute for Advanced Study by the Ambrose Monell
Foundation.
72

References

Bertschinger, E., &: Jusllkiewicll, R., 1988. A6troph1l6. J., 334, L59.
Binggeli, B., 1989. in Large Scale Structure and Motion6 in the Univer6e, p. 47, ed. M. Mellzetti et
al., Kluwer Academic.
Carignan, C., &: Freeman, K. C., 1988. A6troph1l6. J., 332, L33.
Davis, M., Efstathiou, G., Frenk, C. 5., &: White, S. D. M, 1985. A6troph1l6. J., 292, 371.
Efstathiou, G., &: Rees, M. J., 1988. Mon. Not. R. a6tr. Soc, 230, 5.
Efstathiou, G., Sutherland, W. J., &: Maddox, S. J., 1990. Nature, 348, 705.
Frenk, C. 5., White, s. D. M., Efstathiou, G., &: Davis, M., 1990. A6troph1l6. J., 361, 10.
Melott, A. L., 1983. Mon. Not. R. a6tr. Soc, 202, 595.
Ostriker, J. P., &: Cowie, L. L, 1981. A6troph1l6. J., 243, L127.
Park, C., 1990. Mon. Not. R. a6tr. Soc, 242, 59P.
Peebles, P. J. E., 1982. A6troph1l6. J., 263, L1.
Peebles, P. J. E., 1989a. J. R. A6tr. Soc. Canada, 83, 363.
Peebles, P. J. E., 1989b. in The Epoch of Galaz1l Formation, p. 1, ed. Frenk, C. S., Ellis, R. S.,
Shanks, T., Heavens, A. F., &: Peacock, J. A., Kluwer Academic.
Peebles, P. J. E., 1990. A6troph1l6. J., 362, 1.
Peebles, P. J. E., Daly, R. A., &: Juszkiewicz, R., 1989. A6trophY6. J., 347, 563.
Sandage, A., &: Bedke, J., 1988, Atla6 of Galazie6, NASA.
Saunders, W. et al., 1991. Nature, 349, 32.
Teague, P. F., Carter, D., &: Gray, P. M., 1990. A6tron. J. Suppl., 72, 715.
Turner, E. L., 1991. A6tron. J., 101, 5.
White, S. D. M., 1989. in The Epoch of Galaz1l Formation, p. 15, ed. Frenk, C. S., Ellis, R. S.,
Shanks, T., Heavens, A. F., &: Peacock, J. A., Kluwer Academic.
Zel'dovich, Ya. B., Einasto, J., &: Shandarin, S. F., 1982. Nature, 300, 407.
73

DISCUSSION:

Rees: Your ideas on galaxy formation in a baryon-dominated universe require spatial


variations in the baryon-to-photon ratio. H such fluctuations existed, the main problem
with the (otherwise highly attractive) neutrino-dominated model would go away - because
galaxies could then form before the large-scale structure. Would you agree that this provides
a further motivation for urging our particle physics colleagues to come up with a plausible
theory that allows these entropy fluctuations?

Peebles: I agree.

White: I have two questions about dark halos which are concerned with the first and
last parts of your talk. (1). I'm puzzled by the conclusion from your Local Group model,
that in a flat universe most of the local galaxies are required to have moved outwards from
their initial comoving positions. I though that an earlier paper of yours with Adrian Melott
showed good agreement with the velocities of Local Group members in a model where M31
and the Milky Way accreted their halos in an n = 1 universe. (2). While I agree that
"typical" galaxy halos (ve = 300 km/s at r = 200 kpc) can be put into the universe at 1 +
z = 20, when t = 2 X 108 years, the orbital time (and so the collapse time) at this radius
is 4 x 109 years, so I don't see how they can collapse and make their dark matter so early.

Peebles: The numerical action solution and the Melott solution for the relative motion
of Local Group galaxies are quite similar and insensitive to n. The expansion effect I
mentioned in the n = 1 case refers to the relative motions of the Sculptor, Maffei and Local
groups.
I think your estimate of collapse time assumes spherical collapse. My guess is that
previrialisation in real collapse considerably lowers the collapse factor.

Melott: What kind of assumptions do you make in your isocurvature model to create the
kind of rich large-scale structure found in the observed Universe?

Peebles: That has to be an ad hoc adjustment of the assumed spectrum of primordial


density fluctuations, because we do not have an underlying physical theory. That is (was)
the great attraction of the biased cold dark matter theory.
Hoyle: Chairman's remark: This seems to me a good example of the extra freedom that
comes from inverting the time order of the epochs of galaxy formation and of the last
thermalisation of the microwave background. H galaxy formation came ahead of the last
thermalisation the possible models are widened considerably.
The Topology of Galaxy Clustering

PETER COLES
Astronomy Unit, School of Mathematical Sciences
Queen Mary and Westfield College,
Mile End Road,
London El 4NS.

MANOLIS PLIONIS
SISSA-International School for Advanced St'udy
Strada Costiera 11,
Trieste 94014,
Italy

ABSTRACT. We discuss an objective method for quantifying the topology of t.he gala..,,)" <iist.rilmt.ioll
using only projected galaxy counts. The method is a useful complement to fully t.hree-dimcllsional
studies of topology based on the genus by virtue of the enormous projected dat.a sets availahle.
Applying the met.hod to the Lick counts we find no evidence for large-scale non-gaussiall heh<lvionr,
whereas the small-scale distribution is strongly non-gaussian, with a shift in the llwathnll din'dioll.

1 Introduction

Recent large-scale galaxy redshift surveys have revealed a wealth of strllct1ll'e a.bove tha.t
which one might expect to occur in standard models of gala..xy forma.tion (De Lapparent.
Geller & Hllchra 1986; Geller & Huchra 1989; Broadhurst et al. 1990). TIl<' visnal illl-
pression gained from the latest 3-dimensional data sets is that gala..xies are dist.rihut('(1 ill a.
bubbly or cellular pattern, with walls and filaments surrounding large empty Y(litls. at. least
some of which appear to interlock.
Such visual impressions can be seriously misleading, however. The redshift data s<llllpk
rather limited volumes; data from the CfA survey, for example, is available for a few slices
through the spatial distribution. Although one may have a strong visual impression of
a bubbly pattern, one must remember that a slice through a sponge (where voids a.nd
overdensities interlock) would present a similar impression to a slice throngh a hOll('~'(,()lllh
(where they do not). It is important therefore to use objective statistical terIllliqut,s. not
only to quantify the degree to which pattern is actually present in the data. but also t.o test.
models of large-scale structure.
In an important series of papers, Gott and collaborators have developed a. s('t. of tech-
niques that can be used to measure objectively the topology of large-scale strnctnl'<' (Got!
et al. 1989). These methods are not only interesting complements to other st.atistical
studies of galaxy clustering but also provide a ra.ther direct constraint on on(' important.
component of inflationary cosmology, that primordial density fluctua.tions are g<'Il('rat.<'<] hy

75
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 75-79.
© 1991 Kluwer Academic Publishers.
76

quantum fluctuations in a scalar field and are therefore gaussian. In this paper, we review
the strengths and weaknesses of such topological measures in describing galaxy clustering
and then go on to show how to define analogous quantities for projected galaxy data and
mention results obtained for one particular data set-the Lick galaxy catalogue.

2 Measures of Topology in Three Dimensions

An obvious way to study the behaviour of a point pattern is to study the properties of regions
where the local nunber-density of points exceeds some specified threshold level. To look at
the topology of such regions, it is necessary first to smooth the point set thus generating
a continuous random field. The portion of space where this fluctuation field exceeds some
specific density threshold, say 'U, is termed the excursion set at the level It. The excursion
set at an arbitrary u may consist of a number of regions, some of which are disjoint and
some of which are multiply connected, containing holes where the field dips below 'U. A
way of quantifying the topology of excursion sets in three dimensions was suggested in
a cosmological context by Gott, Melott & Dickinson (1986) and has subsequently been
investigated in great depth, the state of the art being well described by Gott et at. (1989)
aud Melott (1990). These authors quantify the topology of excursion sets in terms of the
genus, per unit volume of space, of isodensity contour regions. Strictly speaking, the genus
is a property of the contour surfaces (i.e. the boundaries of the excursion set) and is defincd
iu terms of an integral of the local gaussian curvature over each compact two-dimensional
surface in the set (Weinberg, Gott & Melott 1987). Roughly speaking, the genus of a
snrfacc is the number of handles the surface possesses; a sphere has no handles and has
zero gcnus, a torus has one handle and has unit genus. It is worth remarking, however,
that for technical reasons involving the treatment of boundaries, it has become conventiollal
1l0t to use the genus itself but the quantitity G. which is simply one less than the geuus.
Multiply connected contour surfaces have a positive G s whereas simply connected regions
have negative G s •
Thc great advantage of this formalism is that all gaussian random fields show the salllC
hehaviour of (Gs(v )), where v is a dimensionless form of the density threshold u = (p) +II(T p,
the only depcndence being a cons taut multiplier determined by the power spectrum of the
gaussian field. The curve is symmetric in v so that high and low density regions are
cquivalcnt, is positive at v = 0 and negative for large v. This behaviour has becn likeucd
to a sponge. Two (non-gaussian) alternative scenarios to the (gaussian) sponge arc the
meatball (where the pattern is dominated by isolated dense clusters) and the Swiss dwcs('
(isolated voids, surrounded by high-density regions).
In models where the large-scale structure is generated by the gravitational evolutioll
of initially gaussian density perturbations, such as the standard CDM model, one expects
the small-scale distribution to be highly non-gaussian, whereas smoothing on larger <1.1\(1
larger scales should produce a density field closely resembling the gaussian. Intriguingly
this is what has been found in practice when this method has been applied to real galaxy
clustering data. No significant departures from gaussian topology on large scales have yet
heen detected, whereas there is some evidence for the meatball shift at small scales thought
to he the signature of gravitational evolution (Gott et al. 1989).
77

3 Topology in Two Dimensions

Unfortunately, studies of the topology of the 3-d galaxy distribution are severely hampered
by resolution problems. The galaxies appearing in a volume-limited catalogue are bright
and therefore rather rare objects. When one smooths the data set to recover a continuous
field, one typically has to smooth on a scale greater than the galaxy-galaxy correlation
length and this can drastically alter the underlying statistics. One has to fight hard against
this effect to see any non-gaussian behaviour at all, even in models where the initial density
perturbations are intrinsically non-gaussian. To circumvent this problem we han~ t.ried
an alternative approach to topology, using only projected data (Coles & Plionis 1991).
Although a catalogue of galaxy positions on the celestial sphere contains less information
than a redshift catalogue, it will contain very many more gala.xies (hundreds on thousa.nds
rather than mere thousands). Importantly, this means that the mean angular separation of
galaxies is very much less than the galaxy-galaxy correlation angle so one need not. WOlTY so
much about the smoothing problem mentioned above. The technical details of 0111' a.pproach
can be found in Coles & Plionis (1991); we give an outline here.
We suppose that the celestial sphere can be represented as a plane of area -1" steJ'<ulians.
We can define a two-dimensional density field on this plane and consider its eXClll'SiOll s<'t.s
in the same way as the 3-d case outlined above. A quantity related to the gellus ('QuId he
defined in terms of curvature integrals over the 1-d contours that define the edges of the
excursion set. In fact we choose not to do this. It is more convenient to use a slight.ly difkr-
ent topological quantity, the Euler Poincare (EP) characteristic, of the exclUsion s('( (Colt's
1988). This quantity is defined in terms of the excursion set itself not via its b011lldar)'.
We define the EP characteristic, r, to be the number of disjoint regions in the ('xrlUsion
set minus the number of holes in such regions, The Gauss-Bonnet theorem shows tha.t till'
EP characteristic is equivalent to the genus for sets defined on an infinite flat 2·(\ snrfan'.
This second definition is algorithmically much easier to implement and has a 1l11lltl)('r "I'
advantages of the method favoured by other workers (Coles & Plionis 19!)1). As wi th the
genus the mean value of r can be computed exactly for any gaussian field:

(r) = Avexp(-v 2 /2) (1 )

where the threshold level is again expressed in dimensionless form: 1/. = II + 1m. Note
that this is an odd function of v. An important fact of 2-d topology is t.hat th('r<' is uo
distinct equivalent of a sponge; all closed curves in 2-d have one 'inside' and Oll(' 'o11tsi(\(",
It is, however, still possible to define analogous meatball and Swiss-cheese shifts to those
discussed above (lVIelott et al. 1989).

4 Application to the Lick Catalogue

As the first stage in a large programme of work involving this method, we han' a.pplic(\
our algorithm to the Shane-Wirtanen counts of the Lick gala.xy catalogue (Shau<' ~ Wirta-
nen 1967). The corrected version of the catalogue we use contains approxilllatd~' 810,000
gala.xies, covering about 5 steradians after we remove areas of low galactic latitude to pn~·
vent the systematic effects of galactic obscuration. Because the statistics of a proj<'ct('(\
78

data set are expected to depend rather strongly on the angular resolution one employs we
re-binned the original 10 by 10 arcmin cells into (a) 20 by 20, (b) 50 by 50 and (c) 160 by
160 arcmin cells. We refer to these as high-, medium- and low-resolution cases respectively.
U sing Monte Carlo simulations to assess the statistical errors and possible systematic effects
due to boundary corrections, we were able to demonstrate clear departures from gaussian
behaviour in the high-resolution case (a). The intermediate resolution results can be ade-
quately modelled by a simple transformation of a gaussian field such as a lognormal (Coles
1988; Coles & Jones 1991). In both cases the non-gaussian shift we see is in the meatball
direction, as expected from non-linear gravitational evolution. At low resolution, the results
are consistent with gaussian behaviour, although the statistical errors here are rather large.
Further details and graphical results can be seen in Coles & Plionis (1991).

5 Discussion

We expect the small-scale distribution of matter to be highly non-gaussian even if primordial


density fluctuations were gaussian, simply by virtue of the actioll of non-linear gravity. The
fact that the intermediate scale distribution can be modelled by a local transformation of a
gaussiall field is suggestive that the departure from gaussian statistics may just be due to
the action of weakly nOll-linear gravitational effects, although we cannot conclude this for
certain without detailed simulations. We have much further work to do in this area; the
modelling of projection and selection effects in the available catalogues must be performed
using both simple stochastic models, such as the Voronoi tesselation, and detailed N-body
simulations. We must suspend judgement on the implications of our results until these
simulations have been completed.

References

Broadhurst, T.J., Ellis, R.S., Koo, D.C. & Szalay, A.S., 1990. Nat'ure, 343, 726.
Coles, P., 1988. Mon. Not. R. astr. Soc, 234, 509.
Coles, P. & Jones, B.J.T., 1991. Mon. Not. R. astr. Soc, 248, 1.
Coles, P & Plionis, M., 1991. Mon. Not. R. astr. Soc, 000, 000.
De Lapparent, V., Geller, M..1. & Huchra, J.P., 1986. Astrophys. J., 302, L1.
Geller, M.J. & Huchra, J.P., 1989. Science, 246, 897.
Gott, J .R., Melott, A.L. & Dickinson, M., 1986. Astrophys. J., 306, 341.
Gott, J .R., et al., 1989. Astrophys. J., 340, 625.
I."Iclott, A.L., 1990. Physics Reports, 193, 1.
Melott, A.L. et al., 1989. Astrophys. J., 345, 618.
Shane, C.D. & Wirtanen, C.A., 1967. Pub!. Lick Observatory, 22, 1.
Weinberg, D.H., Gott, J.R. & Melott, A.L., 1987. Astrophys. J., 321, 2.
79

DISCUSSION:

Sahni: Which other topological invariants, apart from the genus, can be used to charac-
terise Gaussian and Non-Gaussian distributions?

Coles: I don't know. Can you think of any?

Melott: I would like to point out that if there is biasing, i.e. if MIL is allowed to vary
V2 permits direct comparison with observation so long as MIL varies monotonically with
scale. VI cannot be used unless one knows the mass associated with each galaxy.

Coles: I don't agree that "VI cannot be used". If there is biasing, VI will contain in-
formation about its form. I would advocate that both VI and V2 be used since they are
not equivalent for a non-gaussian field such as a biased gaussian field can be regarded as
complementary.

Frenk: The amplitude of the genus curve depends, as you mentioned, on the slope of the
power spectrum. Are there any constraints you can place on this spectrum from your data?

Coles: In principle, the answer to your question is yes. In practice this is very difficult
with the Lick data. The reason is that, when the smoothing scale is large enough for the
curves to look gaussian (which they would have to be for such an analysis to be worthwhile)
the scatter in amplitudes in the Monte-Carlo simulations is very large. Furthermore, we
are not convinced that there is a well-defined magnitude limit which rather devalued any
qualitative statements about the amplitude of any inferred 3D spectrum. Looking purely
at the 2D field, our results are consistent with a white-noise spectrum.
CAN NON-GAUSSIAN FLUCTUATIONS FOR
STRUCTURE FORMATION ARISE FROM INFLATION!

D.S. Salopek
NASAl Fermilab Astrophysics Center
P.O. Box 500 MS-209
Batavia, Dlinois, USA 60510

ABSTRACT
Non-Gaussian fluctuations for structure formation may be generated during the inflationary
epoch from the nonlinear interaction of two scalar fields with gravity. Semi-analytical
calculations are given describing nonlinear long wavelength evolution in 3+ 1 dimensions.
Long wavelength fields are governed by a single equation, the separated Hamilton-Jacobi
equation (SHJE). I discuss complete analytic solutions of the SHJE for two scalar fields
with a potential whose logarithm In V(~i) is linear. More complicated potential surfaces
may be approximated by continuously joining various linear In V(~i) potentials. Typically,
non-Gaussian fluctuations arise when one passes over several sharp ridges in the potential
surface. One can input this richer class of initial conditions into N-body codes to see the
effects on the large scale structure in the Universe. The cleanest test of non-Gaussian
fluctuations will hopefully occur in the near future from large angle microwave background
anisotropy experiments.

1. INTRODUCTION
Although inflation predicts that n = 1, it does not give a unique prediction for the primor-
dial fl.uctuations for structure formation. The simplest inflation models utilizing a single
scalar field yield a flat fluctuation spectrum for the Newtonian potential with Gaussian
distributed Fourier amplitudes. Here, however, I discuss stochastic inflation calculations
(see, for example, Salopek and Bondl ) describing non-Gaussian primordial fluctuations.
The models presented may shed some light on the problems of large scale structure in the
Universe. 3 Hopefully, cosmic microwave background (CMB) anisotropy experiments3 will
provide the most definitive tests of these scenarios.
In this report, I will not use the full stochastic machinery. I will assume that the fluc-
tuations for structure formation leave the horizon when the value of the Hubble parameter
is much smaller than the Planck scale. In addition, it will be assumed that the potential
does not change rapidly at that time. I will consider only models where nonlinear effects
become important later when all scales of cosmological interest are very much larger than
the Hubble radius. Different comoving points are then no longer in causal contact, and the
analysis is tractable.
Quite remarkably, the full nonlinear dynamics oflong wavelength fields is contained in a
single equation, the separated Hamilton-Jacobi equation (SHJE) which is a radical starting
point for numerical calculations in general relativity. It is useful for several reasons. The
equation is truly covariant in that it makes no reference to either the time parameter nor
to the spatial coordinates. When one performs calculations, it is not necessary to make any
gauge choice. The Hubble parameter is calculated in field space, H == H( (h.) rather than
on a spatial lattice. Loosely speaking, the determinant of the 3-metric is the natural time
parameter because it separates from the full Hamilton-Jacobi equation. The new canonical
81
T. SluJnks et al. (etis.), Observational Tests o/Cosmological Inflation, 81-89.
© 1991 Kluwer Academic Publishers.
82

variables may actually be taken to be the spatial coordinates because they are constants
along comoving spatial points.
For multiple scalar fields, the method of calculation is based on some analytic tricks
rather than brute force numerics. IT the logarithm of the potential for multiple scalar fields
is linear, one can obtain a complete solution of the SHJE by utilizing hidden symmetries. 4
More complicated potential surfaces can be modelled by continuously matching linear In V
potentials along straight line boundaries. In fact, one can construct viable non-Gaussian
models consistent with current CMB limits if one employs a potential with three ridges. 5

2. LONG WAVELENGTH EQUATIONS


Given Bome initial conditions, it will be shown how to solve for the long wavelength evolution
of scalar fields tPi(t, z) with potential V(tPi) interacting through gravity. It will be assumed
that the metric has the form,

(2.1)
which describes an isotropic Universe with inhomogeneous scale factor ea(t,z). The lapse
function N is determined when one decides the time hypersurface, although an explicit
choice is not necessary. Gravitational radiation has been neglected which is typically an
excellent approximation, although its evolution, too, is tractable if one employs the powerful
machinery of canonical transformations} In what follows, H( t, z) == a/N is the Hubble
parameter and 1["tI>i(t, z) = e3a~i/N are the momentum densities of the scalar fields.
All second order spatial gradients in the Lagrangian of Einstein gravity with scalar
fields will be neglected, but it is necessary to retain first order spatial gradients otherwise
one returns to homogeneous minisuperspace. The energy constraint,

(2.2a)

and the evolution equations are valid at each comoving spatial point, and they are the
same as those for homogeneous flat cosmologies. The new ingredient is the momentum
constraint, which joins together the independent spatial points to make one Universe:

H. - -~ e- 3a r tl>i"J. .• (2.2b)
"- 2 "'3,'-
m."

The solution of this equation is familiar to those who study fluid mechanics. The Hubble
parameter is assumed to be a function of the scalar fields, and the momentum densities are
given by partial derivatives with respect to the scalar fields,

8H
= _~e3a_.
m2
H == H(tPi), 1["tI>i (2.3)
4r 8tPi
When these are substituted into the energy constraint, one obtains the separated Hamilton-
Jacobi equation,
(2.4)
83

This remarkable equation governs the nonlinear dynamics of the long wavelength gravita-
tional system. It is covariant in that is does not refer either to the time hypersurface nor
to the spatial coordinates.
In a significant improvement for calculations based on Hamilton-Jacobi methods4 (here-
after known as SI), I gave an complete analytic solution of the SHJE for two scalar fields
(;;., ~) interacting with an exponential potential,

(2.5)

The coupling parameter p controls the steepness of the potential. The SHJE for two scalar
fields interacting through arbitrary linear In V (th, t/1J) potential,

(2.6)

may then be solved by rotating the fields (;;., ~~) by a mixing angle (J:

(2.7)

A complete solution which depends on two arbitrary parameters, 6, mis,

l'o)1/2
= ( -81rmop (3p)(m2 + 1) [_ (41r( -
f/>tsinfJ+ t/1JcosB)]cosh(u)
H(~ht/1J;p,(J;6,m) 3'I.
m2(3p-l) + 3p exp VP mop
(2.74)
where u is a function of 6, m, ~1 and t/1J which is defined implicitly through,

v'121r [(cosB _ msinfJ)t/1J _ (mcosB + 8infJ)~1 - 6] = _3p-l


v'3P X
mop
m
[uJm2(3p - 1) + 3p + In Icosh(u) - sinh( uh/r --=-2(::-3p----:l):-+-3-pl]. (2.76)

All solutions of the SHJE with potential (2.6) may be derived from eq.(2.7). Surfaces of
constant Hubble parameter are plotted as solid curves in Fig.(l) for the case m = 1, B = O.
The family of orthogonal lines are the physical trajectories. This solution of the SHJE may
be verified by differentiation; it is actually derived by looking for symmetries in the SHJE.
The complete solution (2.7) generates a transformation to new canonical variables, 6,
m, with conjugate momenta, 1r" and 1rm, given by differentiation ofthe Hubble parameter,

8H
1rm =m~ 3a
-e - . (2.8)
41r 8m
The Hamiltonian density actually vanishes strongly at each spatial point when expressed
in terms of these new variables. Hence, they are constants in time, although they may be
spatially dependent. Moreover, the new canonical variables are constrained through the
new version of the momentum constraint,

(2.9)
84

Once may invert (2.3) and (2.8) to determine the 4 constants of integration as a function
of the original variables, a, 4>;, 1(''''1:

(2.10)

These complicated analytic expressions are explicitly given in Sl, and they will playa
crucial role in what follows. In terms of the new canonical variables,4 one may write an
explicit analytic expression for the variable ( which characterizes adiabatic fluctuations at
long wavelengths,8
(2.11)
It is the quantity of primary interest for the simplest models of structure formation. For
example, in the Cold Dark Matter Model, microwave background anisotropies at angular
scales greater than '" 3° are directly proportional to it, il.TcMBITcMB = (/15.

5
...
e
"'-..
--
"$.
0
S
I::
t\l

-5

-1 0 L....I....,1...;w:::.....JL-.oI:;...J...~l.....<L......L...~'-""'-....J,..,o<:l-JL.oQ>......L......I..-.J
-10 -5 1/20 5 10
(12rr) CPI/ mp

Fig. (1): The complete solution, H( tP;i b, miP, 9), of the separated Hamilton-Jacobi equa-
tion is shown for two scalar fields interacting through an linear In V potential, eq.(2.6).
Here, the mixing angle 9 vanishes, and the constant parameters are chosen to be b = 0
and m = 1. Surfaces of constant potential are just horizontal lines. The broken lines are
trajectories of the fields, which are orthogonal to the surfaces of constant Hubble parameter
(solid lines). One can ask what happened to space and time coordinates in the SHJE (2.4)?
Loosely speaking, a, which enters in the canonical transformation (2.8) is the most natural
time parameterization of the trajectories. The new canonical variables, b, m, 1('~, 1('m then
serve as the spatial coordinates because they have no time dependence. In this sense, the
Hamilton-Jacobi formalism yields the the choice of coordinates, both temporal and spatial,
in which the evolution appears simplest.
85

3. INITIAL CONDITIONS
In Sec. 4, models are constructed by joining several linear In V potentials together. It
will be assumed that GaU8.ian fluctuations arising from short wavelength quantum noise
are generated in region 1 where the potential parameters are Pt and 81, Non-Gaussian
fluctuations are produced when one passes over other regions of the potential.
The Hubble function, H(tP;), in region 1 is taken to be the attract or solution,

(3.1)

corresponding to b = -00 and m = 0 in (2.7) having homogeneous values. The new


momentum constraint (2.9) is then satisfied at early times, and the evolution equations
guarantee that it will be satisfied at late times. In region 1, the fields then evolve in time
a according to,

4>t(z,a) ~asin81 + tPI0(Z),


= - y411"Pt (3.2a)

11" .1( z,a ) -_ - mop


n==-e '1'; sin810
3aHaU ("") (3.2b)
y411"Pt
The initial values of the scalar fields, tPiO (z ), are classical Gaussian random fields with power
spectrum
'P . (k) == ~ltPiO(k)I' = (HO)' (_k_)-'/(1'1-1) .
•• 0 211"' 211" Hoeao
The amplitudes of the homogeneous k = 0 modes are arbitrary. IT the value of the Hubble
parameter, H o, when the longest modes in the simulation left the horizon during inflation,
is much smaller than the Planck scale then the assumption of Gaussian initial conditions is
justified. I will not give the full lattice results& here; I will show only the l:t.TCMB/TcMB
distribution at large angular scales.

4. MODEL CALCULATIONS
Two sets of calculations will be given. For illustration purposes, I will describe in de-
tail a potential with a single ridge although it generically produces Gaussian fluctuations.
When there are three ridges in the potential, then one can indeed generate non-Gaussian
fluctuations.

4.1 POTENTIAL WITH A RIDGE


Consider a potential obtained by joining two linear In V potentials continuously along the
line tP2 = tPltanXn (see Fig.(2a». Continuity of the potential at the boundary implies that
the potential parameters, 1'2 and 8, in the upper half-plane (region 2), are related to those
in the upper half (region 1) through,

!iii. = sin(X12 - 8,) . (4.1)


VP; sin(X12 - 8t}
86

For a given spatial point in the lattice, the value of the fields at the interface follow from
(3.2),
(4.2a)

cosXu .
tPll = ( (J ) ( tPlOCOS(Jl
cos Xu - 1
+ tP20Sm(Jl ),

Using these values, one can determine the constants of integration (2.10) in region 2, and
then calculate (, eq.(2.l2), which yields the CMB temperature anisotropy at large angular
scales. All the equations are algebraic.
In Fig.(2a), I show the potential as well as some typical trajectories. In Fig.(2b), I show
the distribution of ( on a 2-D slice from a 643 lattice calculation. For plotting purposes, the
initial value of the Hubble parameter was chosen to be Ho = 10-2 m 7', although microwave
background anisotropy limits would require Ho < lO- t m7'. Here, the potential parameters
in the lower half-plane are 1'1 = 20, (Jl = -60; in the upper half-plane, (J2 = -300. The
ridge is inclined at an angle of Xu = 1650 to the tP1 axis. Surprisingly, one still obtains
Gaussian statistics (Fig.(2b)) although the fields mix at the interface. For a potential with
a single ridge, this result may be proved quite generally. Apparently, each trajectory evolves
in a self-similar way.

4.2 POTENTIAL WITH THREE RIDGES


One can obtain non-Gaussian fluctuations on cosmologically observable scales from a po-
tential created by joining three linear In V regions as shown in Fig.(3). Trajectories begin
in the lower half-plane, region 1 (PI. = 20, 81 == -50°), with initial Hubble parameter,
H o = 10-8m 7', chosen to satisfy CMB limits. If the trajectories pass into the upper right
hand area, region 2 «(J2 = 0°), they receive an upward kick from the potential, which forces
them into region 3 «(J3 = -30 0 ). (If this diagram were extended, one would find that trajec-
tories actually cross in region 3.) The angles of the ridges starting with the lower right and
proceeding counterclockwise are, Xu = 100 , X23 = 39 0 , and X3l = 1580 • The calculations
here are more complicated than in Sec. 4.1 because one must determine where a trajectory
strikes region 3 after it has passed through region 2. ( is not constant in passing through
region 2. The distribution of , in region 3 is plotted in Fig.(3b). For the parameters shown,
it was found that non-Gaussian fluctuations can occur if the fields passed sufficiently near
the origin, which can be arranged through the arbitrary choice of the homogeneous mode
amplitudes in eq.(3.2). A thorough search of parameter space has yet to be made.

5. CONCLUSIONS
Because inflation probes arbitrarily small distance scales, it is quite likely that non-Gaussian
fluctuations may arise from the interactions of various fields. However, the greatest obstacle
is actually calculating these effects. As the first non-trivial improvement over homogeneous
minisuperspace models, one should calculate nonlinear effects at long wavelengths. How-
ever, there is currently nO strong evidence for non-Gaussian fluctuations. 7 The goal here
then is to generate some simple models that can be used to test the non-Gaussian hypothesis
87

.04 Gaussian Fluctuations


From Inflation
.02 400
B
......... 0
-l. 200
-.02

-.04
OWW~~~~WW~~LU

-.1 -.08 -.06 -.04 -.02 0 -4 -2 0 2 4


"'1 /mp
«(" _<("»/ ( «(" _<("»2> )1/2
Fig. (2): (a) Some trajectories in field space are shown for a potential with a single ridge.
The solid curves are lines of uniform potential, whereas the very heavy line is the ridge.
The trajectories (broken lines) begin in the lower half plane with Gaussian initial conditions
generated from short wavelength quantum noise. Even when they pass over the ridge, the
nonlinear metric fluctuation ( still remains Gaussian distributed, as shown in the histogram
of Fig.(b) which displays the results of a 643 lattice simulation.
Fig. (3): (a) Non-Gaussian fluctuations consistent with CMB limits may be generated
when one passes over several ridges in the potential. The light solid curves are lines of
uniform potential, whereas the heavy lines are the ridges. H the scalar field trajectories
(broken lines) pass sufficiently near the origin, nonlinear effects at long wavelengths are
important. For a 643 lattice simulation, the histogram of resulting microwave background
fluctuations at large angular scales (IlTCMB/TcMB = (/15) is shown in Fig.(b). For
comparison, a Gaussian distribution (smooth curve) with the same mean and dispersion as
the histogram is also shown.

10'" POTENTW. WITI:I 3 RIDGES 600 r-T"'"I"'"T"T-r-r..................-r-r-r-r-...........,..,....,

400

200

OL.L..&.....L...IotL..&.....L..J...L..LL..L.L1I...l:=Io..L.1...L...l
-4 -2 0 2 4
«(" _<("»/ ( «(" _<("»2> )1/2
88

with observational data. The major achievement of this report over earlier workS - lO ,I,11 is
an efficient way of calculating ob.eMlable non-Gaussian models using Hamilton-Jacobi the-
ory. (Previously, the only fully calculable non-Gaussian model consistent with microwave
background limits was proposed by Bardeenu who considered the square of a Gaussian
random field whose power spectrum peaked at cluster scales.)
In the future, the models presented here will be compared with large scale structure
observations. The cleanest test of models of non-Gaussian fluctuations will hopefully come
in the near future from the Cosmic Background Explorer satel1ite3 which should be in a
position to test the Cold Dark Matter Model prediction. 13
I would like to thank J .R. Bond for several stimulating discussions. This work was
supported by the U.S. Department of Energy and NASA at Fermilab (Grant No. NAGW-
1340).

6. References

1 Salopek, D.S. and Bond, J.R., Phys. Rev. D42, 3936 (1990), D43, 1005 (1991).
2 Maddox, S.J., Efstathiou, G., Sutherland, W.J. and Loveday, J., Mon. Not. R. Astr.
Soc., 242, 43P (1990).
3 Smoot, G., in these proceedings (1990).

• Salopek, D.S., submitted to Phys. Rev. D Fermilab-Pub-90/213-A (1990). Referred to


as S1.
5 Salopek, D.S., to be submitted to the Ap.J. (1991). Referred to as S2.

II Bardeen, J.M., Steinhardt, P.J. and Turner, M.S., Phys. Rev. D28, 610 (1983).

T Melott, A.L., in these proceedings (1990).

S Allen, T.J., Grinstein, B. and Wise, M., Phys. Lett. B191, 66 (1981).

II Matarrese, S., Ortolan, A. and Lucchin, F. Phys. Rev. D40, 290 (1989).

10 Vi, I., Vishniac, E. T. and Mineshige, S., University of Texas Preprint (1990).

11 Mollerach, S., Matarrese, S., Ortolan, A. and Lucchin, F., SIS SA 143A preprint (1990).

12 Bardeen, J .M., talk presented at W ork.hop on Inflation and Ezotic Co.mic Structure
Formation, Canadian Institute for Theoretical Astrophysics, Vancouver, Canada,
(1990).
13 Bond, J.R., in Proc. of The Early Universe, ed. W. Unruh and G. Seminoff, Dordrecht:
Kluwer (1988).
89

DISCUSSION:

Guth: My understanding is that the linearized calculation always gives a Gaussian distri-
bution, so non-Gaussian effects arise only in higher order. Since the fluctuations are small,
one would expect the non-Gaussian contributions to generally be smaller yet, unless some
coefficient blows up to make them large. Is this picture correct? In the calculation that you
did which gave non-Gaussian behaviour, do you know if it was crucial that the potential
function was not Taylor expandable?

Salopek: For the parameters chosen, non-Gaussian fluctuations arise if the scalar field
passes near the origin where non-linearities are certainly important. In the future, I will
investigate whether this is essential.
White: Is the non-Gaussian nature of the fluctuations you predict associated with any
particular spatial scale or are the statistics scale-free?
Salopek: The statistics are scale free in these models which are the simplest scenarios for
non-Gaussian statistics. One can construct more complicated models as well.
Sahni: Do you have to fine tune the initial conditions to get a non-Gaussian distribution
on a specific length scale?
Salopek: I do not tune any parameters to accentuate any specific length scale. But as A.
Guth has mentioned, the results may be sensitive to the form of the potential.
NON-BARYONIC DARK MATTER

S. SARKAR
Department of Theoretical Physics,
University of Oxford,
1 Keble Road, Oxford OXl 9NP,
England.

ABSTRACT. We discuss the theoretical motivation for new particles which may constitute the
cosmological dark matter and review the progress of experimental efforts at their detection.

1 Introduction

My brief is to discuss particles whose very discovery would provide strong support for
inflation, in that their relic cosmological abundance would be naturally expected to provide
the critical density (in the form of dark matter). This, of course, is in no way a test of
inflation since failure to find such particles would not be particularly exciting. Furthermore,
one may question the effectiveness of this enterprise given that new particle candidates for
dark matter are announced with monotonous regularity. Surely we cannot expect difficult
experimental searches to be undertaken at the whim of theoretical fashion. Rather than
delight in a 'zoo' of weird particles, I will argue instead that there are o~y a few well
motivated candidate particles, and that nearly all of these are capable of being found, or
excluded, by ongoing or forthcoming laboratory experiments. We may well resemble the
proverbial drunk looking for the lost keys under the lamp post, but the lamps shine with
theoretical insight! To this end, I will briefly discuss why particle physicists believe in the
existence of new particles which are likely candidates for dark matter.
Particle dark matter involves physics beyond the 'standard model'. We are concerned
here with two such models - the Standard SU(3) i8l SU(2)i8l U(l) Model of particle physics
(which should really be called the Standard Theory) and the standard Big Bang model of
cosmology (which is still only a model). 1 The two models are satisfactorily consistent upto
the begining of the nucleosynthesis era, at t '" 10- 2 sec after the Big Bang, beyond which
there is no direct observational basis for cosmology. However our faith in the Standard
Model of particle physics, brilliantly confirmed by all experiments to date, allows us to
extrapolate further back to t '" 10- 12 sec. At least two phase transitions are believed
to have occurred in this interval - one associated with the confinement of quarks into
hadrons and the dynamical breaking of chiral symmetry at AQCD '" 200 MeV, and another
at TEW '" 300 GeV associated with the spontaneous breaking of the electroweak symmetry
where all known particles are supposed to have received their masses through the Higgs
mechanism. To go beyond this point requires an extension of the Standard Model. Indeed,
the very success of the model demands such new physics.
IThe distinction I am making, following Manfred Eigen, is that a theory must be either right or wrong
whereas a model has a third alternative ... it may be irrelevant!

91
T. Shanks et al. (eds.). Observational Tests o/Cosmological Inflation. 91-102.
© 1991 Kluwer Academic Publishers.
92

Just as there is an apparent 'fine tuning' of the initial conditions of the Big Bang model
which motivates the idea of an early era of inflation (Guth 1991), there are similar prob-
lems with the Standard Model whose solution must be sought in new physics at higher
energies. For example, the Higgs boson mass cannot be much larger than its vacuum ex-
pectation value which sets the Fermi scale, v =(..j2GF)-1/2 ~ 247 GeV. This creates
a 'naturalness' problem, viz. why is its mass so small relative to say, the Planck scale
Mp (= G"N1 / 2 ~ 1.22 X 1019 GeV), given that it receives quadratically divergent radiative
corrections due to its couplings to all massive particles? 2 Supersymmetry addresses this
problem by imposing a symmetry between bosons and fermions which makes such radiative
corrections cancel to zero. This then requires all known particles (boson/fermion) to have
supersymmetric (fermion/boson) partners distinguished by a new quantum number called
R-parity; the lightest supersymmetric particle (LSP) would be stable given R conservation.
Supersymmetry must be broken in nature since known particles do not have supersymmet-
ric partners of the same mass. However the Higgs mass would still be acceptable if the scale
of supersymmetry breaking (and the mass of the LSP) is not much beyond the Fermi scale.
When such breaking is realised locally, in analogy to gauge theories, a link with general
coordinate transformations, i.e. gravity, emerges; this is supergravity. Technicolour is an
alternative approach in which the offending elementary Higgs particle is made composite;
electroweak symmetry breaking is now seen as a dynamic phenomenon, akin to the break-
ing of chiral symmetry by the strong interaction. This too predicts new particles called
technibaryons, the lightest of which may be stable on cosmological time scales.
A different problem is that charge-parity (CP) is known to be well conserved by the
strong interaction given the stringent experimental upper limit on the neutron electric
dipole moment, whereas QCD, the successful theory of the strong interaction, contains an
arbitrary CP violating parameter. An attractive solution is to replace this parameter by
a field which dynamically relaxes to zero - the axion. This is a pseudo-Goldstone boson
generated by the breaking of a new chiral global U(l) 'Peccei-Quinn' symmetry; it acquires
a small mass and may be cosmologically stable if it is sufficiently light.
Another motivation for going beyond the Standard Model is the unification of forces.
Grand unification of the strong and electroweak interactions at high energies provided the
original context for discussion of an inflationary phase transition, motivated partially by the
need to dilute the embarrassingly large relic abundance of magnetic monopoles expected in
these theories. Unification naturally provides for baryon number (B) and lepton number (L)
violation, which allows generation of the cosmological baryon asymmetry as well as masses
for neutrinos. Going further, superstrings are the most ambitious attempt yet towards a
finite quantum theory of gravity and its unification with all other forces. 'String-inspired'
models incorporate grand unification and supersymmetry and relate 'low energy' physics
in four dimensions to the very different mathematical world in which superstrings exist.
These models contain gauge singlet particles which couple very weakly to matter.
The reason behind the proliferation of particle candidates for dark matter is now clear.
It is a common feature of new physics beyond the Fermi scale to predict the existence of new
particles which may be stable by virtue of new conserved quantum numbers and whose weak
2By contrast, it is 'natural' for fermion and gauge boson masses to be small relative to the Planck
scale, since, letting such masses go to zero reveals new symmetries (chiral symmetry and gauge symmetry,
respectively); there is no such symmetry which 'protects' the mass of the Standard Model Higgs.
93

interactions ensure a cosmologically significant relic density. In addition, known particles


such as neutrinos, though strictly massless in the Standard Model, may acquire masses
from such new physics, enabling them also to be candidates for dark matter.
What about baryons, which after all are the only form of 'matter' we know about? It has
become customary to argue that baryons cannot be the cosmological dark matter based on
considerations of primordial nucleosynthesis, galaxy formation, et cetera. This will shortly
be discussed by Carr (1991), so I would like to suggest a rather different reason, viz. that
it is more difficult to create baryons from a theoretical point of view! If we calculate the
expected relic abundance of stable baryons, i.e. nucleons, in the same manner as for any
other massive particle, assuming chemical equilibrium at sufficiently high temperatures, we
find, because of their strong self annihilation, a very small value fiN ,... 10-11 • By contrast,
just the luminous nucleons in galaxies contribute at least fiN ,... 10- 2 • To explain this
huge discrepancy, we have to postulate a chemical potential for baryons - a cosmological
asymmetry between matter and antimatter. This requires new physics to violate Band
CP at high temperatures and, additionally, the relevant processes are required to be out
of equilibrium, in order to ensure time asymmetry. While there are indeed plausible ways
to achieve all the above objectives (see Dolgov 1991), it is very difficult to obtain the
required asymmetry and, furthermore, to protect it from being washed out due to (recently
identified) B-violating effects in the Standard Model itself at high temperatures. As we will
see below, it is possibly easier to account for the unknown 99% of the universe than the
known 1%! Both explanations need physics beyond the Standard Model. 3

2 Candidate Particles

Let us consider the various types of cosmological relics adopting the classification suggested
by Hall (1988). For further details see Kolb & Turner (1990) and Sarkar (1991).

2.1 'PLASMA' RELICS

These are particles which became collisionless while still relativistic, the canonical example
being light neutrinos, the first suggested candidate for dark matter (Cowsik & McClelland
1973). The weak interaction rate rw( ~ G}T5) which keeps neutrinos in kinetic equilibrium
(i.e. T" = T) through reactions such as ve- -+ ve- falls behind the Hubble expansion rate
H(~ T2/Mp) at the 'decoupling' temperature Td ~ (G}Mp)-1/2 ,... O(MeV). Neutrinos
which are relativistic at this time (i.e. which have m" ~ Td) are nearly as abundant as
photons; taking into account the subsequent increase in the comoving photon number due to
the annihilation of e+e- pairs at T;:, me, the present number density of such neutrinos (with
= =
g" 2 helicity states) is n" (3/11)n'"( ~ 114 cm- 3 for a present blackbody temperature
To ~ 2.74 OK. Hence their contribution to the present density parameter is

fi h2~~(g,,) (1)
" 93 eV 2 .
3While non-perturbative effects in the Standard Model at T ~mw can erase any pre-existing (B-L
preserving) baryon asymmetry, the observed asymmetry cannot be created by such effects without further
extending the Higgs sector (e.g. McJ,erran et al. 1990), given the bounds on the Higgs mass from LEP.
94

Comparison with the present experimental bounds on neutrino masses (Particle Data Group
1990), viz. mv. < 9.6 eV, m v,. < 270 keY, mVr < 35 MeV, shows that whereas a massive
Ve cannot provide the critical density, the v,. and V T are still plausible candidates.
Similar considerations apply to light particles which are more weakly interacting; they
may have a higher decoupling temperature Td and consequently undergo more dilution
relative to photons so that critical density would correspond to a higher mass, upto O(keV).
One candidate is a light photinofhiggsino (1/ if), expected in supersymmetric models
with global supersymmetry breaking. Though experiments provide no useful constraints,
it is possible to exclude such a particle from considerations of the energetics of SN 1987a.
These particles should have been profusely created by nucleon bremsstrahlung in the dense
core of the collapsing star and the resultant energy loss would have been too large to be
compatible with the observed energetics of neutrino emission (Ellis et al. 1988a). It is
more difficult to rule out a light gravitino (G) which may occur in 'no-scale' supergravity
models. The sole experimental constraint comes from an analysis of e+ e- annihilation and
requires only that its mass exceed 0(10- 5 ) eV (Dicus & Roy 1989).
An exciting recent indication that neutrinos may indeed be massive comes from analysis
of solar neutrino observations. The combined data from the Homestake and Kamiokande
experiments (and preliminary indications from the SAGE experiment) suggest that the
Mikheyev-Smirnov-Wolfenstein mechanism of enhanced oscillations in matter between Ve
and some other neutrino can account for the observed energy dependent depletion of solar
neutrinos below that predicted by the standard solar model if ilm 2 sin 2 8 '" 10-8 eV2,
where sin 2 8 is the 'mixing' between the two neutrinos and ilm 2 the mass difference squared
(Bahcall & Bethe 1990). In conjunction with the 'see-saw' model in which neutrinos masses
are expected to scale as the square of the associated charged fermion (say, quark) mass,
this suggests the following hierarchy (taking 8 = 8Cabbibo '" 0.2):

(2)

These masses are far below the sensitivities of present or future laboratory experiments
but there may be an astrophysical signature of such a massive V T • It has been conjectured
(Sciama 1990, see also Mellott et al. 1988) that the ionisation of the intergalactic and
interstellar medium is due to photons from decaying relic ",30 eV neutrinos (e.g. VT -+ Vel),
given that conventional astrophysical sources have difficulty in explaining the observations.
The required radiative lifetime of '" 2 x 10 23 sec is 106 times longer than the age of the
universe but is still too fast (!) from a theoretical point of view and is only attained in
rather non-standard extensions of the Standard Model, e.g. supersymmetry with R-parity
breaking (Gabbiani et al. 1990). A recent search for the decay line at ",15 eV from the
Abell cluster A665 by the ASTRO-l satellite has been unsuccessful, but this may be due
to absorption by cold neutral matter (Fabian et al. 1991).
An equally exciting recent development 4 is the experimental detection of a 17-keV
neutrino mixed at the'" 1% level with the electron neutrino in studies of the f3 decay
of 35S (Hime & Jelley 1990) and 14C (Sur et al. 1990), in support of an earlier study of
3H (Simpson 1985). Experimental bounds on neutrinoless f3f3 decay suggest a Dirac rather
than Majorana mass for this state (i.e. it cannot be its own antiparticle). Laboratory limits
{This was actually announced just after this meeting but has been included in view of its topicality.
95

on neutrino oscillations imply that this cannot be the v,. but it could be the v", or perhaps,
a right-handed 'singlet' which does not couple to the ZO boson. This neutrino must be
unstable for otherwise its present energy density would be greatly in excess (see eq. 1) of
the observational limit flh2 < 1 which follows (for a matter-dominated FRW cosmology)
from bounds on the present age, to > 1010 yr, and Hubble parameter, h == HollOO Km
seC 1 Mpc 1 > 0.4. By decaying into relativistic particles, the excessive energy density
can be red-shifted away; the decay products have then subsequently radiation-dominated
the universe - they constitute the dark matter! In this case, the energy density limit is
tightened to flh2 < 1/3 and this imposes an upper limit on the lifetime (Dicus et al. 1978)

T < 1.5 X 1012 (9;) -2 sec. (3)

Because of the small observed mixing (sin2 8 '" 0.01) with the electron neutrino, such a
short lifetime is very difficult to obtain in any particle physics model where the decay
occurs into lighter neutrinos; it is also difficult to adequately inhibit concommitant decays
into photons which would violate observational bounds from the ultraviolet and 2.7 OK
radiation backgrounds and from SN 1987a (Hime et al. 1991). In order to allow fast decays
into 'invisible' particles it is necessary to invoke new particles called majorons - Goldstone
bosons created by the global violation of lepton number (see Glashow 1991). There are
important implications for studies of the growth of structure in the universe. In the past
such studies have concluded that the above 'scenario' of a decaying particle cosmology is
ruled out; however the (sub-dominant) radiative decays may have reionised the universe
thus allowing these constraints to be evaded (Fukugita 1988). It is also claimed that a Dirac
neutrino of mass 17 keY conflicts with constraints from primordial nucleosynthesis and the
energetics of SN 1987a (Kolb & Turner 1991), although the seriousness of the conflict is
debatable (Hime et al. 1991). Clearly an exciting time lies ahead should the experimental
results receive further confirmation!

2.2 'FREEZE-OUT' RELICS

These are massive particles which are non-relativistic when they become collisionless and,
consequently, whose present density is much smaller than that of the blackbody photons.
The equilibrium number density of a species i becomes sharply depleted due to a Boltzmann
factor cm;/T when the temperature drops below its mass mi. Annihilation reactions,
e.g. i",
+-+ e+e-, which maintain chemical equilibrium, proceed at a rate rA = ni(uAv),
where the cross-section has been averaged over the thermal velocity. This rate decreases
exponentially when the particles turn non-relativistic, hence the species 'freezes-out' (i.e.
its number density stops tracking the equilibrium value) when rA drops below H, at a
temperature well approximated by TJ ~ m;j[17 + In{(m;jGeY)«(uAV)/lO-lO Gey-2)}].
The more strongly the particles interact, the longer they stay in equilibrium and deplete
themselves, hence the surviving abundance varies inversely as the annihilation cross-section,
3 x 10- 10 Gey2
flvh2~ ( ) (4)
UAV
The relic density is therefore higher the more weakly interacting the particle, as had been
anticipated ( ... the meek shall inherit the Earth ... ).
96

There are several features of interest about eq. 4. First of all critical density is naturally
attained for a massive weakly interacting stable particle, for which there is a plethora of
candidates in theories going beyond the Standard Model. There is a certain irony here
in that the more difficult a particle is to detect in the laboratory, the more likely it is to
dominate the universe! For example, supersymmetric particles have so far remained unob-
served in accelerator experiments. This implies that the LSP, if it exists at all, is sufficiently
weakly interacting that nLSP ~ 0.01, i.e. it contributes as much as all the visible matter.
Baryons, as we have discussed earlier, have a present abundance greatly in excess of that
given by the above analysis, hence they cannot be freeze-out relics but must have resulted
from non-equilibrium processes. What about particles even more strongly interacting than
baryons? The unitarity constraint, {UAV} :5 81r/m?, implies an upper bound of 340 TeV
on the mass of any elementary stable relic (Hall 1988, Griest & Kamionkowski 1990). This
bound can only be violated if freeze-out occurs during the phase transition which generates
a mass for the particle 5 or if there is significant generation of entropy after freeze-out, for
example by a first-order phase transition or a decaying massive particle.
The originally proposed candidate for a 'freeze-out' relic was a (fourth generation)
heavy neutrino of mass O(GeV) (Lee & Weinberg 1977); however this was hardly well
motivated since there is no reason why such a particle should be stable. It would be un-
natural if there were no mixing at all with lighter neutrinos, while a mixing as small as
10- 14 would make it cosmologically unstable. In any case this possibility is now ruled out
by measurements of the ZO ~ vii decay width at LEP which finds the number of (weak
isodoublet) neutrino generations lighter than '" mz/2 to be (see Dydak 1990)

Nil = 2.89 ± 0.10. (5)


This requires a fourth generation neutrino to have mil ~ mz/2 which would make it cos-
mologically unimportant since for non-relativistic neutrinos annihilating through the ZO,
{UAV} '" G'j..m~ implying nil '" (m ll /GeV)-2. However, above the ZO pole UA begins de-
creasing as '" m;2 which would appear to make the relic abundance significant again at
mil '" O(TeV) (Dolgov & Zeldovich 1982). This is averted by additional annihilation chan-
nels (e.g. vii ~ W+W-) which open up for such massive neutrinos, greatly reducing their
relic abundance (Enqvist et al. 1989). In any case, heavy Dirac neutrinos (which have vec-
tor couplings, hence coherent weak interactions) have been directly excluded as halo dark
matter by low background semiconductor detectors originally built to study the (3(3 decay
of germanium. These are sensitive to the energy deposited by elastic scattering against nu-
clei of throughgoing dark matter particles and the negative results exclude the mass range
milD'" 10 GeV - 4.5 TeV (Ahlen et al. 1987, Caldwell et al. 1988).
Though a tau neutrino of mass O(MeV) is experimentally allowed, it would nominally
have a relic density well in excess of observational limits if it were stable. Decays cannot
reduce the abundance for they would occur into 'visible' channels (e.g. v., ~ e-e+ve ) and
are excluded by a combination of cosmological and laboratory constraints (Sarkar & Cooper
1984). However if the v., possesses a large magnetic moment of'" 10-6 J-lB, the resulting
enhanced annihilation cross-section can reduce the relic density to nil.. '" 1 (Giudice 1990),
if it is somehow kept stable. This interesting possibility is ruled out by a search for the
5The recent LEP bound on mHO implies that T, < TEW for all particles which obtain mass from the
Standard Model Higgs, hence their abundance cannot have been thus affected (Dimopoulos et al. 1990a).
97

implied 1/".e- elastic scattering events in a 'beam dump' experiment at CERNj this sets a
preliminary limit I'll.. < 3 X 1O-7 1'B (Cooper-Sarkar et al. 1991). Studies of rare meson
decays may improve this bound further (Bergstrom & Rubinstein 1991).
A more interesting candidate is the LSP, which, as we have seen, is well-motivated
theoreticallr. It is most likely to be the neutralino, a linear combination of the photino
en, zino (ZO) and higgsino (if) or a sneutrino (Ii). The LEP bound (eq. 5) constrains
sneutrinos in a manner similar to heavy neutrinos since each Ii would contribute ",0.5 to the
Nil count. The neutrallno would also contribute but by a small amount dependent on the
actual admixture of its components since they couple differently to the Zo. The parameter
space for such mixing is constrained by the failure to find charged or strongly interacting
supersymmetric particles (charginos and gluinos) at LEP and at the FERMI LAB collider.
Detailed analyses taking all such constraints into account find that in order for it to be a
candidate for dark matter, the neutralino mass must exceed at least 20 GeV and can extend
to O(TeV) (Krauss 1990, Ellis et al. 1990a). There are also bounds coming from searches in
underground nucleon decay detectors for high energy neutrinos produced by annihilations
of heavy dark matter particles which have accreted in the Sun. A recent analysis excludes
sneutrinos in the range mj; '" 3 - 90 GeV (Sato et al. 1990), and the excluded mass range
is extended to several TeV by the aforementioned f3f3 decay experiments. However the
neutralino is not particularly constrained by these bounds (Gelmini et al. 1991).
Another interesting candidate is an electroweak gauge triplet neutrino with zero
hypercharge, stabilised by some new conserved fermion number. Depending on the model;
such a particle would be cosmologically relevant for a mass'" 30 - 80 GeV (Salati 1991)
or several TeV (Dimopoulos et al. 1991). Alhough this neutrino does not contribute to the
ZO decay width, it is expected to have charged partners which should be detectable at the
proposed LHC or sse colliders, or possibly even at LEP II.
So far we have restricted our attention to neutral weakly interacting particles. It is,
in fact, unlikely that dark matter particles are charged or strongly interacting since such
particles would have bound to ordinary matter and thus shown up in the sensitive searches
which have been conducted for anomalously heavy isotopes of ordinary nuclei (see Rich
et al. 1987). These searches have set upper limits several orders of magnitude below the
expected abundances for such particles with masses up to ",10 TeVj this rules out, in
particular charged technibaryons (Chivukula & Walker 1990). Nevertheless, there has
been recent interest in even more massive charged (De Rujula et al. 1990, Dimopoulos
et al. 1990b) or strongly interacting (Starkman et al. 1990) particles. There are many
astrophysical arguments against such particles, the most severe of which is imposed by
the existence of neutron stars which indicates that such particles do not congregate inside
them sufficiently to form black holes which would swallow up their hosts (Gould et al. 1990).
Direct searches in cosmic rays for charged dark matter have also been negative (e.g. Barwick
et al. 1990), making this possibility quite unlikely.
It had been suggested that the annihilation of heavy freeze-out relics in the Galactic halo
into antiprotons and positrons may provide a distinctive signature in cosmic rays. However
the observed flux of p and e+ is found to be consistent with spallation secondaries. Unfor-
tunately, this does not yield any useful constraints on dark matter, given the uncertainties
in cosmic ray propagation (Ellis et al. 1988b). The prospects for detection of annihilation
'Y rays are also bleak (e.g. Giudice & Griest 1989).
98

2.3 'ASYMMETRIC' RELICS

These are relics which have a cosmological asymmetry between particles and antiparticles,
the best example being, of course, nucleons which have (nN - nji;)/(nN + nji;) ~ 10-8 •
Can other particles also develop such an asymmetry so as to be plausible dark matter
candidates even when their freeze-out abundance is insignificant? Recently, a physical
mechanism which can do this has emerged with the discovery that non-perturbative effects
in the Standard Model cause unsuppressed fermion number violation at T ~ mw, which
would distribute any pre-existing fermion number among all electroweak doublets (Kuzmin
et al. 1985, Arnold & McLerran 1987). Given that nN ~ 10-2, this would allow a particle
of mass 0(100 GeV) with a similar asymmetry to provide the critical density. This would
also explain the 'cosmic' ratio of ",100 between the amounts of dark and visible matter as
being just the ratio of the particle mass to the nucleon mass.
The suggestion that a neutral technibaryon of mass O(TeV) may acquire such an
asymmetry (Nussinov 1985) has been examined recently in this context (Barr et al. 1990).
However the (J(J decay experiments rule out any such particle with coherent weak inter-
actions as halo dark matter. Another problem is that technibaryons may, in fact, be
metastable, albeit with a lifetime longer than the age of the universe. High energy neutri-
nos from such decays are detectable in underground nucleon decay detectors through the
upward-going muons they would generate in the surrounding rock. The present limits on
this signal require the lifetime of aTeV (or higher mass) relic to be ~ 1016 yr if it is to have
n '" 1 (Ellis et al. 1990b). Another possibility is a heavy Dirac neutrino6 , which was
suggested (Gelmini et al. 1987) as a candidate for the 'cosmion' - dark matter particles
which, when trapped in orbits inside the Sun, lower the core temperature by transporting
heat outward, thus solving the solar neutrino problem. (The required large elastic scatter-
ing cross-section nominally implies a large annihilation cross-section which would result in
the relic abundance being negligible, hence the necessity of an initial asymmetry.) However
this possibility is ruled out by the underground (J(J decay detector results referred to earlier.
In fact all cosmion candidates which have coherent weak interactions are essentially ruled
out by a similar dedicated experiment using a silicon detector (Caldwell et al. 1990).

2.4 'SOLITON' RELICS

The term soliton is used here informally to imply energy that is spatially localized. The
canonical example is a topological defect in a Higgs field - domain walls, strings,
monopoles and textures - which may result from symmetry breaking in the early uni-
verse. These are regions of the symmetric vacuum which are stabilized against the (energet-
ically favoured) asymmetric vacuum due to some non-trivial topological property. Among
these the only remotely plausible candidate for dark matter is the magnetic monopole, as-
suming that its initial abundance (which is either enormous, or zero due to inflation!) is
somehow adjusted to the appropriate level. The motivation for this was originally provided
by the claim of the experimental detection of a magnetic monopole but subsequent experi-
ments have set flux limits ",4000 times smaller and ruled out monopoles with masses upto
Mp as halo dark matter (e.g. Huber et al. 1990, Bermon et al. 1990).
6A Majorana fermion cannot have an asymmetry, being its own antiparticle.
99

There may also be non-topological soliton relics which are regions of false vacuum sta-
bilized by some conserved charge. The best discussed example are quark nuggets -
bubbles of quark matter stabilized by the presence of strange quarks - which may have
been created at a first-order QCD phase transition (Witten 1984). However these are likely
to have evaporated during the subsequent evolution and are otherwise constrained by direct
searches in cosmic rays and various astrophysical phenomena (see Alcock & Olinto 1988).

2.5 'OSCILLATION' RELICS

The Higgs scalar field in the Standard Model is strongly coupled to matter, so that when
electroweak symmetry breaks and it rolls to the minimum of its potential (the asymmetric
vacuum), it can radiate the energy density of the symmetric vacuum very easily. However
if a scalar field ¢> has only very weak couplings, it will oscillate about the minimum of its
potential (at (01¢>10} = (J') without significant damping, according to ¢>( x, t) = (J' + ¢>oeimt ,
where m is the mass of the quanta which the field creates. The energy density in these
oscillations corresponds to a uniform distribution of ¢> quanta at rest, i.e. it is diluted by
the Hubble expansion in the same manner as non-relativistic particles. Hence it grows with
time relative to the energy density in radiation and may well dominate by the present epoch
if the relevant phase transition occurred in the radiation-dominated era.
The best example of such a relic is the axion, which arises, as discussed earlier, from
the spontaneous breaking of a U(l) symmetry at a scale fa. Because this symmetry is also
explicitly broken by QCD instanton effects, the axion acquires a small mass ma '" Nri fa
when the temperature drops to T '" AQCD. This mixing with the pion allows for the axion to
decay into two photons with the lifetime Ta-+2-y '" (m'l,:/m a )5T"o-+2-y '" 4x 1024 sec( m a /eV)-5.
Accelerator beam dump searches for such decaying axions have set the limit fa > O(TeV)
i.e. ma < 0(10) keY. This implies that axions are light enough to be produced in stellar
interiors such as those of red giants and supernovae. Considerations of stellar cooling
(particularly SN 1987a) through axion emission then require fa > 0(10 10 ) GeV, the precise
bound depending on the assumed couplings of the axion to fermions (see Raffelt 1990). In
that case the energy density in the coherent oscillations of relic axions is cosmologically
significant; in fact the present energy density would be excessive unless fa < 0(10 12 ) GeV
(see Turner 1990). Hence only a small 'window' is left for the axion:

ma ~ 10-5 - 10- 3 eV. (6)

It has been argued that axions are radiated in significant numbers by the global strings
which are also created by the same U(1) symmetry breaking, and that their relic energy
density, in fact, dominates over that in coherent oscillations (Davis & Shellard 1989). In
that case the bound from the cosmological energy density would be fa < 0(101°) GeV,
leaving very little, if any, allowed region for the axion. However this is disputed by Harari
& Sikivie (1987) who argue for a different energy spectrum of the radiated axions which
would imply that their present energy density is ",100 times smaller. In either case it is
clear that ifaxions do exist then they definitely do contribute to the dark matter. Direct
searches for halo axions have looked for their conversions to monoenergetic (microwave)
photons in a magnetic field; however the presently attained experimental sensitivity is still
a factor of ",100 too small (DePanfilis et al. 1989, Hagmann et al. 1990).
100

3 Experimental Searches

As we have seen above, results from accelerator as well as non-accelerator experiments


(which were designed for other goals) have already excluded many particle candidates for
dark matter. However the indirect detection techniques have nearly reached their limiting
sensitivity and we cannot expect further results in this direction. New particles too appear
unlikely to be discovered before the next generation of hadron colliders turn on.
The search for particle dark matter must now be pursued through direct detection exper-
iments. Shielded nuclear recoil detectors sensitive to heavy particles and microwave cavity
detectors for axions have been mentioned earlier. These and other detection ideas have
been considered in some detail (Primack et al. 1988, Smith & Lewin 1990). The present
limits suggest that a freeze-out relic would be very heavy, with only axial couplings hence
incoherent weak interactions. This implies a low event rate in recoil detectors, which may
necessitate cryogenic operation to reduce the background. Several such experiments are
presently being constructed. Although axion detection experiments are already underway,
they require significant improvement in sensitivity. As for light neutrinos, the relic back-
ground appears almost impossible to detect, but a time-of-flight mass measurement may
be possible with operational detectors if a supernova occurred nearby.

4 Summary

Inflation and dark matter have been the two dominant themes in the dialogue between
cosmologists and particle physicists in recent years. In both cases cosmology provides the
'problem' while particle physics attempts to provide the 'solution'. It might be argued that
stimulating though such a dialogue is, it has not actually led to any actual progress; we
still talk of hypothetical particles, cosmological scenarios ... It is certainly true that many
possibilities which at first seemed promising have been found wanting when subjected to
closer scrutiny. This, I believe, is a sign of the growing maturity of the field - an indication
that these possibilities are being taken seriously by cosmologists and particle physicists alike.
After all the usual tribute we pay to a good idea is to try to show that it is wrong! One
need not be overly concerned, however, about recent reports that all forms of 'cold dark
matter' are ruled out (as was 'hot dark matter' earlier!) on the basis oflarge-scale structure
(Efstathiou 1990). It must be emphasized that these arguments refer to specific models. 1
There has certainly been progress in the field. Though the precise amount of dark
matter and its location is still a matter for debate, even distinguished sceptics have come
to accept that there is dark matter in the universe. Similarly, no one seriously doubts that
there must be physics beyond the Standard Model, though the nature of such physics is
not clear. The link between the two is certainly a very plausible working hypothesis and
one that has motivated many experimentalists to embark on ingenious experiments to test
it. The risks are indeed high but so are the possible rewards. We wish the experimentalists
good luck and look forward to learning what our world is made of.

Acknowledgements. I would like to thank John Ellis, Dennis Sciama and Peter Smith for many
discussions and the organisers for the opportunity to participate in this stimulating meeting.
101

References

Ahlen, S.P., Avignone, F.T., Brodzinski, R.L. et al. , 1987. Phys. Lett., B195, 603.
Alcock, C. & Olinto, A., 1988. Ann. Rev. Nucl. Part. Sci., 38, 161.
Arnold, P. & McLerran, L., 1987. Phys. Rev., D36, 581.
Bahcall, J.N. & Bethe, H.A., 1990. Phys. Rev. Lett., 65, 2233.
Barr, S.M., Chivukula, R.S. & Farhi, E., 1990. Phys. Lett., B241, 387.
Barwick, S.W., Price, P.B. & Snowden-Ifft, D.P., 1990. Phys. Rev. Lett., 64, 2859.
Bergstrom, L. & Rubinstein, H.R., 1991. Phys. Lett., B253, 168.
Bermon, S., Chi, C.C., Tsuei, C.C. et al. , 1990. Phys. Rev. Lett., 64, 839.
Caldwell, D.O., Eisberg, R.M., Grumm, D.M. et al. , 1988. Phys. Rev. Lett., 61, 510.
Caldwell, D.O., Magnusson, B., Witherell, M.S. et al. , 1990. Phys. Rev. Lett., 65, 1305.
Carr, B., 1991. these proceedings.
Chivukula, R.S. & Walker, T.P., 1990. Nucl. Phys., B329, 445.
Cooper-Sarkar, A.M., Hulth, P.O., Hultqvist, K. & Sarkar, S., 1991. to be published.
Cowsik, R. & McClelland, J., 1973. Astrophys. J., 180, 7.
Davis, R.L. & Shellard, E.P.S., 1989. Nucl. Phys., B324, 167.
De Panfilis, S., Melissinos, A.C., Moskowitz, B.E. et al. , 1989. Phys. Rev., D40, 3153.
De Rujula, A., Glashow, S. & Sarid, U., 1990. Nucl. Phys., B333, 173.
Dicus, D.A., Kolb, E.W. & Teplitz, V.L., 1978. Astrophys. J., 221,327.
Dicus, D.A. & Roy, P., 1989. Phys. Rev., D42, 938.
Dimopoulos, S., Esmailzadeh, R., Hall, L.J., & Tetradis, N., 1990a. Phys. Lett., B247, 601.
Dimopoulos, S., Eichler, D., Esmailzadeh, R. & Starkman, G., 1990b. Phys. Rev., D41, 2388.
Dimopoulos, S., Tetradis, N., Esmailzadeh, R. & Hall, L.J., 1991. Nucl. Phys., B349, 714.
Dolgov, A.D., 1991. in Primordial Nucleosynthesis f1 Evolution of the Early Universe, ed.
Sato, K. & Yokoyama, J., Kluwer Academic.
Dolgov, A.D. & Zeldovich Ya.B., 1982. Rev. Mod. Phys., 53, 1.
Dydak, F., 1990. in Proc. 25th International Con/. on High Energy Physics, Singapore, ed.
Phua, K.K., World Scientific.
Efstathiou, G.F., 1990. in The Birth f1 Early Evolution of Our Universe, Nobel symposium 79,
Griiftavallen, OUAST/90/IB.
Ellis, J., Olive, K., Sciama, D.W. & Sarkar, S., 1988a. Phys. Lett., B215, 404.
Ellis, J., Flores, R. A., Freese, K. et al. , 1988b. Phys. Lett., B214, 403.
Ellis, J., Nanopoulos, D.V., Roszkowski, L. & Schramm, D.N., 1990a. Phys. Lett., B245, 251.
Ellis, J., Gelmini, G.B., Nanopoulos, D.V. & Sarkar, S., 1990b. CERN-TH.5853/90.
Enqvist, K., Kainulainen, K. & Maalampi, J., 1989. Nucl. Phys., B317, 647.
Fabian, A.C, Maylor, T.W. & Sciama, D.W., 1991. Mon. Not. R. astr. Soc, 249, 29p.
Fukugita, M., 1988. Phys. Rev. Lett., 61, 1046.
Gabbiani, F., Masiero, A. & Sciama, D.W., 1990. Phys. LeU. B, in press, SISSA 154/90/A.
102

Gelmini, G.B., Hall, L.J. & Lin, M.J., 1987. Nucl. Phys., B281, 726.
Gelmini, G.B., Gondolo, P. & Roulet, E., 1991. Nucl. Phys., B351, 623.
Giudice, G.F., 1990. Phys. Lett., B251, 460.
Giudice, G.F. & Griest, K., 1989. Phys. Rev., D40, 2549.
Gould, A., Draine, B.T., Romani, R.W. & Nussinov, S., 1990. Phys. Lett., B238, 337.
Griest, K. & Kamionkowski, M., 1990. Phys. Rev. Lett., 64, 615.
Guth, A., 1991. these proceedings.
Hagmann, C., Sikivie, P., Sullivan, N.S. & Tanner, D.B., 1990. Phys. Rev., D42, 1297.
Hall, L.J., 1988. in Proc. 16th SLAG Summer Institute on Particle Physics, p. 85, ed.
Brennan, E.C., SLAC-336.
Harari, D. & Sikivie, P., 1987. Phys. Lett., B195, 361.
Hime, A. & Jelley, N., 1991. Phys. Lett. B, in press, OUNP-91-01.
Hime, A., Ross, G.G., Phillips, R.J.N. & Sarkar, S., 1991. Phys. Lett. B, in press, OUTP-91-02P.
Huber, M.E., Cabrera, B., Taber, M.A. & Gardner, R.D., 1990. Phys. Rev. Lett., 64, 835.
Glashow, S.L., 1991. Phys. Lett., B256, 255.
Kolb, E.W. & Turner, M.S., 1990. The Early Universe, Addison Wesley.
Kolb, E.W. & Turner, M.S., 1991. FERMILAB-Pub-91/50-A.
Krauss, L., 1990. Phys. Rev. Lett., 64, 999.
Kuzmin, V., Rubakov, V. & Shaposhnikov, M., 1985. Phys Lett., B155, 36.
Lee, B.W. & Weinberg, S., 1977. Phys. Rev. Lett., 39, 165.
McLerran, L., Shaposhnikov, M., Turok, N. et al. , 1990. PUPT-90-1224.
Melott, A.L., McKay, D.W. & Ralston, J.P., 1988. Astrophys. J., 324, L43.
Nussinov, S., 1985. Phys. Lett., B165, 55.
Particle Data Group, 1990. Phys. Lett., B239, 1.
Primack, J.R., Seckel, D. & Sadoulet, B., 1988. Ann. Rev. Nucl. Part. Sci., 38, 751.
Raffelt, G.G., 1990. Phys. Rep., 198, 1.
Salati, P., 1991. Phys. Lett., B253, 173.
Sarkar, S., 1985. in Supersymmetry, Supergravity f1 Unified Theories, p. 465, ed. Furlan, G. et al. ,
World Scientific.
Sarkar, S., 1991. Rep. Prog. Phys., to appear, .
Sarkar, S. & Cooper, A.M., 1984. Phys. Lett., B148, 347.
Sato, N., Hirata, K.S., Kajita, T. et al. , 1990. KEK 90-166.
Sciama, D.W., 1990. Gomm. Astrophys., 15, 71.
Simpson, J.J., 1985. Phys. Rev. Lett., 54,1891.
Smith, P.F. & Lewin, J.D., 1990. Phys. Rep., 187,203.
Starkman, G.D., Gould, A., Esmailzadeh, R. & Dimopoulos, S., 1990. Phys. Rev., D4l, 3594.
Sur, B. et al. , 1990. LBL-30107.
Turner, M.S., 1990. Phys. Rep., 197, 67.
Witten, E., 1984. Phys. Rev., D30, 272.
ARE GALACTIC HALOS MADE OF BROWN DWARFS OR BLACK HOLES?

B.J.CARR
Astronomy Unit,
Queen Mary & Westfield College,
Mile End Road,
London El 4NS.

ABSTRACT. If galactic halos are baryonic, they comprise either brown


dwarfs or black holes. The first possibility can be tested by microlensing
observations, the second by searches for an infrared background.

1. INTRODUCTION

The most recent cosmological nucleosynthesis arguments 1 indicate that


the baryonic density parameter must lie in the range 0.04 to 0.06 for a
Hubble parameter of 50, whereas the density associated with visible
galaxies is only 0.01. This indicates that a large fraction of the baryons
must be dark. It is possible that the dark baryons are in a hot
intergalactic medium, although the Gunn-Peterson test and COBE
constraints on the Compton distortion of the microwave background
require that its temperature be in the range 10 4 to 10e K. More likely,
the baryons have been processed into the dark remnants of a first
generation of "Population III" stars. In this case, their density could
suffice to explain the dark matter in galactic halos and the term
"Massive Compact Halo Object" or MACHO has been coined in this context.
On the other hand, unless one invokes fluctuations at the quark-hadron
phase transition 2, one would still need non-baryonic dark matter to
explain the critical density required by inflation and the term "Weakly
Interacting Massive Particle" or WIMP is then used. The Universe may
therefore need both machos and wimps.
The relative distribution of the baryonic and non-baryonic dark
matter depends on the epoch at which the Population III stars form and
on whether the wimps are hot or cold. If they are hot, then halos
consist exclusively of machos because the wimps cannot cluster. If they
are cold, then halos contain both machos and wimps. In this case, if the
Population III stars form before galaxies, one might expect their
remnants to be distributed throughout the Universe with the wimp-to-
macho density ratio being the same everywhere and of order 10. On the
other hand, if they form at the same time as galaxies, perhaps in the
first phase of proto galactic collapse, one would expect the machos to be
confined to halos and clusters. In this case, their contribution to the
halo density could be larger since the baryons would probably dissipate
and become more concentrated. These possibilities are illustrated in Fig.l.

103
T. Shanks et al. (etis.). Observational Tests of Cosmological Inflation. 103-109.
© 1991 Kluwer Academic Publishers.
104

COSMOL061 CAL ____ Quc.rk-hadron __ ~ __ NON-BARYONIC


fluctuations?
NUCLEOSYNTHESIS DARK MArTER
(0.010(.A bh2< 0.C16)

Hot
/~
Cold

Prega I act i c WIMPS+~CHOS


MACI!OS
Pop III (10:1)
BARYONIC ,--..-::r------ Ptotogalactic WIMPS+MACHOS
Pop III MACHOS
(3:1)
~
><
DARK MATTER
Hot 16M WIMPS

FIGURE (1): halo contribution from wimps and machos in various models

A variety of arguments 3 show that the Population III objects must be


either the black hole remnants of "Very Massive Objects" (VMOs) larger
than 100 Me which collapse during their oxygen-burning phase or
"brown dwarfs" which are too small to burn hydrogen at all. Both
situations require the Population III stars to have a very different mass
spectrum from the ones forming today. We do not really understand why
this should be, or indeed which situation is more plausible, but the
evidence from starburst galaxies and cooling flows suggests that both
situations may be possible. Here we will consider the observational
consequences of the two scenarios.

2. BROWN DWARFS

X-ray observations suggest that the cores of many clusters contain hot
gas which is flowing inwards because the cooling time is less than the
Hubble time 4. The mass flow rate varies from a few ~1a y-l to 103 M1:) y-l
and the mass appears to be deposited over a wide range of radii.
However, the gas cannot be forming stars with the same mass spectrum
as in the solar neighbourhood, else the central region would be bluer
and brighter than observed. This suggests that the cooling flows
produce very low mass stars, possibly because the high pressure
reduces the Jeans mass. The important feature of a cooling flow is that
it is quasi-static, in that the cooling time exceeds the local dynamical
time and it is this condition which preserves the high pressure.
Although cooling flows provide a natural way of turning gas into low
mass stars with high efficiency, those observed in cluster cores could
not generate dark halos for galaxies outside clusters. However, one could
expect analagous high pressure quasi-static flows to occur at earlier
cosmological epochs and these would have been on much smaller scales
than clusters s • This is best illustrated for the hierarchical clustering
scenario, in which as time proceeds increasingly large gas clouds bind
and virialize. The mass fraction of a cloud cooling quasi-statically is
maximimized when the cooling time tc is comparable to the free-fall time
If: collapse does not proceed at all for tc»lf, whereas it is not
quasi-static for tc«lf. In any particular variant of the hierarchical
clustering scenario, one can specify the mass binding as a function of
redshift. For a cloud of mass M, the dynamical time will just be of order
the Hubble time at that redshift, whereas the cooling time will depend
105

upon the density and virial temperature of the cloud (which are
themselves determined by M and z). Thus one can specify a region in
the (M,z) plane of Fig.2 in which bound clouds will cool within a
dynamical time. This applies above a lower mass limit associated with
molecular hydrogen or Lyman-IX cooling and below an upper mass limit
associated with atomic hydrogen cooling. The condition tc ~tr will be
satisfied at the boundary of the region (shown shaded) and the
intersection of this boundary with the binding curve M(z) singles out
two characteristic mass-scales and redshifts. These correspond to what
are termed "Pervasive Pre galactic Cooling Flows" (PPCFs) since the
amount of gas cooling quasi-statically is maximimized. The associated
mass-scales are always of order 104-lOs~ and 1011~ but the redshifts
depend on the particular scenario. Fig.2 shows the binding curves
corresponding to the Cold Dark Matter (CDM) scenario, the broken curve
corresponding to the biased version, and the isocurvature clustering
scenario (with 1"\ specifying the exponent in the mass dependence of the
density fluctuations at decoupling).
One might anticipate most of the dark matter being made on the
smaller scale since much of the gas will have been consumed by the time
atomic cooling becomes important. However, this does not apply in the
Cold Dark Matter picture because the spectrum of fluctuations is very
flat on subgalactic scales. Also, in most isocurvature models, M(z) is
never small enough for low mass PPCFs to occur after decoupling 6 • Both
these features are indicated in Fig.2. This suggests that most of the
dark matter would need to be made by high-mass PPCFs. However, the
problem with this is that most of the baryons will by then have gone
into clouds with tc<tf and such clouds should make ordinary stars
because they would not fragment at high pressure. One way round this
is to invoke supernovae to reheat the gas so that most of it can avoid
cooling until the proto galactic epoch7. Another is to argue that even
clouds with tc <tr can make a lot of dark matters. The idea here is that
gas always drops out at such a rate as to preserve the PPCF condition
tc~tf for the surviving gas. One thus gets a two-phase medium, with cool
dense clouds embedded in hot high pressure gas. This was originally
proposed as a mechanism to make globular clusters at a proto galactic
epoch 9 but one can argue that sufficiently small clouds would fragment
into dark clusters rather than visible clusters, at least in the presence
of molecular hydrogenS.
One of the most useful signatures of clusters of brown dwarfs would
be their gravitational lensing effects. Data from m8crolensing, the
multiple-imaging of a distant source like a quasar, already provide
evidence for dark halos, since the image separations usually require a
lens mass exceeding the visible galaxy mass. Observations of
microlensing, the intensity variations in one of the images due to
individual halo objects within the lensing galaxy, may provide evidence
for brown dwarfs with mass down to 10-4M0. Indeed, such evidence may
already exist 10 for the quasar 2237+0305. Here the timescale of variation
requires a lensing mass in the range 0.001 M0 to 0.1 ~, although this
conclusion is tentative since the variable image is almost exactly aligned
with the centre of the lensing galaxy, where the density should be
dominated by ordinary stars rather than dark matter.
106

decoupling

"'":l
N

....+
bO
o
......

12 14

log (M!Ma)
FIGURE (2): indicating the (M,z) region in which a cloud cools within a
Hubble time. PPCFs occur where the binding curve intersects this region.

Another possibility is to seek microlensing by objects in our own


halo by looking for intensity variations in stars in the Large Magellanic
Cloud l1 • In this case, the timescale for the variation is smaller, so one
can seek for lensing masses all the way down to 10- 9 Me. However, the
probability of an individual star being lensed is very low, so one has to
look at many stars for a long time. Such observations are already
planned and could soon confirm or eliminate the brown dwarf model.

3. THE VMO SCENARIO

The most striking signature of the VMO scenario would be background


radiation generated during the stars' hydrogen-burning phase, so we
begin by summarizing the observational limits on the background light
intensity I(A) in various bands. These are conveniently expressed in
critical density units by defining a quantity 0RIA)=4rrAI(A)/c3Pcrit. For
comparison, the microwave background radiation (t.1BR) peaks at
Apeak=1400~ with a densitv 0R=2xlO- s h- 2 where h is the Hubble
parameter in units of 100 km/s/Mpc. The limits are summarized in Fig.3.
The current FIRAS results 12 from COBE imply that the MBR is so well
fit by a black-body spectrum that any extra background must have an
intensity less than 1% of the peak MBR intensity over the range
500-5000~. This implies a constraint 0R(A)<2xlO- 7 h- 2 IA/ A.peak)-1 and the
limit could eventually be strengthened by a factor of 10. The DIRBE
results at the south ecliptic pole 13 give upper limits in the J, K, L, M,
12~, 25~, 60,-" 100~, 120-200,-, and 200-300,-, bands. However, the limits
indicated in Fig.2 are very conservative since they do not include any
subtraction for the foreground backgrounds from interstellar and
interplanetary dust. Stronger but less firm constraints in the far-IR
band come from the most recent analysis 14 of the Nagoya-Berkeley
data 1S • After careful modelling of the foreground contributions, one gets
0R(275,-,)<7xlO- 7 h- 2 , 0R(135,-,)=(1.4 0.4)x10- 6 h- 2 and 0R(100,-,) <8x10- 7 h- 2.
The 135~ result may contain an isotropic component, although it would
not necessarily be extragalactic.
107

-.
I ...
T

••
-.
~ ~---r
Dlut rllM II.AS If'

1]
L
L __ •
-s

-. ,I
;- Tl t

§
T
I
I
,J..
+

,,'" '"
... ,, ,
," ,
/
, '" ,
\
,,
-,
.- .- ,
, ,
,
\

.- .- , ,
\ \

/ V1'tO U~ClOUND \
/
-I
z.: bO
I I I I I I I I I

1.0 l.S t.S t.o o.s

FIGURE (3): comparing the background radiation limits from DIRBE,


FIRAS, IRAS and NB with the background expected in the VMO scenario

Some of the IRAS limits are also shown in Fig.:;. At one stage IRAS
data seemed to indicate 16 a 1001-1 background with 0R(1001-l1=3x10- 6 h- 2 but
this is inconsistent with both the Nagoya-Berkeley and DIRBE results.
The DIRBE and IRAS limits are very weak at 121-1 and 251-1, about 1O- 4 h- 2 ,
because the interplanetary dust emission is so large. Indeed it may
never be possible to get good limits in this waveband without a satellite
outside the Solar System. In the near-IR, a Japanese rocket experiment 17
gives a limit 0R(1-51-1) <3x10- s h- 2 • with the possible detection of a "line"
at 2.21-1 with 0R(2.21-1)=3xlO- 6 h- 2 • However, this claim is controversial
because of the problem of subtracting starlight and rocket exhaust, and
the line interpretation is particularly insecure.
Let us now compare these limits with the background expected from
a population of VMOs. This can be predicted rather precisely 19 since all
VMOs have a surface temperature T s of about lO s K and generate
radiation with efficiency "-=0.004. We assume that the VMOs have a
density parameter O*=0.02h- 2 =0.lh so-2, corresponding to the sort of
density required for galactic halos, and that they produce black-body
radiation with temperature Ts' If the radiation is affected only by
cosmological redshift, it should then have a peak wavelength and density

(1)

at the present epoch, where z* is the redshift at which the VMOs burn.
We can place an upper limit on z* by noting that the main-sequence time
of a VMO is tMS=2x10 6 y (independent of mass I, so that z* cannot exceed
the redshift when the age of the Universe is tMS' This implies
z*<240h- 2 / 3 and hence >'peak<151-1 and 0R>5x10- 7 h- 2 • We can place a lower
limit on z* from UV-IR background light limits. As discussed by
McDowell 19 , these imply a constraint on the density of VMOs burning at
108

any redshift z*. If one requires ~=O.lhso -z, this places a lower limit on
z*, mainly because one needs the radiation to be redshifted into the
near-IR band, where the background light limits are weaker. In the
absence of neutral hydrogen absorption, one requires z*>30 and this
implies ),peak>l~ and 0R<3x10- 6 h- 2 • The peak of the VMO background
must then lie somewhere on the heavy line in Fig.2 and the spectrum
must lie within the region to the right of the broken line. If the
radiation from the VMOs were absorbed by neutral hydrogen, the
spectrum would be cut off shortward of 9[( l+z*) /100] )~, with most of the
radiation appearing in redshifted recombination lines. The UV-IR limits
would then allow z* to be as low as 4 and the limit on OR would be a
factor of 6 higher than indicated above.
Let us first ask whether the direct light from VMOs could explain
the background reported at 2.2~. If this is a line, as claimed, then it
would be natural to identify it with the redshifted 1200A Lyman-<x line.
In this case, the redshift can be specified uniquely as z*=16 and the
predicted density is 0R=6xlO- s h- 2 • To compare this with the observed
background (OR=3x10- S h- 2 ), one must multiply by the ratio of the
bandwidth of the actual line to that of the observations. The observed
background is too small unless the line is exceedingly narrow but that
would require the VMOs to form over a very narrow range of red shifts.
If the claim that one needs a line is unjustified, one should identify
),peak rather than the redshifted Lyman-<x wavelength with 2.2~. One
then needs z*=50 and hence 0R=2xlO- s h- 2 , which fits the data somewhat
better. If the 2.2~ background is not confirmed, Fig.3 shows that one
could still hide the VMO light at around 10~
We now consider the effects of dust. One possibility is that the light
from VMOs is reprocessed by dust inside the clouds where they form (cf.
IRAS galaxies where most of the radiation emerges at 100~). However, this
would still give too large a far-IR background if the VMOs were
proto galactic , so VMOs numerous enough to explain galactic halos would
need to be pre galactic. We therefore consider the possibility that the
VMO light is reprocessed by cosmological rather than local dust. Such
dust could either be pre galactic in origin or confined to galaxies
themselves if galaxies cover the sky. If the dust cross-section for
photons of wavelength), is assumed to be geometric (TTrdz for a grain
radius rd) for ),«rd but to fall off as ),-1 for :>"»rd' then the spectrum
should peak at a present wavelength ZO

(2 )

where zd is the epoch of dust production and we have used eqn (1) to
express OR in terms of z*. The crucial point is that the wavelength is
very insensitive to the various parameters appearing in eqn (2) because
the exponents are so small. At one time the Nagoya-Berkeley experiment
appeared to find a submillimetre excess peaking at almost exactly the
wavelength predicted. However, the Nagoya-Berkeley excess has now been
disproved and the question arises of whether the VMO-plus-dust scenario
is still compatible with COBE. Since the FIRAS limit extends all the way
down to 500~, one clearly needs rd to be as small as possible. However,
I'd could hardly be less than O.Ol~, so the situation is clearly marginal.
109

In order to examine the issue in more detail, Bond et al. 21 have


carried out a more sophisticated analysis, in which the dust
cross-section is assumed to scale as A-O: for A>rd where 0: may be
wavelength-dependent. They also introduce a more realistic model for the
source luminosity history, allowing for both "burst" and "continuous"
models. Comparison with the far- IR and COBE constraints is shown in
Fig.3 for models with z*=100, zd=10 and rd=O.OlJ..l. The broken and dotted
curves have 0:=1 and 0:=2, respectively, while the upper and lower ones
have 0d=lO-s and 0d=10- 6 • This shows that, until the FIRAS constraint
improves, the VMO-plus-dust scenario is still viable for 0d=10- 6 •
If the COBE results do eventually eliminate the VMO-plus-dust
scenario, it should be stressed that one does not necessarily expect dust
reprocessing anyway. Pre galactic dust with density 0d would only absorb
UV photons for Zd>10(Od/lO-S)-2/3(rd/O.1J..l)2/3, where 0d is normalized to
the sort of value appropriate for galaxies. The dust in galaxies
themselves could suffice only if galaxies cover the sky and this requires
the redshift of galaxy formation to exceed about 10. Even if galaxies do
cover the sky, the analysis of Fall et al. 22 indicates that the dust-to-gas
ratio in primordial galaxies may only be in the range 1/20 to 1/4 that of
the Milky Way for 2<z<3, which makes the opaqueness condition difficult
to satisfy. Therefore, if COBE does eventually eliminate the predicted
far- IR background, the VMO light could still reside in the near- IR.

REFERENCES

1. Olive, K.A., Schramm, D.N., Steigman, G. & Walker, T., 1990,


Phys.Lett.B., 236, 454.
2. Applegate,J. & Hogan, C.J., 1985, Phys.Rev.D., 31, 3037.
3. Carr, B.J., 1990, Comm.Astrophys., 14, 257.
4. Fabian, A.C., Nulsen, P.E.J. & Canizares, C.R., 1984, Nature, 320, 733.
5. Ashman, K.M. & Carr, B.J., 1988, Mon.Not.R.Astr.Soc., 234, 219.
6. Ashman, K.M. & Carr, B.J. (1991). Mon.Not.R.Astr.Soc., in press.
7. Thomas, P. & Fabian, A.C., 1990, Mon.Not.R.Astr.Soc., 246, 156.
8. Ashman, K.M. (1990). Mon.Not.R.Astr.Soc., 247, 662.
9. Fall, S.M. & Rees, M.J. (1985). Astrophys.J., 298, 18.
10. Irwin,t-1.J., Webster,R.L. & Hewett, P.C. (1989). Astron.J., 98, 1!j89.
11. Paczynski, B. (1986). Astrophys.J., 304, 1.
12. Mather, J.C. et al., 1990, Ap.J.Lett., 354, L37.
13. Mather, J.C. et al., 1991, Ap.J., in press.
14. Lange, A.E. et al., 1991, Ap.J., in press.
15. Matsumoto, T. et al., 1988, Ap.J., 329, 567.
16. Rowan-Robinson,t-1., 1986, Mon.Not.R.Astr.Soc., 219, 737.
17. Hatsumoto, T., Akiba, M. & Murakami, H., 1988, Ap.J., 332, 575.
18. Bond, J.R., Arnett, W.D. & Carr, B.J., 1984, Ap. J., 280, 825.
19. HcDowell, J.C., 1986, Mon.Not.R.Astr.Soc., 223, 763.
20. Bond, J.R., Carr, B.J. & Hogan, C.J., 1986, Ap.J., 306, 428.
21. Bond, J.R., Carr, B.J. & Hogan, C.J., 1991, Ap.J., 367, 420.
22. Fall, S.M., Pei, Y.C. & McMahon, R.G., 1989, Ap.J.Lett., 341, L5.
AGES OF GLOBULAR CLUSTERS

PIERRE DEMARQUE, CONSTANTINE P. DELIYANNIS, AND ATA SARAJEDINI


Department of Astronomy,
Center for Solar and Space Research, and
Center for Theoretical Physics
Yale University
P.O. Box 6666
New Haven, CT 06511, USA

ABSTRACT. A brief historical introduction is given. Recent advances in precision photometry of faint
stars in globular clusters and in the modeling of turnoffs and synthetic horizontal branches, indicate that
there is a spread in age of at least 4 Gyr among Galactic globular clusters. Absolute ages are more difficult
to determine and the sources of uncertainty are discussed. On the observational side, chemical abundances,
[Fe/H] and [OIFe], are the major sources of uncertainty. Two effects hitherto ignored in evolutionary
calculations and currently under study, rotation and diffusion, are discussed. Internal rotation affects the ages
only in a minor way. The diffusion of helium near the main sequence could be important in reducing the
derived ages of the oldest clusters; its efficiency, which is poorly known, can however be constrained and
evaluated by lithium abundance observations. Finally, an approximate chronology of the Galaxy is
presented on the basis of our current best understanding of cluster dating. The ~ age derived for the
cluster M92 is 14 Gyr, which puts a strong constraint on the cosmological timescale.

1. Brief Historical Introduction on Stellar Dating

Let me start with a few historical comments. The first step was of course the recognition that it is
nuclear fusion in the cores of stars that drives the slow phases of stellar evolution. An equally
important development in our ability to date star clusters is the realization by Opik (1938a,b) that
stars do not evolve mixed, and that it is the chemical inhomogeneity due to the conversion of
hydrogen into helium in the core that is responsible for the evolution of dwarf stars into red giants.
The identification of the main sequence turnoff with core exhaustion is the basis of our cluster
dating techniques. Note that this point had not yet been generally recognized when Chandrasekhar
(1939) wrote his classic monograph. But not long afterwards, Opik's idea was further investigated
by ScMnberg and Chandrasekhar (1942), who explored the properties of stellar models near the
main sequence with a progressively larger isothermal helium core. They found that there is a limit
to the growth of the fractional mass of such a core, the well-known ScMnberg-Chandrasekar eSC)
limit. The SC-limit is only approximately reached in real stars, because its derivation assumes that
the perfect gas law holds throughout the star, and ignores the possible contribution from
gravitational contraction of the core. However, for the first time, the existence of the SC-limit did
permit a realistic quantitative estimate of the main sequence lifetime of a star, given its mass and
chemical composition.
The SC-limit provided the time scale calibration in the historic studies of the

111
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 111-129.
© 1991 Kluwer Academic Publishers.
112

color-magnitude diagram of the globular cluster M3 (Sandage 1953) and the old disk cluster M67
by Sandage (1957). The use of the SC-limit, together with the early evolutionary tracks of Sandage
and Schwanschild (1952), which considered the evolution beyond the SC-limit phase, formed the
basis of Sandage's semi-empirical approach to stellar evolution, making use of the cluster
luminosity function to evaluate relative evolutionary timescales along the observed giant sequence in
the color-magnitude diagram.
Hoyle and Schwarzschild (1955) provided the first correct theoretical account of the
evolution oflow-mass stars from the main sequence to the onset of helium burning at the tip of the
giant branch. Here, the giant branch is represented by a sequence of static models whose helium
core mass increases as the luminosity increases. Electron degeneracy is included in the isothermal
helium core, which is surrounded by a thin hydrogen-burning shell and an extended convective
envelope. In this paper, Hoyle and Schwanschild explain for the first time why the giant branch
position (now often called the Hayashi line) is a function ofmetallicity. Improvements in the
models, including the use of the mixing length theol)' in the convective envelope and more realistic
surface boundal)' conditions were made by IGppenhahn, Temesv3l)' and Biermann (1958), and by
Demruque and Geisler (1963). The latter further explored the dependence of the giant branch
position on chemical composition, including the helium content
The first automatic computations of stellar evolution around the turnoff and up the giant
branch were performed by Hoyle (1959), using a technique developed by Haselgrove and Hoyle
(1956a,b). As in the previous calculations, these models were constructed by the fitting of inward
and outward integrations (the shooting technique). Soon after, the development of more powerful
computers made the Henyey method the preferred technique for stellar evolution calculations
(Henyey et al. 1959; Larson and Demruque 1964).
The age derived for M3, then the best studied globular cluster, increased rapidly during that
brief period. Sandage's age for M3 was 5 Gyr (1953) and the Hoyle-Schwarzschild (1955) age
was 6 ± 2 Gyr, both uncomfortably close to that of the Sun at 4.5 Gyr. The age derived by
Sandage (1962) on the basis of the Hoyle (1959) models was 26 Gyr. The Hoyle models had
radiative cores and did not make use of the SC-limit, and the nuclear energy generation rates in the
models were explicitly calculated. Two factors later reduced the Hoyle (1959) age scale: (1) a
more realistic formulation of the energy generation rates (Hoyle used simple power laws), and (2)
a helium abundance closer to that of the Sun (about 0.25 by mass; Demruque 1964, 1966, 1967).
This latter point is important and deserves some comment. Until the discovel)' of the cosmic
microwave background, there was little guidance as to what the helium abundance, hereafter Y, of
the globular clusters should be. Low values ofY, around 0.10 or less were generally assumed.
Early calculations by Gamow of big bang nuc1eosynthesis that predicted Y=OA (Tayler 1964) were
largely ignored. Note that evolution3l)' tracks produced in the mid-1960's yielded 17 Gyr for the
age ofM3 when one assumed Y=D.25 (Demruque 1964).
The first age estimates were based on a few short model sequences, which at the time
represented a major computational task.. Since then, a great deal of work has been done by many
investigators to improve the physics of standard stellar models and the calculation of stellar
evolution3l)' tracks (lben 1967, 1974, and references therein; Mengel et al. 1979; VandenBerg
1985; Demruque et al. 1988 and references therein; Straniero and Chieffi 1991). Ages are
determined with the help of detailed isochrones constructed from grids of stellar evolutionary tracks
(Ciardullo and Demruque 1977; VandenBerg and Bell 1985; Green et a1.1987). In addition, the
shapes of the isochrones themselves are now much more faithful to the observations. However,
the standard theoretical calibration of the ages of the oldest clusters has not changed appreciably in
the last twenty five years. Differences in the age scales of different authors can be traced primarily
113

to differences in the adopted chemical abWldances for individual globular clusters or in the fitting
procedures (Demarque and McClure 1977; Demarque 1980; Janes and Demarque 1983; Sandage
1986; Zinn 1986; King et al. 1988; Sarajedini and King 1989). The systematics of these
uncertainties are now much better understood (Sarajedini 1991). Precision photometry is making
spectacular progress with ceo detectors. Recent theoretical advances related to the modeling of the
horizontal branch and to the detailed physics of evolving halo stars are very promising. Already, it
has proven possible to establish with some confidence the existence of an age spread among halo
globular clusters (§ 3), and a much sharper picture of the relative chronology of our Galaxy is
emerging.
The following sections will review the efforts currently Wlderway to improve our ability to
date individual globular clusters. Finally, for the cosmologist. we summarize the chronology of the
Galaxy, and present our best current estimate for the ages of the oldest globular clusters.

2. Dating Techniques
2.1 CLUSTER PARAMETERS

Determining the age of a globular cluster usually requires knowledge of several parameters, and in
what follows, we consider each in tum.
(1) By far the most straightforward to deal with is the helium mass fraction (y) of the stars
that comprise globular clusters. Since the Galactic globulars formed early in the history of the
Galaxy, it has become a custom to ascribe to them the primordial value of Y as determined via
observations of extragalactic HII regions (Davidson and Kinman 1985), the standard model of big
bang nucleosynthesis (e.g. Boesgaard and Steigman 1985), the R method (Buzzoni et al. 1983),
and models of the horizontal branch (hereafter HB; Lee, Demarque, and Zinn 1990, and references
therein). This adopted value for Y is usually 0.23 or 0.24; however, to the degree that Y is
Wlcertain ( - 0.01- 0.02), virtually all of the results presented here are relatively insensitive to the
choiceofY.
(2) Another cluster parameter important to the determination of the age is the cluster heavy
element (metal) abWldance as represented by the observed iron abWldance, [Fe/II]. In recent years,
the controversy regarding the globular cluster metallicity scale, which existed previously, has been
somewhat resolved. Many authors who conduct surveys of the globular clusters now use the
abundances derived by Zinn and West (1984). The majority of halo clusters have [Fe/H] values
from -1.0 to -2.3 dex with a typical uncertainty of 0.15 dex. The determination of the absolute
[Fe/H] is usually achieved using line strength analysis along with model atmospheres, which is
rather uncertain As with other derived parameters, the absolute metallicities are more uncertain
than relative ones.
(3) The morphology of cluster main sequence turnoffs (and therefore cluster dating) is also
greatly dependent on the oxygen abundance. If the controversy regarding the cluster [Fe/H] scale
has been more or less resolved, the problem of the oxygen enhancement in cluster stars, as
measured by [O/Fe], is still an open question. That is to say, given [O/Fe] = 0 for the Sun, is the
[O/Fe] of halo stars significantly different, and if so, does it vary with [Fe/H] ? The problem arises
when one wants to obtain spectra of the faint cluster turnoff stars (generally no brighter than about
V= 16) to measure the very weak oxygen lines. As a result, various authors have adopted two
different strategies both of which require a leap of faith. First, observations of giant stars in 13
globular clusters by Pilachowski et al. (1983) reveal a mean value of [O/Fe] = 0.25, regardless of
114

[Fe/H]. Gratton and Ortolani (1989) studied giant stars in 11 clusters and found [O/Fe] '" 0040;
therefore, the giant stars in at least some globulars are enhanced in oxygen relative to the Sun, and,
more importantly, show no trend of [O/Fe] with cluster [Fe/H]. However, are the oxygen
abundances of the giants in a given cluster indicative of the dwarf star abundances? Second, most
observations of field dwarf stars, which are comparatively brighter than cluster dwarfs, show a
relatively constant level for [O/Fe] of", 0.40 dex (Bond and Luck 1988; Gratton 1990) and thus no
variation with [Fe/H]. However, the work of Abia and Rebolo (1989) on the oxygen abundances
of the field dwarfs shows a marked variation of [O/Fe] with [Fe/H]; they find [O/Fe] '" +1.2 at
[Fe/H] =-2.3 and [O/Fe] '" +0.6 at [Fe/H] =-1.0. Such a large slope, if valid, may have
significant consequences for the chemical enrichment and formation history of the Galactic halo (see
below). However, are the oxygen abundances of field dwarf stars indicative of those in cluster
dwarfs? Regardless, there is now general agreement that the main sequence stars in the Galactic
halo are enhanced in oxygen relative to the solar value. The question of whether the oxygen
abundances of the metal-poor ( [Fe/H] - -2 ) systems is greater than those of metal-rich
( [Fe/H] - -1 ) systems is still the subject of vigorous research. In general, investigators adopt
either [O/Fe] '" 0.0 or +0.5 to +0.7 when dating globular clusters.
(4) As with the oxygen abundance problem, the question of the globular cluster distance
scale has also been the object of intense controversy. The distance modulus of a globular cluster
reduces to comparing the apparent magnitude of the cluster RR Lyrae variables or horizontal branch
(HB) stars (V(RR) or V(HB), respectively) to the absolute magnitude of these same stars
(Mv(RR) or Mv(HB), respectively). The values of V(RR) and V(HB) are relatively
straightforward to derive observationally (but see below). Theoretically, there are two alternatives
for the dependence of Mv(RR) on cluster [Fe/H] , from which Mv(RR) is computed. (a) The
theoretical work of Lee et al. (1990, and references therein) based on synthetic HB calculations
with solar [O/Fe] has suggested a value of dMv(RR) /6[Fe/H] = 0.17. This value for the slope of
the Mv(RR) '" [Fe/H] relation is supported by the observational work of Cacciari et at. (1989), Liu
and Janes (1990), Fusi Pecci et al. (1990), and Sarajedini and Lederman (1991). (b) Sandage
(1990, and references therein) has analyzed the period behavior of cluster RR Lyraes and concluded
that LlMv(RR) 16[Fe/H] = 0.39. This value is supported by the observational work of Buonanno
et at. (1989) and Longmore et at. (1990). As in the case of the oxygen abundance controversy,
conclusions regarding the history of the Galactic halo strongly depend on the precise form of the
relation.
(5) The inclusion of helium diffusion in stellar evolution models has shown that isochrone
turnoffs which include the effects of diffusion are redder than those that do not (§4.2). As a result,
ignoring helium diffusion overestimates the actual age. Therefore, diffusion has become another
parameter that potentially affects age estimates for globular clusters. At this time, nearly all
published ages for clusters do not include helium diffusion..
(6) Finally, some knowledge of the cluster reddening is required. In general, there are
enough low-reddening ( E(B-V) < 0.1 mag) globular clusters with a range of metallicities up to
[Fe/H] '" -1.0 to use in tracing the evolution of the Galactic halo. The typical error in a reddening
measurment ofless than 0.1 mag in E(B-V) is about 0.02 mag, and usually increases as the
reddening itself increases.

2.2 ISOCHRONE FITTING

In order to fit theoretical isochrones to photometry of a globular cluster, we must first select the
highest quality photometry, most likely done with a CCD detector. Then, isochrones of various
115

ages and appropriate Y. [Fe/H] • and [O/Fe] for the cluster are superimposed on the photometry in
the color-magnitude plane after being shifted in magnitude by the adopted distance modulus and in
color by the reddening. Thus. to do isochrone fitting. all of the parameters discussed above are
necessary. and this. in fact. is a major weakness of the teclmique. In addition. rarely is it clear
from the comparison of the isochrones to the photometry exactly which isochrone best fits the data.
However. all of the adopted parameters have errors associated with them. Therefore. parameters
such as [Fe/H] and distance modulus. whose errors contribute most to the quality of the isochrone
fit, are varied slightly within the typical error until an optimum match is achieved between the
photometry and the theoretical isochrones. Therefore. deriving ages via isochrone fitting is one of
the most uncertain methods we will describe.
Most studies that have attempted to apply the isochrone fitting teclmique have assumed that
[O/Fe] = O. This is fine if one is concerned only with the relative ages and is convinced that [O/Fe]
does not vary significantly with [Fe/H). Such studies find ages anywhere from 10 - 12 Gyr for
the cluster Palomar 12 ([Fe/H) =-1.1. Sarajedini and King 1989. Straniero and Chieffi 1991) to
16 - 18 Gyr for clusters like M92 and M68 ([Fe/H) == -2.1. King et al. 1988. see also McOure et
al. 1987. Stetson and Harris 1988. Stetson et al. 1989. Straniero and Chieffi 1991). If [O/Fe) is
enhanced at a constant level regardless of [Fe/H) • these ages would be reduced by 1 to 2 Gyr
(VandenBerg et al. 1990. and references therein). If [O/Fe) varies significantly with [Fe/H) • then
the conclusions are less clear. In addition. the most recent WOlX regarding the effect of helium
diffusion on isochrone fitting indicates a possible age reduction of 2 to 3 Gyr (§4.2).

2.3 tN METHOD

The 11V method of cluster dating depends on the fact that the magnitude of the main sequence
turnoff becomes fainter as a cluster evolves, although the HB luminosity remains unchanged.
Figure 1 illustrates the quantity 11V. which is defined to be the V magnitude difference between the
turnoff and the HB (usually the RR Lyrae variables) as measured at the color of the turnoff. The
HB shown in Figure 1 is schematic and is referred to the scale of the upper abscissa. The solid-line
isochrone is for an age of 10 Gyr and the dashed-line isochrone is for 20 Gyr. It is thus clear that
11V increases as a cluster ages. The fmal step is to calibrate observed values of 11V with theoretical
ones in order to derive ages.
The 11V method is attractive observationally because it is independent of reddening and
attractive theoretically because the luminosity of the turnoff, which is mostly a function of the
conditions in the stellar core, is well detennined. There is the additional advantage that the mean
magnitude of the RR Lyrae variables and the turnoff are probably both weak: functions of [O/Fe]
and the effects of helium diffusion. Conversely, because some clusters do not have RR Lyraes and
some clusterHB's are simply not horizontal, detennining the value ofV(RR) or V(HB) can be an
uncertain process with typical errors of ±O.I 0 mag. Furthennore, the turnoff (TO) region is
essentially vertical in the color-magnitude plane thus making the value of V(TO) uncertain as well
with typical errors of ±O.lO mag so that the error in observed 11V becomes ±O.14 mag. Finally,
there is the above-mentioned controversy regarding the absolute magnitude of the HB stars and
how it varies with [Fe/H). Ideally, Mv(RR) and Mv(TO) are derived from theory to fonn a
calibration between 11V and age at each metallicity. Then the observed values of 11V and cluster
[Fe/H] are used to derive an age.
It is now well established that the values of 11V are not correlated with cluster metallicity
(Gratton 1985; Peterson 1987; Sarajedini and King 1989; Buonanno et al. 1989), but there is
some scatter in 11V at a given metal abundance. This does not mean however that there is no
116

relation between age and metallicity among the globular clusters. In the discussion above. we
mentioned that the slope of the Mv(RR) - [Fe/H] relation has important implications for the
fonnation scenario of the Galactic halo because the slope sets the scale of the relative ages. In
particular. since ~v(TO) I ~[Fe/H] =- 0.40. if ~v(RR) I MFe/H] = 0.39. then the ~V method
will yield no significant age-metallicity relation among the clusters considered. On the other hand.
if ~v(RR) I ~[Fe/H] = 0.17. then a significant relation between age and metallicity will result
(Zinn 1986; Sarajedini and King 1989; Lee etal. 1990; Sandage and Cacciari 1990). In addition,
the zeropoint of the Mv(RR) - [Fe/H] relation is most important in setting the scale of the absolute
ages; unfortunately. it is not well detennined.
Numerous investigators have applied the ~V method to various data sets and have made
conclusions regarding the age range of the Galactic globular clusters. Because both the slope and
zeropoint of the HB luminosity relation are uncertain. there is as yet no general consensus from the
~ V method regarding the absolute ages of the globular clusters.

2.4 L\(B-V) METHOD

The ~(B-V) method is a relatively new invention in the area of cluster dating (Sarajedini and
Demarque 1990; VandeoBerg et al. 1990). Figure 1 defines ~(B-V) to be the color difference
between the turnoff and the base of the giant branch. This decreases as the cluster age increases.
As with ~ V. observed values of ~(B-V) can be calibrated with theoretical ones to yield ages.

0.0 0.2 0.4 0.6

/
/
/
/

l
I
2 L\V . ,I 2
, , '/

3
L\(B-V) ~I/ 3

C-!
1

5 5
Isochrones
6 lOGyr 6
--- 20Gyr

0.2 0.4 0.6 0.8


(B-V)o

Figure 1. Definitions of the ~V and ~(B-V) quantities used for globular cluster dating. in the
absolute magnitude - intrinsic color plane. The horizontal branch luminosity stays constant with
time (note the shifted scale for the schematic HB). whereas the turnoff becomes fainter and redder.
Thus ~V increases and ~(B-V) decreases with age.
117

TIle advantages of A(B-V) are that it is independent of distance, reddening, photometric


zeropoint, helium abWldance, and largely insensitive to variations in [Fe/H]. Measuring the
A(B-V) of a cluster is also a relatively simple process; if all clusters in a sample are measured in
the same manner and similar to the theoretical values used in the calibration, no measurement-
related problems arise. The major disadvantage in using A(B-V) is that there are uncertainties in
the theoretical colors used to convert the observed values to ages. These Wlcertainties are the result
of uncertainties in the model stellar radii and the conversion of effective temperatures to colors.
Another shortcoming of using A(B-V) is its sensitivity to the cluster [OIFe] and the degree of
helium diffusion. Ata given age, an increase in [OIFe] makes the color of the tumoffredder
(Simoda and lben 1968; Renzini 1977) so that A(B-V) becomes smaller. Therefore, an increase in
age causes A(B-V) to decrease and so too does an increase in [OIFe]. Similarly, inclusion of
helium diffusion in the models of Population II stars makes the color of the turnoff redder. Thus,
A(B-V) is reduced by an increased age, an enhancement in oxygen abundance, and through the
effects of helium diffusion.
No studies have yet been published with absolute ages derived using the A(B-V) method.
VandenBerg et al. (1990), Sarajedini and Demarque (1990), and Sarajedini (1991) have fOWld that
there is an age range of at least 2.5 Gyr among the Galactic globular clusters (see also §3 for an
even larger spread). Furthenoore, one of us (A. S.) has compiled A(B-V) values for 21 globular
clusters in the Galaxy and used the Revised Yale Isochrones (Green et al. 1987, RYI), to derive
ages for this sample of clusters. As yielded by the RYI, the age of the oldest clusters is =18 Gyr;
the RYI assume [OIFe] = 0 and do not include the effects of helium diffusion. If [OIFe] is
enhanced at a constant level regardless of [Fe/H] and a reasonable amount of diffusion is added,
this age would be reduced by about 3 to 4 Gyr.

3. An Age Spread Among Globular Clusters

The significance of globular clusters to cosmology goes beyond providing a lower limit to the age
of the Universe. The chronology of the globular cluster system is also a key input in our
understanding of the fonoation and the evolution of galaxies. Both of these topics are important
parts of cosmology. The problem of a possible age spread among Galactic globular clusters has
often been considered within the context of a possible age-meta1licity relation among globular
clusters (see Demarque 1980; Zion 1984; Sarajedini and King 1989 for a recent discussion). This
question is related to the more difficult problem of absolute ages, and is still open.
From a different point of view, Searle and Zion (1978) have made the hypothesis that age
is the second parameter that controls the morphology of globular cluster horizontal branches
(meta1licity being the first parameter), primarily because it is related to the global properties of the
Galactic halo (see also Rood and lben 1968). Recently, thanks to precision CCD photometry
within globular clusters, and to advances in horizontal branch modeling, it has become possible to
contino the Searle-Zion suggestion that age is likely to be the second parameter (Lee et al. 1990;
LDZ). Although other second parameter candidates may playa role in explaining some features of
HB morphology, age provides the only explanation of the second parameter that is compatible with
both standard stellar evolution theory and the observed distribution of RR Lyrae variables.
The strongest test is provided by pairs of clusters with very nearly the same metallicity, but
with unmistakingly different HB types. One well-studied pair is NGC 288-NGC 362 (Bolte 1989),
with [Fe/H] = - -1.4, shown in Figure 2a, for which precise relative turnoff photometry was
obtained by Green and Norris (1990). The age difference between the clusters is near 3 Gyr, as
1\8

determined from both isochrone fitting and the ~(B-V) method (Sarajedini and Demarque 1990).
Another striking pair is provided by NGC 6397 and the newly swdied cluster Ruprecht 106
(Buonanno et al. 1990), with [FelHJ = - -1.9, shown in Figure 2b. In this case, the age difference
is about 4 Gyr. These two age differences do not necessarily overlap; therefore, the 4 Gyr
difference seems to represent a minimum time scale for the formation of the halo. The interpretation
of age as the second parameter thus rules out the Eggen et al. (1962) picwre of a rapid formation of
the Galactic halo as a single event Rather, it lends support to the idea of a prolonged phase of halo
formation, possibly involving mergers and the accretion of large fragments with independent
dynamical and nucleosynthetic histories (see recent review by Larson 1990).

4. Recent Developments Absolute Ages


4.1 ROTATION

Early stellar models with rotation suggested that rotation could lengthen the main sequence lifetime
relative to standard models through its effects on hydrostatic equilibrium, thus yielding larger ages
for globular clusters (Mengel and Gross 1976). However, angular momentum loss from the
surface and redistribution in the interior was ignored. Young stars are observed to spin down, and
the old halo stars are known to rotate very slowly.
Following the approach pioneered by Endal and Sofia (1976, 1981), it has recently been
possible to include physical descriptions in stellar evolutionary models of the effects of: a) surface
angular momentum loss due to a magnetic wind, which is calibrated to match the observed spin-
down of stars, b) the internal redistribution of angular momenmm triggered by a variety of
rotationally-induced instabilities, and c) associated material mixing. Models of globular cluster stars
lose enough angular momenmm so that rotation provides onI y a minor perturbation on their internal
strucwre, and therefore ages «1 % effect, Deliyannis et al. 1989). Rotation does not affect globular
cluster ages as estimatedfrom either isochronefitting, the LiV method, or the Li(B-V) method. The
models do, nevertheless, retain enough angular momentum in the interior in the form of a rapidly
rotating core, to account for the high surface velocities observed in horizontal branch stars.
Agreement between the models and the observations strongly suggests that, contrary to what has
often been suggested, stars do not rotate rigidly as they evolve (Pinsonneault, et al. 1990).

4.2 DIFFUSION

In halo stars, helium can diffuse inward relative to hydrogen (e.g. gravitational settling, thermal
diffusion). This diffusion can reduce the age scale for globular clusters in at least two ways. First,
lower relative central hydrogen abundance can reduce the main sequence lifetime of the models
(Stringfellow et al. 1983). Second, a higher relative hydrogen abundance in the envelope results in
a larger main sequence radius (lower Teff), while post main sequence dredge-up of most of the
diffused helium results in only a small change of the giant branch position; therefore, the size of the
subgiant portion of an isochrone is reduced, and the main sequence turnoff is redder. Thus, a
lower-aged isochrone is required to fit a given cluster (Deliyannis et al. 1990). With helium
diffusion included, all three dating methods (isochrone fitting, ~V, and ~(B- V) ) yield lower ages.
The two papers mentioned previously in this section (§4.2) and also that of Proffitt et al. (1990)
suggested that the age reduction might be quite substantial, perllaps 20 - 30%.
HB (B-V)o HB (B-V)o
0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8

o 0
o 0 0

1 f- • • •• , -I 1
• •
• •
2f- 2

3
10, 12, 14, 16, 18 Gyr 10, 12, 14, 16, 18 Gyr Mv
~] 4
[Fe/H] =- 1.3 [Fe/H] = - 1.9
5f- Mge = 4Gyr ~5
' - _ Mge=3Gyr
• NGC288 • NGC6397
o NGC362 o Ruprecht 106
61- -16
(a) (b)

0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.3 0.4 0.5 0.6 0.7 0.8 0.9
(B-V)o (B-V)o

Figure 2. A significant age spread in the Galactic halo. Shown are two pairs of clusters of the same metallcity for two different
metallicities. The age differences of 3 and 4 Gyr, derived using the isochrones of Green et al. (1987), need not overlap; therefore the halo
took at least 4 Gyr to form. Note the difference in scale for the horizontal branches, which are shown only to illustrate the morphology.. The
fiducial sequence shown for NGC 362 is taken from VandenBerg et aI. 1990. Those for NGC 288, Ruprecht 106, and NGC 6397 were
derived from the unpublished data of Bolte, of Buonanno, and of Stetson, respectively (1990,1991,1991, private communications), who
have generously provided their data to us before publication. \0
120

Unfortunately it is not possible to test the helium diffusion predictions directly because it is
very difficult to observe helium in low mass stars. To complicate matters further, it is possible that
rotationally-induced motions render diffusive motions inefficient; thus, the quoted age reduction
should be treated as a maximal possible reduction.

4.3 L1TIllUM COMES TO THE RESCUE

Fortunately, indirect constraints may be put on the possible effects of helium diffusion by using the
observations of lithium in extreme halo stars (Deliyannis and Demarque 1991 b). Since lithium
diffuses at a similar rate as helium, lithium observations yield infonnation about helium diffusion.
What is required is a sample of halo dwarf and turnoff stars with a range of masses (Teff) and with
a uniform initial abundance. It just so happens that these requirements are satisfied by extreme field
halo stars, whose initial lithium abundance very likely reflects that synthesized during the big bang
itself (Deliyannis 1990). Furthennore, lithium observations in these stars exhibit a remarkably
striking and well-defined pattern in the Li-Teffplane (Spite and Spite 1982; compilations in
Deliyannis et aI. 1990 and Deliyannis and Demarque 1991a): a unifonn (or nearly unifonn) plateau
from 6400 K to 5500 K (possibly with a slight increasing slope toward higher Teff), and
progressive depletion relative to the plateau for stars progressively cooler than 5500 K, which is
nearly certainly due to destruction of 7Li through the (p,a) reaction. Figure 3a shows how a
standard Li isochrone can reproduce these features in detail. Acceptable diffusive Li isochrones in
the Li-Teffplane must pass this Liflatness contraint.
Diffusion depletes the surface lithium abundance by draining lithium out of the bottom of
the convection zone. A shallower convection zone is more conducive to diffusion; therefore, in a
model sequence, diffusion rates increase with time as the surface convection zone recedes during
main sequence evolution. Furthennore, higher mass models have shallower convection zones;
therefore, a diffusive isochrone (in the Li-Teffplane) exhibits increasing downward curvature
toward higher Teff (Figure 3b). The degree of this curvature can be constrained by the lithium
observations.
Isochrones from the relevant works mentioned above do not pass the Li flatness constraint
(at >4u level); therefore, the effects oj helium diffusion have been overestimated. This may be due
in part to uncertainties in the detennination of the depth of the surface convection zone. For
example, reported opacity enhancements (Iglesias et aI. 1990) relative to previous work could result
in a deeper model convection zone. Another possibility, hinted at from diffusive solar-calibrated
models, is that smaller envelope Y requires a larger mixing length. In halo star models, this would
result in a deeper convection zone; diffusion might then require an increasing mixing length with
time, and thus a still deeper convection zone. Regardless, passing the Li flatness constraint implies
a smaller diffusion-induced reduction in globular cluster age estimates. Recent models of Chaboyer
et al. (1991) that use Los Alamos opacities (Huebner 1978) with the Grevesse (1984) mixture,
[O/Fe] = 0.4, and higher than solar-calibrated mixing length have been able to pass the Li flatness
constraint (though not the depleted cool star constraint), and result in -2 Gyr age reduction for
M92, one of the oldest clusters. More theoretical work is required to gauge the importance of
helium diffusion.
Note that while improved opacities yield improved isochrone fits, age estimates remain
unaffected (Demarque et al. 1988). Note also that high precision luminosity functions near the
turnoff might be able to constrain the diffusion theory.
121

6000 5500 5000

0.0
2.0 (a)

,-... 0.5
.f
--
1.5

:; t:
--
1.0

f
'-'
[] Detections with -2.7 ~ [Fe/H]
~ 1.0
• Detections with -1.7 ~ [Fe/Hl > -2.7
+ o Detections with -1.3 ~ [FeIHl > -1.7 v
v Upper Limits 1.5
0.5 Same Star Multiple Observations
- - Z = 10-4 Standard Li Isochrone
2.0
0.0 standard isochrone (Deliyannis et al. 1990) v

0.0

2.0 (b)
0.5

.f
--
1.5
~
:;
1.0
typical uncertainty
t:
'-'
~ 1.0
--t:
1.5 ~:

-
+ '-'
N - - DeJiyannis and Demarque 1991b
- - - Proffitt et al. 1990
0.5 - - Chaboyeretal.I991
.'V 2.0

0.0 diffusive isochrones v


2.5

6000 5500 5000

Teee
Figure 3. The Li flatness constraint. A standard Li isochrone (right scale) reproduces in detail
the observed plateau (left scale), and the decreasing abundance with decreasing Teff (a). Diffusion,
on the other hand, results in a downward curvature toward higher T eff (b). The effects of diffusion
(e.g. age reduction) can thus be observationally constrained by the Li abundances. Observations
are from the works of Spite, Hobbs, Rebolo, and collaborators, and are compiled in
Deliyannis et al. 1990 and in Deliyannis and Demarque 1991a.
122

4.4 A MINIMUM AGE FOR ONE OF THE OLDEST CLUSTERS

Using the physics assumed in Chaboyer et al. (1991) and completely standard models, the best age
derived for one of the oldest clusters, M92, is -17 Gyr (Figure 4). Using [OIFe] = 0.4 brings this
down to -16 Gyr, and assuming unimpeded helium diffusion brings it further down to -14 Gyr.
In spite of the many uncertainties discussed, it is difficult to reduce this age further. For example,
no further age reduction is found when assuming even higher [OIFe] because it would then be
necessary to reduce the amount of He diffusion in order to pass the Li flatness constraint. We must
conclude, then, that 14 Gyr is the minimum age of the Galactic halo.

5. Galaxy Chronology and Cosmological Implications

Our educated guess for Galactic chronology is shown in Figure 5. The best detennined age is that
of the Sun, at 4.49 Gyr (Guenther 1989). The age of the oldest disk open cluster for which we
have a good color-magnitude diagram and a reliable chemical composition is NGC 188. Its age has
been estimated as 6.5 ~.; Gyr (Demarque et al. 1990). The age of the oldest disk stars can be
estimated from the white dwarf cooling times (e.g.lben and Laughlin 1989). The ages of most
disk globular clusters are unknown at this point. Most of these relatively metal-rich objects (with
[FelH] - -0.7 or higher) are situated in crowded fields of the Galactic bulge, and as such will be
major targets for HST observations. The best studied member of the group at this point is 47 Tuc
(Hesser et al. 1987), for which age estimates fall in the range 11-14 Gyr. It is possible that still
older metal-rich clusters exist in the Galactic bulge. Finally, we show the metal-poor cluster M92.
As a representative of the oldest known stellar populations in the Galactic halo, it is of particular
cosmological interest since it sets a lower bound to the age of the Universe. The age range
indicated is 14-19 Gyr. We emphasize here that these age ranges do not represent standard errors in
the statistical sense. Rather they reflect the limits of currently accepted uncertainties in input
parameters of the stellar models. For M92 to be much younger than 14 Gyr would require an
extreme set of circumstances including both an [OIFe] ratio higher than is currently believed likely,
and more effective helium diffusion than seems presently allowed by the lithium observations in
field dwarfs of the same metallicity. We note also that a relatively long phase of halo fonnation and
development is likely to have taken place, possibly extending to at least 5 Gyr, in contrast with the
rapid collapse (0.1 Gyr) proposed by Eggen et al. (1962, §3). This factor puts further constraints
on attempts to squeeze down the scale of Galactic chronology. We also emphasize that the age
ranges presented here have been generously stretched, particularly at the low age end. Note that
ages of other authors tend to be systematically higher than the ages of our best models.
The implications for cosmology are summarized by the two lower scales on Figure 5. The
bottom scale shows the inverse of the Hubble parameter Ho which sets the universal expansion
timescale. Just above it, we also show the quantity 2/3Ho, which is the relevant timescale if the
Universe is flat. In the case of a flat Universe with no cosmological constant ( 0= I, A=O ), for
values of Ho in the vicinity of 90 km s-1 Mpc-l, there is a significant timescale incompatibility. The
problem appears to go beyond the application of a simple scale factor to the Galactic age scale since
it conflicts with a large and varied body of astronomical knowledge on stellar and Galactic
evolution. At this point, the problem seems unlikely to disappear: more than two decades of
progress on both the theoretical and observational sides, by a host of authors, have sharpened this
result. On the other hand, low values of Ho, around 45 km s-1 Mpc-l or less are compatible with
123

0.4 0.5 0.6 0.7 0.8

Standard
(m-M)y = 14.75 [O/Fe] = 0.0
17 E(B-V) = 0.03 14, 16, 18, 20 Gyr 17

18 18

V 19 19

20 " 20

.,
21 21

(m-M)y = 14.80 [O/Fe] = 0.4


17 E(B-V) =0.02 13, 15, 17, 19 Gyr 17

18 18

V 19 19

20 20

21 21

0.4 0.5 0.6 0.7 0.8


B-V
Figure 4. Minimal age for one of the oldest clusters, M92. Standard isochrones with
[O/Fe] = 0.0 yield an age of - 17 Gyr; including both oxygen enhancement and helium diffusion
yields - 14 Gyr (but with a fit oflesser qUality). Note that the degree of oxygen enhancement, if
any, is controversial (Pilachowski 1988). The data are from Stetson and Harris (1988).
124

the stellar nuclear ages. In addition, for values of 0 closer to those derived for O(baryon)
( - om - 0.2), an Ho as high as 70 Ian s-1 Mpc-l can be accomodated.
Along with the constraints set by big bang nucleosynthesis, the age of the oldest globular
clusters thus seems to provide one of the best established constraints to cosmology.

GALACTIC CHRONOLOGY

Local Galactic Disk:


White Dwarfs, Oldest Disk Stars Estimated Minimal Age Spread for Galactic Halo
". c
I. ~

NGC. 188'1 47 ,Tuc ~'''''' M92 ....... 1


0 I.. ·L·-II-·· ..:'·· ..I ,····::···.. ··t
• • • • • •
Gyr 0 5 10 15 20 25
2/3Ho 133 67 44 33 27
Ho (kms·1Mpc·l) 200 100 67 50 40

Figure 5. Besides the age of the Sun, shown are probable ranges for the old Galactic disk cluster
NGC 188, the disk globular 47 Tuc, and one of the oldest globular clusters M92. Also shown is a
range for the oldest stars in the solar neighborhood (with stars in the central portion of the disk and
in the bulge possibly being much older), and a likely minimum time offonnation for the Galactic
halo (§3).

Acknowledgements :
We would like to thank M. Bolte, R Buonanno, and P. Stetson for graciously allowing us
to use their new unpublished data on NGC 288, Ruprecht 106, and NGC 6397, respectively, to
derive the fiducial sequences presented here.
This work was supported in part by NASA grants NAGW-777 and NAGW-778.

6. References

Abia, C. and Rebolo, R (1989) 'Oxygen Abundances in Unevolved Metal-poor Stars :


Interpretation and Consequences', Ap. J., 347,186-194.
Boesgaard,A. and Steigman, G. (1985) 'Big Bang Nucleosynthesis: Theories and Observations',
Ann. Rev. Astr. Ap., 23,319-378.
Bolte, M. (1989) 'The Age ofNGC288, the Fonnation of the Galactic Halo, and the Second
Parameter", A. J., 97, 1688-1698.
Bond, H. E. and Luck, R E. (1988) 'Chemical Compositions of Population II Mid- and Late-type
Stars', in G. Cayrel de Strobel and M. Spite (eds.), The Impact of Very High Signal to Noise
Spectroscopy on Stellar Physics, IAU Symp.132, Reidel, Dordrecht, pp. 477-483.
Buonanno, R, Corsi, C. E., and Fusi Pecci, F. (1989) 'The Ages of Globular Ousters and the
Sandage Period Shift Effect', Astr. Ap., 216, 80-108.
Buonanno, R, Buscema, G., Fusi Pecci, F., Richer, H. B. and Fahlman, G. G. (1990) "Ruprecht
106: A Young Metal-Poor Galactic Globular auster", A. J., 100, 1811-1840.
125

Buzzoni, A., Fusi Pecci, F., Buonanno, R., and Corsi, C. E. (1983) 'Helium Abundance in
Globular Clusters', Astr. Ap., 128, 94-101.
Cacciari, C., Clementini, G., and Buser, R. (1989) 'The Baade-Wesselink Method Applied to
Field RR Lyrae Stars. TIl. YZ Capricorni, RV Phoenicis, V440 Sagittarii', Astr. Ap., 209,
154-164.
Chaboyer, B., Deliyannis, C. P., Demarque, P., Pinsonneault, M. H., and Sarajedini, A., (1991)
"Effects of Helium Diffusion on Globular Custer Ages Inferred from Isochrones", presented at
the 177th AAS meeting in Philadelphia Abstract in Bulletin MS, 22, p. 1205.
Chandrasekhar, S. (1939) An Introduction to the Study 0/ Stellar Structure, University of Chicago
Press, Chicago.
Ciardullo, R. B. and Demarque, P.(1977) ''Tables ofIsochrones", Yale Trans., Vol.33.
Davidson, K. and Kinman, T. D., (1985) 'Primordial Helium, Spectrophotometric Teclmique, and
I Zwicky 18', Ap. J. Supp., 58, 321-340.
Deliyannis, C. P. (1990), Primordial Lithium and Stellar Evolution: Lithium in Halo Stars/rom
Standard, Diffusive, and Rotational Stellar Evolution and Implications/or Cosmology, Ph. D.
Dissertation, Yale University, pp. 1-323.
Deliyannis, C. P. and Demarque, P. (1991a) 'Lithium in the Most Extreme Halo Stars; Trends
With Metallicity', Ap. J.(Letters) ,370, (in press, April).
Deliyannis, C. P. and Demarque, P. (1991b) 'Lithium in Halo Stars: Constraining the Effects of
Helium Diffusion on Globular Cluster Ages and Cosmology', Ap. J. , submitted.
Deliyannis, C. P., Demarque, P., Kawaler, S. D. (1990) 'Lithium in Halo Stars from Standard
Stellar Evolution', Ap. J. Suppl., 73, 21-65.
Deliyannis, C. P., Demarque, P., and Pinsonneault, M. H. (1989) 'The Ages of Globular Cluster
Stars: Effects of Rotation on Pre-Main Sequence, Main Sequence, and TumoffEvolution',
Ap. J. (Letters), 347, L73-L76.
Demarque, P.(1964) "Additional Meeting of the R. A. S., Dec.1964", Observatory, 84, 249.
Demarque, P.(1966) "Evolution of Stars Near One Solar Mass", in R. F. Stein and A.G. W.
Cameron (eds), Stellar Evolution, Plenum Press, New York, pp. 231-236.
Demarque, P. (1967) "The Early Evolution of Population II Stars", Ap. J., 149, 117-129.
Demarque, P.(1980) "Ages and Abundances of Globular Clusters and the Oldest Open Clusters",
in 1.E.Hesser (ed), Star Clusters, IAU Symp.85, pp.281-298.
Demarque, P. and Geisler, 1. E. (1963) "Models for Red Giant Stars", Ap. J., 137, 1102-1120.
Demarque, P., Guenther, D. B., King, C. R., and Green, E. M. (1988) 'Isochrone Comparisons,
Stellar Physics, and Implications for Stellar Ages', in Calibration o/Stellar Ages, ed. A. G.
Davis Philip (L. Davis Press), Schenectady, p. 101-115.
Demarque, P., Guenther, D. B., and Green, E. M. (1990) "Age Estimates for the Old Open
Clusters M67, NGC188 and NGC6791", in The Formation and Evolution o/Star Clusters, A.
S. P. Conference Series, K. A. Janes (ed), in press.
Demarque, P. and McClure, R. D. (1977) "Stellar Populations in the Disk and Halo of the
Galaxy", in B. M. Tinsley and R. B. Larson (eds),The Evolution 0/ Stellar PopUlations, Yale
Univ. Obs., New Haven, pp.199-215.
Eggen, O. J., Lynden-Bell, D. and Sandage, A. (1962) "Evidence from the Motions of Old Stars
that the Galaxy Collapsed", Ap. J., 136, 748-766.
Endal, A. S. and Sofia, S. (1976) 'The Evolution of Rotating Stars. I. Method and Exploratory
Calculations for a 7 M0 Star', Ap. J., 210,184-198.
Endal, A. S. and Sofia, S. (1981) 'Rotation in Solar-Type Stars. I. Evolutionary Models for the
Spin-Down of the Sun', Ap. J., 243,625-640.
Fusi Pecci, F., Ferraro, F. R., Crocker, D. A., Rood, R. T., and Buonanno, R. (1990) 'The
Variation of the Red Giant Luminosity Function "Bump" with Metallicity and the Age of the
Globular Clusters', Astr. Ap., 238, 95-110.
Gratton, R. G. (1985) 'Deep Photometry of Globular Clusters. V. Age Derivations and Their
Implications for Galactic Evolution' , Astr. Ap., 147, 169-177.
Gratton, R. G. (1990), in G. Michaud and A. Tutukov (eds.), Evolution 0/ Stars: The
Photospheric Connection, IAU Symp. 145, in press.
Gratton, R. G. and Ortolani, S. (1989) 'Metal Abundances in Metal-poor Globular Clusters', Astr.
126

Ap., 211, 41-55.


Green, E. M., Demarque, P., and King, C. R. (1987), The Revised Yale Isochrones and
Luminosity Functions (New Haven: Yale University Observatory).
Green, E. M. and Norris, 1. E. (1990) "Population Studies: The Age Difference Between the
Globular Ousters NGC 288 and NGC362", Ap. J. (Letters), 353, L17-L20.
Grevesse, N. (1984) "Accurate Atomic Data and Solar Photospheric Spectroscopy", Phys. Scripta,
T8,49-58.
Guenther, R. B. (1989) "Age of the Sun", Ap. J., 339, 1156-1159.
Haselgrove, C. B. and Hoyle, F. (1956a) "A Mathematical Discussion of the Problem of Stellar
Evolution, with Reference to the Use of an Automatic Digital Computer", M. N. R. A. S.,
116, 515-526.
Haselgrove, C. B. and Hoyle, F. (1956b) "A Preliminary Determination of the Age of Type II
Stars", M. N. R. A. S., 116,527-532.
Henyey, L. G., Wilets, L., Bohm, K., LeLevier, R., and Levee, R. D. (1959) "A Method for
Automatic Computation of Stellar Evolution", Ap. J., 129, 628-636.
Hesser, 1. E., Harris,W. E., VandenBerg, D. A., Allwright, J. W. B., Shott, P. and Stetson, P.
B. (1987) "A CCD Color-Magnitude Study of 47 Tucanae", Pub. A. S. P., 99, 739-808.
Hoyle, F. (1959) "The Ages of Type I and Type II Subgiants", M. N. R. A. S., 119, 124-133.
Hoyle, F. and Schwarzschild, M. (1955) "On the Evolution of Type II Stars", Ap. J. Supp., 2,1-
40.
Huebner, W. F. (1978) "Solar Opacities and Equation of State", in Proc. o/the Brookhaven Solar
Neutrino Conference, ed. G. Friedlander (BNL 50879), Vol. 1, p. 107.
Iben, I. Jr. (1967) "Stellar Evolution Within and Off the Main Sequence", Ann. Rev. Astr. Ap., 5,
571-626.
Iben, I. Jr. (1974) "Post Main Sequence Evolution of Single Stars", Ann. Rev. Astr. Ap., 12,
215-256.
Iben, I. Jr. and Laughlin, G. (1989) "A Study of the White Dwarf Luminosity Function", Ap. J.,
341, 312-326.
Iglesias, C. A., Rogers, F. 1., and Wilson, B. G. (1990) 'Opacities for Clasical Cepheid Models',
Ap. J. ,360, 221-226.
Janes, K. A. and Demarque, P.(1983) "The Ages and Compositions of Old Ousters", Ap. J.,
264,206-214.
King, C. R., Demarque, P., and Green, E. M. (1988) 'Isochrone Fitting of Color Magnitude
Diagrams of Old Star Ousters: Implications for Stellar Ages', in A. G. D. Philip (ed.),
Calibration o/Stellar Ages, L. Davis, Schenectady, pp. 211-216.
Kippenhahn, R., Temesvary, S. and Biermann, L. (1958) "Die Entwicklung der Sterne der
Population II", ZeitAstrophys., 46, 257-275.
Larson, R. B. (1990) "Galaxy Building", Pub. A. S. P., 102, 709-722.
Larson, R. B. and Demarque, P. (1964) "An Application of Henyey's Approach to the Integration
of the Equations of Stellar Structure", Ap. J., 140,524-540.
Lee, Y. -W. (1990) "On the Sandage Period Shift Effect among Field RR Lyrae Stars", Ap. J.,
363, 159-167.
Lee, Y. -W., Demarque, P., and Zinn, R.J. (1990) 'The Horizontal Branch Stars in Globular
Ousters. I. The Period-Shift Effect, the Luminosity of the Horizontal Branch, and the Age-
Metallicity Relation', Ap. J., 350, 155-172.
Liu, T. and Janes, K. A. (1990) 'The Luminosity Scale ofRR Lyrae Stars With the Baade-
Wesselink Method. II. The Absolute Magnitudes of 13 Field RR Lyrae Stars', Ap. J., 354,
273-294.
Longmore, A. J., Dixon, R., Skillen, I., Jameson, R. F., and Fernley, J. A. (1990) 'Globular
Ouster Distances From the RR Lyrae Log (period) - Infrared Magnitude Relation', M. N. R.
A. S., 247, 684-695.
McClure, R. D., Vandenberg, D. A., Bell, R. A., Hesser, J. E., and Stetson, P. B. (1987) 'CCD
Photometry of the Globular Cluster M68', A. J., 93, 1144-1165.
Mengel, 1. G. and Gross, P. G. (1976) 'Possible Effects of Internal Rotation on Low-Mass Stars',
Astrophys.space Sci., 41,407-415.
127

Mengel, 1. G., Sweigart, A. V., Gross, P. G. and Demarque, P. (1979) "Stellar Evolution from
.. the Zero-Age Main Sequence", Ap. 1. Supp., 40, 733-791.
Qpik, E.(1938a) "Composite Stellar Models", Tartu Obs. Pub. 30, no.3, pp.48.
Opik, E.(1938b) "Stellar structure, Source of Energy, and Evolution", Tartu Obs. Pub. 30, no.4,
pp.118.
Peterson, C. J. (1987) 'Ages of Globular Clusters', Pub. A. S. P., 99,1153-1160.
Pilachowski, C. A. (1988), "The Abundance of Oxygen in M92 Giant Stars", Ap. 1. (Letters),
326, L57-L60.
Pilachowski, C. A., Sneden, C., and Wallerstein, G., (1983) 'The Chemical Composition of Stars
in Globular Clusters', Ap. 1. Supp., 52,241-287.
Pinsonneault, M. H., Deliyannis, C. P., and Demarque, P. (1991) 'Evolutionary Models of Halo
Stars with Rotation: I. Evidence for Differential Rotation wilh Deplh in Stars', Ap. 1.,367,
239-252.
Proffitt, C. R., Michaud, G., and Richer, 1. (1990) 'Diffusion in Metal Poor Low Mass Stars', in
Cool Stars, Stellar Systems, and the Sun, ASPCS Vol. 9, ed. G. Wallerstein, (Provo:
Brigham Young), pp. 351-353.
Renzini, A. (1977) 'The Evolution of Population II Stars and Mass Loss and Stellar Evolution', in
P. Bouvier and A. Maeder (eds.), Advanced Stages of Stellar Evolution, 7th Course of the
Swiss Society of Astronomy and Astrophysics, Saas-Fee, Geneva Observatory, pp. 149-283.
Rood, R. and Iben, I. Jr. (1968) "On the Variation of Globular-Cluster Characteristics with Age",
Ap. I., 154, 215-224.
Sandage, A. (1953) 'The Color-Magnitude Diagram for the Globular Cluster M3", A. 1.,58,61-
75.
Sandage, A. (1957) "Observational Approach to Evolution.II.", Ap. 1.,125,435-444.
Sandage, A. (1962) 'The Ages of M67,NGCI88,M3,M5,and M13 According to Hoyle's 1959
Models", Ap. I., 135, 349-365.
Sandage, A. (1986) "The Population Concept, Globular Clusters, Subdwarfs, Ages, and the
Collapse of the Galaxy", Ann. Rev. Astr. Ap., 24,421-458.
Sandage, A. (1990) 'The Oosterhoff Period Effect: Luminosities of Globular Cluster Zero-Age
Horizontal Branches and Field RRLyrae Stars as a Function of Metallicity' , Ap. I., 350, 631-
644.
Sandage, A. and Cacciari, C. (1990) 'The Absolute Magnitudes ofRR Lyrae Stars and the Age of
the Galactic Globular Cluster System', Ap. I., 350, 645-661.
Sandage, A. and Schwanschild, M. (1952) "Inhomogeneous Stellar Models.!I.", Ap. 1.,116,
463-476.
SchOnberg, M. and Chandrasekbar, S. (1942) "On the Evolution of the Main-Sequence Stars", Ap.
1.,96, 161-172.
Sarajedini, A. and King, C. R. (1989) 'Evidence for an Age Spread Among the Galactic Globular
Clusters', A. 1.,98, 1624-1633.
Sarajedini, A. and Demarque, P. (1990) 'A New Age Diagnostic Applied to the Globular Clusters
NGC 288 and NGC 362', Ap. I., 365,219-223.
Sarajedini, A. and Lederman, A. (1991) 'The Horizontal Branch Luminosity - Metallicity Relation:
Preliminary Work', in K. Janes (ed.),The Formation and Evolution of Star Clusters,
Proceedings of the 102nd Meeting of the Astronomical Society of the Pacific, in press.
Sarajedini, A. (1991) 'Globular Cluster Photometry Near the Tum-off: Relative Ages and the
Horizontal Branch', in A. G. D. Philip (ed.), Precision Photometry: Astrophysics of the
Galaxy, L. Davis, Schenectady, in press.
Searle, L. and Zinn,.R. (1978) "Compositions of Halo Clusters and the Formation of the Galactic
Halo", Ap. 1.,.225, 357-379.
Simoda, M. and Iben, I. (1968) 'Heavy Element Abundances and lhe Interpretation of Globular
Cluster Characteristics', Ap. I., 152, 509-513.
Spite, F. and Spite,. M. (1982) 'Abundance of Lithium in Unevolved Halo Stars and Old Disk
Stars: Interpretation and Consequences', Astr. Ap., 115, 357-366.
Stetson, P. B. and Harris, W. E. (1988) 'CCD Photometry of the Globular Cluster M92', A. I.,
96, 909-975.
128

Stetson, P. B., VandenBerg, D. A., Bolte, M., Hesser, J. E., and Smith, G. H. (1989) 'CCD
Photometry of the Anomalous Globular Cluster Palomar 12', A. J., 97, 1360-1396.
Straniero, O. and Chieffi, A. (1991) 'Isochrones for H-Burning Globular Cluster Stars. II. The
Metallicity Range -2.3 < [Fe/Hl < -0.5', Ap. J. Supp., in press.
Stringfellow, G. S., Bodenheimer, P. Noerdlinger, P. D., and Arigo, R J. (1983) 'Evolutionary
Effects of Helium Diffusion in Population II Stars' ,Ap. J., 264, 228-236.
Tayler, R J. (1964) "Additional Meeting of the RA.S.,Dec.1964", Observatory, 84, 249.
VandenBerg, D. A. (1985) "Evolution of 0.7-3.0 Me Stars Having -1.0 < [Fe/Hl < 0.0", Ap.
J. Supp., 58, 711-769.
VandenBerg, D. A. and Bell, R A. (1985) 'Theoretical Isochrones for Globular Clusters with
Predicted BVRI and Str6mgrenPhotometry',Ap.J. Supp., 58, 561-621.
VandenBerg, D. A., Bolte, M., and Stetson, P. B. (1990) 'Measuring Age Differences Among
Globular Clusters Having Similar Metallicities: A New Method and First Results', A. J., 100,
445-468.
Zinn, R J. (1986) 'Globular Clusters in Local Group Galaxies', in C. A. Nonnan, A. Renzini,
and M. Tosi (eds.), Stellar Populations, Space Telescope Science Institute Symposium Series,
Cambridge University, Cambridge), pp. 73-100.
Zinn, R 1. and West, M. J., (1984) 'The Globular Cluster System of the Galaxy. III.
Measurements of Radial Velocity and Metallicity for 60 Clusters and a Compilation of
Metallicities for 121 Clusters', Ap. J. Supp., 55, 45-66.
Zion, R (1988) "An Overview of the Globular Cluster System of the Galaxy", in Globular Cluster
Systems in Galaxies, IAU Symp.126, J. E. Grindlay and A. G. D. Philip (eds), pp.37-48.
129

DISCUSSION:

Hoyle: This is very beautiful work and Prof. Demarque is to be congratulated on the
persistence and skill needed to obtain such results. You mentioned mass loss briefly. Isn't
this of main concern on the giant branch where it hardly affects cluster ages?
Demarque: Mass loss is not a problem in age determination, because little mass loss
takes place near the main-sequence turnoff; observations of surface Li and Be put strong
constraints on main-sequence mass loss. Near the tip of the giant branch, substantial mass
loss must take place, on a relatively short time-scale. The precise mechanism is not well
understood.

Borner: What is the maximum effect of the assumption on the depth of the convection
layer?

Demarque: The depth of the convection zone affects the amount of He and Li depletion
due to gravitational settling.
Lynden-Bell: What do you think the Galaxy looked like when M92 was there and the
disk was not? And how long did that condition last?
Demarque: Probably quite chaotic. The condition may have lasted several giga-years.
GLOBULAR CLUSTER AGES AND COSMOLOGY

ALVIO RENZINI·
A"tronomy Department,
Univer"ity of Bologna,
CP 596 1-.. 0100 Bologna, Italy

1. Towards Getting Precision Ages for Galactic Globulars

The age of a A = 0, Friedmann universe is given in Table 1, for various combinations


of the Hubble constant Ho and the density parameter no. In turn, it is usual to
assume the present age of the universe as given by its age at the epoch of galaxy
formation, plus the age of globular clusters, i.e.:

to = t(ZGF) + tGC, (1)

where ZGF is the redshift of the epoch of galaxy formation. Preasumably t(ZGF) ~
tGC [say, t(ZGF) ~ 1- 2 Gyrj, so to ~ tGc, Unlike for the case of Ho (for which well
known discrepancies exist in current estimates), virtually every practicioner in the
field of cluster dating would agree that

tGC = (13 - 15) ± 3 Gyr (2)

represents a fair estimate of the age of the oldest clusters in the Galaxy. The
agreement follows from everybody using basically the same input physics in stellar
model calculations and the same observational data, and as such it does not really
ensure that GC ages have been correctly estimated. In any event, this determination
is marginally consistent with the standard inflationary scenario (no = 1, A = 0) if
Ho = 50, while it would clearly require A f:. 0 if Ho = 100. Although the possibility
of living in a A f:. 0 universe is now more widely entertained than a few years ago
(see Weinberg 1989, for a recent extensive discussion), as other lines of evidence may
now suggest A f:. 0, still the obvious questions to answer are "How good are GC age
determination" ~", and also "Do GC ages set tight limit" on the other cosmological
parameters f"

131
T. Shanks et al. (eds.). Observational Tests o/Cosmological Inflation. 131-146.
© 1991 Kluwer Academic Publishers.
132

Table 1

The Ages of A = 0 Universes


Ho no to (Gyr)
100 0 10
100 1 6.7
50 0 20
50 1 13

1.1 THE CLOCK AND ITS UNCERTAINTIES


The output of stellar model calculations are evolutionary sequences, each referring
to a star of given mass and composition. By joining points of the same age lying on
the evolutionary lines of models of different mass one obtains the famous isochrones,
i.e. the loci occupied (usually on the HR, or color-magnitude diagrams) by stars
of the same age and composition, but different mass. Thus isochrones can be used
to infer the age of a cluster by comparing them to the cluster locus, i.e. to the
distribution of cluster stars on the C-M diagram. In practice, the single detail that
can be used to infer the cluster age is the luminosity of the main sequence turnoff
(see insert at the end of this section for a critique of another popular approach),
and e.g. Buonanno, Corsi, and FUsi Pecci (1989) derive from the isochrones of
VandenBerg and Bell (1985):

Log to ~ -0.41 + 0.37 MJo - 0.43 Y - 0.13 [Fe/H), (3)

where to is the cluster age in Gyr units, MJo the absolute visual magnitude of the
main sequence turnoff (TO), Y the helium abundance, and [Fe/H] the metallicity
in standard notations. In turn, we have

MJO = V TO - mod, (4)

where V TO (the TO apparent magnitude) is the directly observable quantity, and


mod is the cluster distance modulus. Equations (3) and (4) now allow to esti-
mate the relative importance of the uncetainty in each of the four input quantities
(namely: VTO, mod, Y, and [Fe/HD in determining the total uncertainty of the
age determination. Table 2 illustrates how such errors propagate into age errors.
Column 1 lists the four IQs, column 2 gives the contribution of the various u(IQ)'s
to u(t)/t, column 3 provides a conservative estimate of the uncertainty affecting
the best current determinations of each the IQs, and finally column 4 gives the
corresponding contribution to the relative error in age.
133

Table 2

Contributions to the Relative Age Error cr(t)/t


Input Quantity
(IQ) cr(t)/t cr(IQ) cr( t)/t
V TO 0.86 cr(VTO) ±o.mlO 9%
mod 0.86cr(mod) ±0.m25 22%
y 0.99 cr(Y) ±0.02 2%
[Fe/H] 0.3 cr([Fe/H]) ±0.3 9%

Clearly the great villain is the error in the distance of the clusters. I estimate that
current distance estimates are typically affected by one quarter of a magnitude error
in the modulus - cr(mod) ~ 0.m25 - which immediately translates into a '" 22% error
in the derived cluster age. All other IQs convey substantially smaller errors. Since
the intoduction of CCDs the high photometric accuracy has allowed to determine
the V TO of a cluster with an accuracy perhaps better than one tenth of a magnitude,
which translates into a '" 9% error in age. The helium abundance of the clusters is
very well known, from either the R method, primordial nucleosynthesis, or empirical
determinations of the pregalactic abundance, which all indicate Y = 0.23 - 0.24
(see Boesgaard and Steigman 1985). Anyway, a ±0.02 uncertainty in Y gives a
negligible 2% error in age. I assume the metal content of the best studied clusters
to be uncertain by perhaps a factor of 2, i.e. 0.3 dex, which translates into a '" 9%
uncertainty in age. There may be a problem with the compo8ition of metallicity, e.g.
[O/Fe], I shall return to it later. All in all, the above discussion clearly indicates that
distance uncertainties dominate, and the question "How good are globular clu8ter
age8f" immediately becomes:
1.2 HOW GOOD ARE GLOBULAR CLUSTER DISTANCES?
All in all, just three main standard candles have been used (or are being considered)
for the determination of GC accurate distances, namely: RR Lyraes, subdwarfs, and
white dwarfs. Pros and cons of each method are discussed below, together with the
best perspectives for a significant improvement of their accuracy.
1.2.1 RR Lyrae di8tance8 have been widely used and there is a vast literature on the
subject (see Fusi Pecci et al. 1990, and references therein). Unfortunately, even the
nearest RR Lyrae star is too distant for a trigonometric parallax to be obtained with
an interesting accuracy, and the calibration of this standard candle relies on indirect
physical or astrophysical methods. A linear relation is usually assumed between the
absolute magnitude of RR Lyraes and metallicity:
M~R = a[Fe/H] + b, (5)
but different methods give different slopes and zero points. Table 3 summarizes
the current situation. The pu18ation method relies on the so-called period-8hijt ef-
fect whose actual size remains matter of debate (see e.g. Lee 1990, and references
134

Table 3

The M~R-metallicity relation: a[Fe/H] +b


METHOD a b References Ages
Pulsation 0.39 1.17 Sandage & Cacciari 1990 15.5-16.3
Baade-Wesselink 0.20 1.05 Liu & Janes 1988 19.3-15.9
0.05 0.90 Carney et al. 1988 22.0-15.0
HB Models 0.18 0.87 Sweigart et al. 1987 17.2-13.8
0.19 0.97 Lee 1990 18.4-15.0
MS fitting 0.37 1.29 Buonanno et al. 1989 17.8-18.2
0.20 0.84 Green et al. 1988 16.2-13.3

therein), and requires an independent estimate of the mass of RR Lyare stars, which
may not be presently obtainable with the necessary accuracy (Petersen 1990). The
Baade-Wesselink method is straightforward in principle, but makes extensive use
of stellar model atmospheres, while horizontal branch (HB) models clearly rely on
stellar evolution theory, and to a minor extent on model atmospheres for the color-
temperature relation and the bolometric correction*. An additional theoretical cal-
ibration, based on the luminosity of the bump feature on the luminosity function
of the red giant branch, is found to be consistent with the theoretical HB calibra-
tion (Fusi Pecci et al. 1990). Finally, main sequence fitting calibrates RR Lyraes
making use of sub dwarfs as standard candles, and therefore will be discussed in the
next subsection. There are three points worth making about the three calibrations:
1) they all intimately rely on theoretical modelling, and so may be all affected by
systematic errors which are difficult to assess; 2) they all agree surprizingly well,
i.e. to within 0.2-0.3 magnitudes; and 3) yet, such small differences have profound
implications for the derivation of globular cluster ages. I tend to conlude that the
three methods have possibly reached their intrinsic limit, and further progress along
these lines may be very slow.
Before passing to discuss the other methods, it is worth briefly commenting
on the implications for GC ages of the various RR Lyrae luminosity-metallicity
relations reported in Table 3. It has been known since a long time that among
galactic globular clusters the magnitude difference between the HB and the MS

* Note that the difference in zero point between Sweigart et al. and Lee cali-
brations arises from the former referring to zero age HB models, while the latter
includes the post-ZAHB evolutionary effects. Moreover, the values for the Sweigart
et al. ZAHB models make use of the bolometric correction from VandenBerg and
Bell (1985), while Lee's result follows from bolometric corrections compiled at Yale.
135
4.0 I I I Figure 1.- The HB to MS turnoff
magnitude difference V&. metallic-
ity for well &tudied galactic globu-
3.B • • lar clu&ter&. The figure is adapted
from Sarajedini and King (1989),

..
...... •
with the inclusion of the two ezcep-
'"
i!l
== 3.6 • • •
•• ••
;;- tional clu&ters Pall! (Gratton and

__ • ___ ...... __ .. _ • _____ • ___ .......... ___ 9 __ .. _____ .. ____ .... __ _
Ortolani 1988) and R 106 (Buo-
0" • nanno et al. 1990) .
3.4 t- •• ••• •

~
;;-
•• •
~

3.2 t--
",R106

I ",PAL 12

I I I
-2.2 -1.6 -1.0 -0.4
[Fe/H)

turnoff is nearly constant, independent of composition, with:

(6)

and Figure 1 illustrates the present situation. Therefore, by using eq.s (5) and (6)
into eq. (3) with Y = 0.23 one obtains:

Logt9 ~ 0.786 + (0.37 a - 0.13)[Fe/H] + 0.37 b, (7)

which provides an estimate of the cluster age for every RR Lyrae calibration (a, b).
The last column in Table 3 gives two such ages, which refer respectively to [Fe/H] =
-2 and -0.5. Clearly, none of the calibrations yield ages less than '" 13 Gyr, but
those with a flat slope (small value of a) produce large trends of the cluster age with
metallicity, with metal rich clusters turning out several Gyr younger than metal poor
ones. What matters for inflationary cosmology is obviously the age of the oldest
clusters, while the age spread is of great importance for our understanding of the
formation process of the Galactic halo and the Galaxy. An age trend with metallicity
such as that of some of the extreme cases reported in Table 3 (many Gyr) can hardly
be reconciled with the classical scenario of Eggen, Lynden Bell, and Sandage (1962),
in which the clusters form in only a few 108 yr. I would say that such a slow process
of metal enrichment and cluster formation can hardly be understood for the case of
a violently relaxing, nearly spherical system such as the Galactic halo might have
been at early times. It is worth emphasizing, however, that the derived trends result
entirely from the particular slope of the M~R - [Fe/H] relation, with a difference
of only rv one tenth of a manitude per dex in [Fe/H] can produce a macroscopic
136

Table 4

The Best Five Calibration Subdwarfs


STAR 1r [Fe/H] Mv u(Mv)
HD 25329 0.0547 -1.62 7.20 0.19
HD 64090 0.0444 -1.83 6.58 0.15
HD 103095 0.1182 -1.31 6.79 0.05
HD 134439 0.0395 -1.57 7.07 0.16
HD 134440 0.0395 -1.52 7.44 0.16

age dispersion. Different is the case of the two exceptional clusters so far discovered
which definitely appear to be younger than the other clusters, namely Pal 12 and R
106 (see Fig. 1). These are sparsely populated clusters which may well have formed
at later times in either accretion events, or in the disruption of small satellites of
the Galaxy triggered by either tidal encounters or starburst activity.
Given the highly indirect nature of the three calibrations so far discussed, a
direct, empirical assessment of the actual slope of the HB luminosity-metallicity
relation would be of great value. The color-magnitude diagrams of globular clusters
in the spiral galaxy M31 should allow such a direct determination, and HST was
supposed to provide the necessary data. Unfortunately, useful HST data could only
be obtained after the well known spherical aberration problem will be eliminated.
For the zero point, other methods such as those discussed below should allow a
precise determination.
1.2.2 Subdwarf$ di$tance$ are currently based on five such stars with reasonably
well known trig parallax. They are listed in Table 4, adapted from van Altena
et al. (1988). The errors in the parallax range from 2.6 to 4.7 milliarcsec, and trans-
late to the error in distance modulus given in the last column. The average error is
< u(mod) >= 0.m15, with the single best case (HD 103095, also known as Groom-
bridge 1830) being remarkably good, i.e. ±0.m05. What is the resulting accuracy of
the globular cluster distances obtained by fitting the cluster main sequence to the
five best subdwarfs? Unfortunately it is significantly larger than 0.m05. This follows
from the fact that we need to know the metallicity of the calibrating subdwarfs, and
this is typically affected by an uncertainty of 0.2-0.3 dex. Moreover, as seen from
Table 4, the five useful subdwarfs span a rather narrow range in [Fe/H], and there-
fore the slope of the main sequence luminosity (at a given color) vs. [Fe/H] needs
to be derived by using the Hyades cluster. In this way, van Altena et al. obtain
Mv(B - V = 0.6) = 4.64 - 0.87 [Fe/H]. So, in the case of Groombridge 1838, a 0.3
dex error in [Fe/H] propagates an error of "" 0.87 X 0.3 = 0.26 mag in the distance
calibration. All in all, while future efforts (and specially the Hipparcos mission)
can improve the trig parallax determinations for very many subdwarfs, the problem
with the uncertainty in metallicity may remain.
137

Figure 2.- 24 local calibrating white dwarfl,


diltinguilhed in DAI (open circlel) and non-

!
My DAI (filled circlel). The error in Mv fol-
lowl /rom the current error in the trig par-
10 allaz of each white dwarf, often much larger
than a few milliarc,econd,. Thil leavel am-
ple room for a ,ub,tantial improvement in
the future.
11

12

-.2 o .2 .4
B-V
1.2.3 White Dwarf al Itandard candlel have been proposed some time ago by Fusi
Pecci and Renzini (1979) as a possibility offered by HST. The basic idea is very
simple: to fit the WD cooling sequence of a globular cluster to either the theoretical
or empirical WD cooling sequence. The procedure is analogous to the classical
main sequence fitting to the local subdwarfs, but with some non-trivial advantages.
Indeed, the method does not involve metallicity determinations which inevitably
bring along their uncertainties, and there is no mixing length calibration involved.
In fact, WDs have virtually metal free atmospheres, coming either in the DA or non-
DA varieties (nearly pure hydrogen or pure helium, respectively), and their radius
is indensitive to the mixing length. In the most straightforward case of the fitting to
an empirical cooling sequence, such as that in Fig. 2, one has only to apply a small
correction taking into account the mass difference between the cluster WDs and the
local calibrators. Moreover, WDs are locally much more abundant than subdwarfs,
and therefore an accurate trig parallax can be obtained for a substantially larger
sample of calibrators. All in all - with HST working at nominal performance -
the distance modulus of a cluster should be obtained with an accuracy better than
,. . ., 0.m1, which translates into a better then 10% accuracy in age. Of course, along
with advantages there are also disadvantages. In particular, globular cluster WDs
are obviously much fainter than the other standard candles (either RR Lyraes or
the main sequence stars). From Fig. 2 one sees that the interesting local calibrating
WDs have Mv ~ 11 - 12, and since the modulus of the nearest globulars is ,....., 13
mag, this implies working at cluster WDs with V ~ 24 - 25 with a few percent
photometric accuracy. The spherical aberration presently affecting HST prevents
us from reaching this kind of performance, in particular in the cluster crowded field
138

which should be observed in order to ensure the collection of a sufficient number of


WDs. Again, we may have to wait for the deployment of WFPC II, unless adaptive
optics on large ground based telescopes comes first. But I believe that, ultimately,
white dwarfs will prove the best standard candles, able to improve the calibration
of all other calibrators (RR Lyraes, sub dwarfs, etc.).
1.3 rOlFe] AND THE LIKE
Equation (3) neglects one important aspect of cluster dating, namely the actual
distribution of the abundances of the elements heavier than helium. The models
used to derive eq. (3) assume a solar proportion, while very soon it was realised
that population II stars may significantly differ in this respect, and that the derived
ages can depend on the adopted distribution, in particular on the CNO /Fe ratio
(Simoda and Iben 1970). Unfortunately, most available opacity tables refer to solar
abundance ratios, and therefore only fewer evolutionary sequences with non-solar
proportions have been computed. On the other hand, observations indicate that halo
stars may be significantly enriched in oxygen (the most abundant of the heavies),
with rOlFe] up to rv +1 for [Fe/H] =-2, then declining to rv 0 for [Fe/H]=O (Abia and
Rebolo 1989). The effect is in the direction of reducing the derived age when rOlFe]
increases for fixed [Fe/H]. For example, VandenBerg (1990) estimates for the cluster
M92 an age of rv 17 and rv 14 Gyr, respectively for [0/Fe]=0.7 and 0.0. This result
allows a first order generalization of eq. (3) to include the effect of rOlFe] by adding
at its r.h.s. the term -0.12 rOlFe]. Clearly, the inclusion of this term allows to have
coeval globular clusters even if the coefficient (0.37 a - 0.13) in eq. (7) differs from
zero: suffice to have a suitable trend of rOlFe] with [Fe/H], such as that suggested
by the observations (see also Sandage and Cacciari 1990). Bencivenni et al. (1991)
have recently explored the effect of a systematic overabundance of all the a-capture
elements (C, 0, Ne, Mg, etc.), which appears more in line with nucleosynthesis
theory compared to an assumed overabundance of oxygen alone. They also find no
significant age-metallicity trend among galactic globulars.
1.4 HOW GOOD IS THE CLOCK?
I have so far discussed the uncertainties affecting the measure of the three input
parameters in eq. (3). But what about the uncertainty in the four numerical
coefficients in the same equation? i.e. how much accurate and reliable are the
stellar evolutionary models used in constructing eq. (3)? Clearly, it would be a
futile effort to improve the globular cluster distances, and then use them into the
wrong dating equation! The question has been extensively addressed elsewhere
(Renzini and Fusi Pecci 1988; Rood 1990), with the conclusion that no clear sign
of inadequacy has so far emerged in the canonical models. However, subtle errors
in the models may be hard to detect, and yet significantly affect the clock. The
check of the accuracy of the evolutionary models should correspondingly become a
cornerstone in our attempt of obtaining precision ages for globular clusters, able to
set strong and believable cosmological constraints.
139

As instructive examples, I will mention three illustrative cases of error, which


may possibly affect the canonical models from which eq. (3) has been derived. The
first example concerns an essential ingredient of the model input physics: radia-
tive opacity. Iben and Renzini (1984) cautioned that this may be the single most
uncertain ingredient, due to the complexities introduced by the very large num-
ber of chemical elements, ionization stages, energy levels and electronic transitions
which must be included in opacity calculations. Recently Rogers and Iglesias (1989)
have reported that their new opacities computed at Livermore show "substantial
differences" over the classical Los Alamos opacities. Could such differences affect
eq. (I)? In general, increasing the opacity makes dimmer stars, which then take
longer to burn their central hydrogen. For given stellar mass the turnoff luminosity
is decreased, and the time needed to reach it is increased. So the two effects tend to
balance each other, and the age-turnoff luminosity is less affected. However, larger
metal opacities favor the expansion of the envelope, and therefore the MS turnoff
may be anticipated (see Renzini and Fusi Pecci 1988), thus producing younger ages
for given MJo. New tracks and isochrones need to be constructed using the new
Livermore opacities in order to quantitatively determine the effect on ages.
The second example refers to the case of an old physical phenomenon which was
not included in canonical models, but which to some extent has to operate in real
stars: ionic diffusion. Stringfellow et al. (1983) argued that the inclusion of helium
diffusion could reduce the derived ages by as much as 15%. The main uncertainty
here comes from the calculation of the diffusion coefficient, which could certainly be
improved in the future, but major reduction of the cluster ages seems to be unlikely
(see also VandenBerg 1988b). The 2 Gyr range in eq. (2) is supposed to represent
the uncertainty introduced by diffusion. Since the case has been already discussed
in detail by Pierre Demarque at this conference, I will turn to my third example.
Finally, I would like to mention the case of the effects that new physics may
have on the dating method. A few years ago - near the peak expansion of the cold
dark matter scenario - an ingegnous way was suggested to solve at once for the
solar neutrino problem and for the dark matter problem. Indeed, Press and Spergel
(1985) advocated the so-called WIMP dark matter constituents being trapped in
the sun as an additional vehicle for the energy flow, thus allowing a reduction of the
solar central temperature and ensuing neutrino flux. Agreement with both the ob-
served solar neutrinos and the no = 1 cold dark matter requirements was achieved
for reasonable values of the basic WIMP parameters, i.e. mass and interaction cross
section with regular matter. Faulkner and Swenson (1988) then went on suggesting
that the inclusion of WIMPs may also lead to younger globular cluster ages, per-
haps even reconciling them with a straight Ho = 100, no = 1, and A = 0 universe.
However, WIMPs of the advocated kind were shown to potentially seriously inter-
fere with the convective energy transport in HB stars, possibly destroying the nice
agreement between the model and observed number ratio of HB to asymptotic giant
branch (AGB) stars in globular clusters (Renzini 1987). A detailed exploration of
the effects of WIMPs in HB stellar models has shown that the WIMP suppression
140

of core convection leads to thermal pulses having macroscopic effects on the mor-
phology and luminosity of the HB (Dearborn et al. 1990a). Dearborn et al. (1990b)
then speculate that the effect on the HB luminosity may lead to younger globular
cluster ages (if the HB models are used as standard candles, d. §1.2.1). How-
ever, the WIMP-induced thermal pulses produce both HB morphologies which have
no counterparts among observed clusters, and RR Lyrae period changes which are
far too large compared to the observations. I would conclude that non-wimpy HB
models compare more favorably with observations than wimpy ones. One argument
heretofore used in support of the WIMP solution to the solar neutrino problem was
the claim that wimpy suns would have been in better agreement with solar seismo-
logical data. However, recent solar seismology observations contradict these early
views, and standard solar models are now found in excellent agreement with the
data, while wimpy models appear to be definitely ruled out (Elsworth et al. 1990).
Perhaps more irrefutably than any astrophysical argument, S. Sarkar has reported
at this meeting that LEP experiments have already eroded the domain of the WIMP
parameter space (mass and cross section) which was advocated to solve the solar
neutrino problem. Surviving heavy cold dark matter candidates must be much more
massive than the 3-10 GeV range, and therefore would not dick when hitting stars,
and in any event would be too centrally concentrated to affect stellar energy trans-
fer. The epoch of widespread hope of using stars as WIMP detectors seems now
definitely concluded.

After over 20 years none of the astrophysical solutions of the solar neutrino
problem has survived. This leaves neutrino oscillations as the best current solution
to the problem. Given this hint in favor of massive neutrinos and the current diffi-
culties of the cold dark matter scenario (so widely pointed out at this meeting), it
seems that we are entering a phase in which the existence of a hot component to dark
matter is not considered so ugly as it was just a few years ago. On the other hand,
neither A '" 0 seems to look ugly anymore, and the pressure that for so long has
been exerted on us to provide younger cluster ages (and weaker neutrino emitting
suns) seems to have now almost vanished. In any event, such pressures didn't really
succeed, which indicates a real difficulty to escape the current estimates given at the
beginning of this paper. In any event, the best test for the accuracy and reliabil-
ity of the stellar evolution clock comes from the photometry of the largest possible
number of globular cluster stars, and in particular the construction of statistically
very representative luminosity functions for stars from near the turnoff to the base
of the giant branch (i.e. the so-called subgiant branch, SGB). This is indeed the
place where the effects of wrong opacities, helium diffusion, non convective mixing,
WIMPs, etc. which might anticipate or delay the turnoff (and then affect the clock!)
should more clearly show up (see Renzini and Fusi Pecci 1988). For example, iffor
some reason real stars were to complete the central hydrogen exhaustion consider-
ably past the main sequence turnoff, then the number of SGB stars per unit cluster
luminosity would be appreciably larger than predicted by the models.
141

Should we fit isochrones to cluster loci?


Having thoretical isochrones in one hand, and observed cluster loci in the other,
it may be difficult to resist the temptation of overlapping isochrones to data
points, pick up the single isochrone which bed fit& the data, read out its age,
and pretend to have accurately determined the cluster age. Still, I believe that
this temptation should be resisted, because the &hape of theoretical isochrones is
seriously affected by our current way of parameterizing the efficiency of the con-
vective energy transfer, via the so-called mixing-length parameter a. So, while
stellar luminosities are fairly insensitive to a, this is not the case for stellar ef-
fective temperatures. This is in particular true for the turnoff (TO) luminosity
and temperature (see for example VandenBerg 1988a, 1990), and therefore the
TO luminosity is a good clock, while the TO temperature (and along with it
the whole isochrone shape) is not. Should we conclude that fitting isochrones
to cluster loci is a futile exercise? No. Indeed, once the cluster distance is in-
dependently known, such fitting provides - instead of age - a calibration of a,
and makes it possible to use isochrone effective temperatures (in particular TO
temperatures) as clocks for those stellar systems in which TO luminosities are
not directly accessible to observations, and can be studied only in integrated
light. Therefore, the isochrone fit is an essential step in the calibration of pop-
ulation synthesis techniques, but it is methodologically incorrect if the goal is
the accurate dating of well resolved star clusters.

2. Magellanic Clouds Clusters and the Epoch of Galaxy Formation


Very high redshift galaxies are now being discovered in what used to be the empty
fields of 1 Jansky radiogalaxies, and objects at redshifts as high as 3.4 (Lilly 1988)
and 3.8 (Chambers, Miley, and van Breugel1990) have been reported. These galax-
ies are characterized by a "flat UV plus red bump" spectral energy distribution, and
Lilly (1988) has proposed that the presence of the red bump requires a rather old
age (>"" 1 Gyr) for the bulk of the stars in the galaxy. At such a high redshift the
universe itself is not much older than '" 1 Gyr, thus pushing the epoch of galaxy
formation back to extremely high redshifts, ZGF "" 10 - 20. In turn, this result pro-
vides further challenge to the standard cold dark matter scenario for the formation
of galaxies and structures, which prefers galaxies to form rather late, at ZGF '" 5.
Lilly's determination of the age of the bulk stellar population of the high redshift
galaxy relies on spectrophotometric models (Bruzual 1983) which assume a 1 Gyr
duration of the star formation process, and in which the red giant component is
not adequately treated. Either of these two aspects may have affected the age
determination, perhaps leading to an overestimate of the real age. More recently
Chambers and Charlot (1990) have re-analyzed the problem, concluding that even
142

the highest redshift galaxy in their sample (with z = 3.8) has an age of only f'V

3.3 x 108 yr, thus somehow releaving the pressure on the dark matter scenario.
What's new in Chambers and Charlot's approach compared to that of Lilly? First
they have assumed a shorter duration for the star formation process (i.e. '" 108
rather than 109 yr) and this certainly helps keeping the derived age low. Second,
they have attempted a somewhat more sophisticated approach to the population
synthesis, in particular for what concerns the treatment of the red giant evolutionary
phases, i.e. precisely of those stars which may be responsible for the red bump.
Although Lilly's red bump isn't really very red, corresponding to a rest frame
wavelength of only f'V 5000 A, still it is important to establish at which age an
evolving stellar population first turn red. From a theoretical point of view the
reddening of an evolving population is related to the first appearence of bright
asymptotic giant branch (AGB) stars, as soon as objects with a degenerate C-O
core appear (Renzini and Buzzoni 1986). A further increase in strength of the red
bump is predicted when later also the red giant branch (RGB) fully develops as a
consequence of the appearence of stars with degenerate helium cores. The precise
age at which these population phase transitions take place is however still rather
uncertain. The age at the AGB phase transition is in fact sensitive to both the
adopted mass loss on the AGB, and to what one is willing to assume about the
size of convective cores (the question of convective overshoot). Seemingly, also the
age at the RGB phase transition is sensitive to the adopted amount of overshoot,
and for the standard case (no overshoot at all) it is '" 6 X 108 yr, independent of
composition (Sweigart, Greggio and Renzini 1990).
Given the uncertainty affecting stellar models for the age of both the AGB and
RGB phase transitions, Chambers and Charlot (1990) have constructed semiem-
pirical spectral energy distributions implanting Magellanic Cloud GC spectra on
Bruzual's models, thus adopting Magellanic GCs as template stellar populations.
Chambers and Charlot assume the AGB and RGB phase transitions to take place
respectively at t = 3 X 108 and 1.2x10 9 yr. These numbers are consistent with quite
a substantial overshoot. Yet, the derived age of the youngest radiogalaxy turns out
to be just 3.3x108 yr, i.e. only 10% larger than the adopted age at the AGB phase
transition. Correspondingly, it would be important to understand whether there is
a direct link between this adopted age (3 X 108 yr) and the derived age of the high
redshift galaxy (3.3 x 108 yr). Such a link mayor may not exist, but if it does
then high SIN spectra of high red shift radiogalaxies do not suffice by themselves
to determine the epoch of galaxy formation. We would also need an accurate and
reliable dating of Mageilanic Cloud clusters, i.e. we also need a good clock.
The cluster NGC 1866 neither shows an extended RGB, nor contains bright
AGB stars (Frogel, Mould and Blanco 1990), and therefore it is younger than the
age at the AGB phase transition. It is a type III cluster in the classification of
Searle, Wilkinson and Bagnuolo (1980). On the other hand, type V clusters show a
well developed RGB and contain bright AGB stars, and one can conclude that both
(AGB and RGB) phase transitions take place within the age range spanned by type
143

3.5 ~I I I I Figure 3.- The integrated color, (V -


K)o v, (U - B)o for a ,ample of rich
Magellanic Cloud clu,ter,. The optical

,
o 8 6 and IR data are from van den Bergh
3 - Il De - (1981) and Peruon et al. (1983), re-
o ,pectively. The ,ymbol u,ed to repre-
!!IDe ,ent each clu,ter give, the type in the
o 80 clu,ter clauijication of Searle, Wilkin-
2.5 t-
dI
-
,on, and Bagnuolo (1980). A "0" in-
~ o dicate, that the SWB type i, not avail-
--
I
>
o able. Note that (U -B)o i, modly,en-
21- 1 1 e - ,itive to the turnoff temperature, i.e.
o age, while (V - K)o reveal, the contri-
bution of cool ,tar, (red ,upergiant"
1.5 t- - AGB, and RGB ,tar,). SWB type 9-~
o and ~ cluder, have been flagged in or-
23 der to ,how that they 'pan the large,t
range in (V - K)o.
I I I
-
-.6 -.4 -.2 0 .2
(U-B}o

IV clusters, i.e within somewhat more than 108 yr to perhaps 2 X 109 yr. This IV IV

is also confirmed by the large range in integrated V - K colors spanned by type


IV clusters, from V - K ~ 1 up to 3.3 (see Fig. 3), and for what concerns the
IV

AGB by the direct estimate of its contribution to the integrated light of the clusters
(Frogel, Mould and Blanco 1990). This jump in the rest frame V - K would appear
in the K -lOlL color in z ~ 3.8 galaxies, which unfortunalety is still out of reach.
Yet, the precise age at which the two phase transitions take place remains to
be accurately determined. Already several years ago an observational and theoret-
ical project has been udertaken aimed at determining the two ages, as well as at
assessing the effects of the two phase transitions on the spectral energy distribution
of stellar populations. The ultimate aim of the project is explicitly that of pro-
viding the astrophysicist and the cosmologist with an accurate, calibrated clock for
estimating the age of high redshift galaxies. To this end a systematic optical and
near-IR study of type IV (and IV-like) Magellanic GCs has been undertaken and
color-magnitude diagrams have been obtained from both the optical and near-IR
photometry (Buonanno et al. and Ferraro et al. in preparation). In parallell with
this observational project fine grids of stellar evolutionary sequences have been con-
structed up to the tip of the RGB (Sweigart, Greggio and Renzini 1990) and are
being extended up to the AGB. All in all this is a very laborious and extensive kind
of work, which for each cluster requires accurate photometry, calibrations, estimates
of the field contamination and completeness, cross identification of stars in optical
144

and IR frames, evaluation of the AGB and RGB contributions to the integrated
light in various bands and in bolometric, and finally the age determination. So we
have still only preliminary results, but the investigation is now close to completion.
One aspect of particular concern is the relative paucity of GCs in this most
interesting range of ages. With perhaps just a dozen type IV-like clusters (LMC
and SMC together don't provide much more than that), spanning some 2 Gyr, it may
be difficult to pinpoint the precise age of the two phase transitions with an accuracy
any better than "" 2 X 109 /12 '" 2 X 108 yr. Small sample size (within individual
GCs) may also be a problem, specially for the AGB phase transition. Ultimately,
it might be that we will have to live with this kind of uncertainty concerning the
age at which stellar populations first turn red, primordial galaxies first develop a
red bump, and perhaps with the uncertainty on the redshift of the el-och of galaxy
formation which may follow.
I am very grateful to Angela Bragaglia for the preparation of Figure 2, and to
Vincenzo Testa for the preparation of Figure 3.

References

Abia, C., Rebolo, R. 1989, Ap. J., 347, 186


Bencivenni, D., Caputo, F., Manteiga, M., Quarta, M.L. 1991, preprint
Boesgaard, A.M., Steigman, G. 1985, Ann. Rev. A&tr. Ap., 23, 319
Bruzual, G. 1983, Ap. J., 273, 105
Buonanno, R., Corsi, C.E., Fusi Pecci, F. 1989, A.dr. Ap., 216, 80
Buonanno, R., Buscema, G., Fusi Pecci, F., Richer, H.B., Fahlman, G.G. 1990 A.
J., 100, 1811
Carney, B.W., Latham, D.W., Jones, R.V., Beck, J.A. 1988, in Calibration of Stellar
Age", ed. A.G.D. Philip (Schenectady: L. Davis), p. 31
Chambers, K.C., Charlot, S. 1990, Ap. J. (Letter,,), 348, L1
Chambers, K.C., Miley, G.K, van Breugel, W. 1990, Ap. J., in press
Dearborn, D., Raffelt, G., Salati, P. Silk, J., Bouquet, A. 1990a, Ap. J., 354, 568
Dearborn, D., Raffelt, G., Salati, P. Silk, J., Bouquet, A. 1990b, Nature, 343, 347
Eggen, O.J., Lynden Bell, D., Sandage, A. 1962, Ap. J., 136, 748
Elsworth, Y., Howe, R., Isaak, G.R., McLeod, C.P., New, R. 1990, Nature, 347, 536
Faulkner, J., Swenson, F.J. 1988, Ap. J. (Letter,,), 329, L47
Frogel, J.A., Mould, J.R., Blanco, V.M. 1990, Ap. J., 352, 96
Fusi Pecci, F., Renzini, A. 1979, in A&tronomical U"e" of the Space Tele"cope, ed.
F. Macchetto, F. Pacini, M. Tarenghi (Geneva: ESO), p. 181
Fusi Pecci, F., Ferraro, F.R., Crocker, D.A., Rood, R.T., Buonanno, R. 1990, A"tr.
Ap., 238, 95
Gratton, R., Ortolani, S. 1988, A&tr. Ap. Suppl., 73, 137
Green, E.M. 1988, in Calibration of Stellar Age", ed. A.G.D. Philip (Schenectady:
L. Davis), p. 81
Then, I.Jr., Renzini, A. 1984, Phy". Rep., 105, 329
145

Lee, Y.-W. 1990, Ap. J., 363,159


Lilly, S.J. 1988, Ap. J., 333, 161
Liu, T., Janes, K.A. 1990, Ap. J., 354,273
Persson, S.E., Aaronson, M., Cohen, J.G., Frogel, J.A., Matthews, K. 1983, Ap.
J., 266, 105
Petersen, J.O. 1990, in Confrontation Between Stellar Pul&ation and Evolution, ed.
C. Cacciari and G. Clementini (San Francisco: A.S.P.), p. 402
Press, W.H, Spergel, D.N. 1985, Ap. J., 286, 679
Renzini, A. 1987, A&tr. Ap., 171, 121
Renzini, A., Buzzoni, A. 1986, Spectral Evolution of Galazie&, ed. c. Chiosi and A.
Renzini (Dordrecht: Reidel), p. 195
Renzini, A., Fusi Pecci, F. 1988, Ann. Rev. A&tr. Ap. 26, 199
Rogers, F.J., Iglesias, C.A. 1989, Bull. A.A.S., 21, 1078
Rood, R.T. 1990, in A&trophy&ical Age& and Dating Method&, ed. E. Vangioni-Flan
et al. (Gif sur Yvette: Ed. FronW~res), p. 313
Sandage, A., Cacciari, C. 1990, Ap. J., 350, 645
Sarajedini, A., King, C.R. 1989, A. J., 98, 1264
Searle, L., Wilkinson, A., Bagnuolo, W.G. 1980, Ap. J., 239, 803
Simoda, M., Iben, I.Jr. 1970 Ap. J. Suppl., 22, 81
Stringfellow, G.S., Bodenheimer, , P., Noerdlinger, P.O., Arigo, R.J. 1983, Ap.
J., 264, 228
Sweigart, A.V., Greggio, L., Renzini, A. 1990, Ap. J., 364, 527
Sweigart, A.V., Renzini, A., Tornambe, A. 1987, Ap. J., 312, 762
van Altena, W.F., Lee, J.T., Hanson, R.B., Lutz, T.E. 1988, in Calibration of Stellar
Age&, ed. A.G.D. Philip (Schenectady: L. Davis), p. 175
VandenBerg, D.A. 1988a, ibid, p. 117
VandenBerg, D.A. 1988b, in The Eztragalactic Di&tance Scale, ed. S. van den Bergh
and C.J. Pritchett (San Francisco: A.S.P.), p. 187
VandenBerg, D.A. 1990, in A&trophy&ical Age& and Dating Method&, ed. E. Vangioni-
Flan et al. (Gif sur Yvette: Ed. Frontieres), p. 241
VandenBerg, D.A., Bell, R.A. 1985, Ap. J. Suppl., 58, 561
van den Bergh, S. 1981, A&tr. Ap. Suppl., 46, 79
Weinberg, S. 1989, Rev. Modern Phy&., 61, 1
Discussion
ROWAN-ROBINSON: So can we have the oldest cluster ages at 10 billion years?
RENZINI: I would say it is not impossible, but unlikely.
DEMARQUE: The ranges usually quoted for globular cluster ages are not standard
errors. Internal errors in the fits are small. The error quoted are estimates of the
systematic effects.
RENZINI: Yes. In my talk I've made an effort to explicitly track the uncertainties
back to where they actually come from, so as to have a map of how errors and
146

systematic effects propagate. Sometimes age uncertainties as small as 1 Gyr have


been reported, or of 6.6% for an age of 15 Gyr. Such an accuracy would imply to
know the distance of the cluster with an accuracy of 0.066/0.86 = 0.m077, i.e. far
better than the distance of any standard candle, which is clearly nonsense. In these
cases the quoted error" actually describe not even an estimate of the systematic
effects, but the aestetic quality of a fit, which is rather uninteresting.
PENNY: Are the gaps which are seen in the luminosity function of some cluster
giant branches - and which disagree with theoretical models - significant?
RENZINI: No. To my knowledge in those clusters further work has shown that such
gaps tend to disappear when larger samples of stars become available.
HOYLE: At our time it was thought that the details of the light curves of RR Lyrae
stars might supply information of evolutionary importance. Has this possibility
been realized?
RENZINI: You may refer to the period changes in RR Lyrae stars, which were
thought to give an indication about the direction in which the horizontal branch
evolution proceeds (i.e. increasing periods for evolution from high to low effective
temperatures, and viceversa for decreasing periods). Both increasing and decreasing
periods have been found in individual clusters, so the period fluctuations are inter-
preted as due to the episodic readjustements of the chemical stratification around
the helium burning core, due to the so-called semi convective process. The average
period change in a cluster is still regarded as giving an indication on the prevalent
direction of evolution (Sweigart and Renzini 1979, A. & A. 71, 66).
D EMARQ UE: I was very interested in your discussion of population synthesis and its
application to Magellanic Cloud clusters. In this context, the super metal rich stellar
populations in the Galactic bulge are particularly significant since they are similar
to the stellar populations in elliptical galaxies. Preliminary theoretical explorations
indicate that contrary to what happens at metallicities lower than solar, where
metallicity increases result in redder populations, at metallicities much higher than
solar there is a reversal in color toward the blue.
RENZINI: I much agree on the relevance of Galactic bulge populations to understand
super metal rich populations so important in elliptical galaxies. Let me show here
the color-magnitude diagram obtained by Ortolani, Barbuy and Bica (1990, A. &
A. 236, 362) for the metal rich globular NGC 6553 in the Galactic bulge, whose
metallicity is within a factor oftwo solar. The most striking effect is in the V -(V -1)
diagram, where the upper part of the red giant branch bents down, getting fainter in
V and redder in V - 1, as severe Ti 0 blanketing strongly depresses the V luminosity.
This is due to the upper RGB and the AGB being both made up of M type giants,
as opposed to metal poor globular clusters where all the RGB and most AGB stars
are K-type giants. It may be worth noting that all population synthesis studies of
el1ipticals have so far assumed RGBs made up of K giants, rather than M giants.
THE LOCAL DISTANCE SCALE:
HOW RELIABLE IS IT?

M.W. Feast
South African Astronomical Observatory
P.O. Box 9, Observatory 7935, Cape, South Africa.

SUMMARY. A review of local calibrators and data for the Magellanic


Clouds, M31 and M33 suggests an increase in the local distance scale of
",10% over that recently adopted by several workers. The uncertainty in
this local scale is also ",10% and this sets a lower limit to the uncertainty in
the Hubble constant.

1. INTRODUCTION.
.... the fix'd stars cannot be less distant from the Earth than fifty
millions of leagues; nay, if you anger an Astronomer, he will set 'em
further. Bernard de Fontenelle: La Pluralite des Mondes translated
by John Glanvill 1688.
The distances to nearby galaxies, those within and just outside the local
group, are the basis for the calibration of large scale distance indicators such
as the Tully-Fisher relation. How reliable are these local distances?

2. DISTANCES FROM CEPHEIDS


Classical Cepheids are almost certainly the most important calibrators of
distances to nearby galaxies. Table 1 lists some of the ways that
observations can be used to determine Cepheid reddenings and distance
moduli. Estimates are given of the intrinsic widths of the various relations
(these are primarily due to the width of the instability strip) and of their
sensitivity to metallicity and reddening. The estimates of the metallicity
corrections are generally from Stothers (1988 = S88) formulation with the
helium abundance Y taken as varying with metal abundance (Z) according to,
6Y = 3.5 Ill. Rather similar metallicity corrections were derived earlier by
Caldwell and Coulson (1986 = CCII) and tests of metallicity effects are
discussed below (Section 6).

147
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 147-160.
© 1991 Kluwer Academic Publishers.
148

In any practical application the best way to proceed will depend on the
circumstances of the particular case including the expected accuracy and
number of the observations as well as knowledge of the relevant metallicity
etc.
Individual reddenings of good accuracy can be obtained from the
(B-V)/(V-I) diagram (Dean, Warren and Cousins 1978, Caldwell and Coulson
1985 = CCI) provided the observational data is of high qUality. The zero
point of the intrinsic line in this diagram is determined from Cepheids of
known reddening (e.g. those in clusters) and the line appears to have a low
intrinsic width. The line is metallicity dependent. However the size of the
metallicity effect can be estimated directly from model atmospheres (e.g.
those of Kurucz 1979 and of Bell and Gustafsson 1978 See CCI). Such
corrections depend only on stellar atmosphere theory (not Cepheid evolution
or pulsation theory) and are practically independent of the helium abundance
(Sonneborn et aL 1979).
Using reddenings determined in this way it is possible to establish
rather clearly the existence of a PLC relation at V and B-V in the LMC
(Martin et al. 1979, CCII) i.e.
My = alog P + {J(B-V)o + 'Y.
The value of {J derived in this way depends on the statistical treatment
used and values ranging from N2.1 to N2.7 have been derived from the LMC
(d. Feast and Walker 1987 = FW87, CCII, Caldwell and Laney 1991).
However this considerable range does not introduce any large uncertainty in
predicted values of My since in each solution the values of Il, {J and 'Y
automatically adjust to give the same My values on the PL or PC ridge lines
and differ in the extreme by ~om.1 even for stars at the edge of the
instability strip (d. FW87). In contrast, since the period-luminosity relation
at V (PL(V)) has an intrinsic width of Nom.8, errors up to ±OmA are
expected using this relation; though this is considerably reduced due to
compensating effect if the reddening is derived from the P-(B-V) relation.
Because of the differing sensitivities of the various Cepheid relations to
metallicity it is possible, with sufficiently accurate data (as in the Magellanic
Clouds), to combine several of them to solve simultaneously for distance,
reddening and metallicity (CCII, Feast 1988).
In future it should be possible to use a PLC relation at V and (V-I) to
~ood effect. This should be dispersionless and quite insensitive to metals
l Table 1). It is also rather insensitive to reddening, viz:
(m-M)o = V+3.66 log P-3.71 (V-I) + 3.39 + 1.07 Ev_r
This follows from the PLC relation of CCII. Notice that the higher the
estimated reddening, the greater the true modulus.
Sometimes reddenings of Cepheids are estimated from observed (B-V)
colours using a P-(B-V)o relation. This introduces a significant metallicity
dependence into moduli derived from a PL(V) relation (which in itself is
rather insensitive to metallicity) (cf. Table 1). The ptc (V, B-V) relation
corrected for reddening in the same way shows a rather similar net effect
(Table 1). The different conclusion of S88 (his equation 37) is not correct.
A change in reddenin~ affects both magnitude and colour in the PLC relation
and S88 equation (34) does not apply in this case.
149

TABLE 1. Metallicity and Reddening Dependence and Intrinsic Width of


Cepheid Relations.
(A) Effect on Reddenings.
(1) (B-V)/(V-I) Reddenings
intrinsic WIdth: Small (~om.07 in E(B-V)*)
Metallicity Effect: ~E(B-v)~m.06 (CCI)
(2) Period/(B-V) Reddenings
Intrinsic Width: ",Om.34 (LMC, mainly for P>10d CCII)
Metallicity Effect: ~E(B-v)",+om.12 (S88)
(3) Period/(V-I) Reddenings
Intrinsic Width: ",Om.24 (LMC, mainly P>10d CCII)
Metallicity Effect: ~E(V-I)",+om.01 (Caldwell from S88)
(B) Effect on Moduli
(4) PL(V) Moduli
Intrinsic Width: ",Om.8 (LMC, mainly P>10d CCII)
Metallicity Effect: -{)m.02 (S88)
Reddening Dependence: 3.3 E(B-V)
Net Metallicity Effect if Reddening from P /(B-V): -{)m.4
Net Metallicity Effect if Reddening from (B-V)/(V-I): +Om.2
(S) PL(H) Moduli
Intrinsic Width: ",Om.3
Metallicity Effect: -{)m.1 (Laney and Stobie 1986**)
Reddening Dependence: ",O.S E(B-V)
Net Metallicity Effect if Reddening from P /(B-V): -{)m.2
Net Metallicity Effect if Reddening from (B-V)/(V-I): -{)m.1
(6) PLC (V, B-V) Moduli
Intrinsic Width: Nearly Dispersionless
Metallicity Effect: ~m.4 (588)
Reddening Dependence: (Rv-P) E(B-V) ",0.8 E(B-V)
Net Metallicity Effect if reddening from P /(B-V): ",-{)m.S
Net Metallicity Effect if reddening from (B-V)/(V-I): ~m.4
(7) PLC (V, V-I) Moduli
Intrinsic Width: Nearly Dispersionless
Metallicity Effect: -{)m.04 (Caldwell after S88)
Reddening Dependence: ",1.07 E(V-I)
Net Metallicity Effect if Reddening from P /(B-V): +Om.OS
Net Metallicity Effect if Reddening from (B-V)/(V-I): -{)m.04

* For consistency with other width estimates this is listed as


0' is the upper limit to the intrinsic dispersions (CCI).
O'~ 12 where
150

** This is an observational estimate. The S88 formulation would predict


a very small effect since this gives an almost zero temperature shift of
the instability strip with metallicity due to the near cancellation of the
helium and metallicity term in S88 equation (22) if 5Y,.,3.5&.
Estimates are approximate since some effects are non-linear.
Metallicity effects are given for a decrease in abundance from solar to
four times lower.

The JHK region is useful for extragalactic Cepheids provided accurate


data on faint objects in crowded fields can be obtained. However there are
metallicity effects at these wavelengths which arise principally through the
metallicity dependence of the temperature of the instability strip (Laney and
Stobie 1986). This affects the bolometric correction.
Given data of sufficiently high quality there seems little doubt that it
would be best to proceed via colour/colour relations to obtain reddenings and
via PLC relations to obtain moduli since these relations seem to be
dispersionless or nearly so. However if the uncertainties in the basic data
are of the same order as the intrinsic scatter in the PL or PC relations, then
the above procedure has a less clear advantage over an approach (e.g.
Freedman 1985) using apparent PL moduli at a series of different wavelengths
to solve for the true modulus and absorption (Av). However some
metallicity corrections are still necessary and this method becomes uncertain
when applied to small samples of stars.
One can reduce the metallicity corrections applicable to any of these
methods to very small quantities by working in the wavelength range VRI.
However to do this with the presently available material (Le. to discard B
observations) would reduce the weight of solutions for extragalactic Cepheids
(including the Magellanic Clouds) significantly.

3. THE CEPHEID ZERO POINT FROM CLUSTERS.


The slopes of the various PL and PLC relations for Cepheids are most
accurately derived from Magellanic Cloud observations lcf. CCII, FW87,
Caldwell and Laney 1991). The best zero points for these relations are
obtained from Cepheids in open clusters in our Galaxy. [If for any reason
one restricted oneself to a very small number of clusters it would be best to
first use them to fix PLC zero points and the moduli of the Magellanic
Clouds and then to obtain PL zero points directly from the Clouds. This
would avoid uncertainties due to the finite width of PL relations]. A
significant part of the uncertainty in the zero points derived from clusters is
due to uncertainties in the cluster reddenings which affect the adopted moduli
rather sensitively through the main sequence fitting procedure (e.g. Walker
and Laney 1987). Because of the steepness of the main sequence an error in
the EB- V reddening of L).E produces an error of ,.,5L).E in the derived true
absolute magnitude of the Cepheid. However if (as is the case in practice)
the reddening scale is also based on Cepheids in clusters, an error in the
151

absorption correction of -3~E is also incurred for a programme star, leaving


a net effect on the Cepheid distance scale of N2~E.
The Cepheid-duster scale is based, ultimately, on the trigonometrical
parallaxes of nearby stars. It is most conveniently described in terms of the
distance modulus of the Pleiades implied by these trigonometrical parallaxes
(i.e. by fitting parallax stars to the Pleiades main sequence) since it is the
Pleiades main sequence (or a sequence of the same shape as that of the
Pleiades) which is the basis of the main sequence fitting technique for
clusters containing Cepheids. FW87 based their discussion on a Pleiades
modulus of 5.57 ± 0.08 which was derived by van Leeuwen (1983) using
trigonometrical parallaxes of 11 nearby stars.
Confidence that this is near the true value has been strengthened by
recent work. (1) VandenBerg and Poll (1989) used a theoretical zero age
main sequence, normalized to the sun and fitted to the mor:Jlhology of the
Pleiades sequence to derive a Pleiades modulus of 5.60. (2) The
semi-theoretical treatment of VandenBerg and Poll together with a
Hyades-Pleiades metallicity difference of ~[Fe/H] = 0.16 (Boesgaard and Friel
1990) gives a predicted Hyades modulus of 3.36 (for a Pleiades modulus of
5.60). The Hanson (1980) value is 3.30 whilst a more recent result (Schwan
1990) from FK5 proper motions is 3.37 ± 0.07. (3) Sandage and Cacciari
(1990) have shown that one can profitably extend the trigonometrical parallax
ap'proach. For the present discussion, 33 near main sequence stars with
t5(U-B) ~ 0.09 and with parallax probable errors ~7% (Jenkins 1952) were
used. Fitting them to the Pleiades main sequence and correcting them to
solar metallicity implies a Pleiades modulus of 5.63 ± 0.04 (the internal
error) or applying Lutz-Kelker (1973) corrections, 5m.68. The true
uncertainty in this value is likely to be larger than 0.04 both due to
uncertainties in the Lutz-Kelker corrections and also to the unexplained
systematic differences between different parallax catalogues (d. Jenkins 1952,
Strand 1963, Gliese 1969). These latter differences translate into absolute
magnitude differences of Om.04 to Om.10 for stars of the parallaxes under
consideration. Until this problem is settled (hopefully by RIPP ARCOS) it
seems best to retain the estimate of the uncertainty of this basic zero point
made by FW87 (=om.08). In the present discussion a distance scale
equivalent to a Pleiades modulus of 5m.60 has been adopted but it should be
remembered that if one based oneself on these 33 parallax stars alone this
modulus would increase to 5m.68. Note that the Lutz-Kelker correction
adopted (om.05) is likely to be an underestimate since the stars are selected
for low percentage errors in their parallaxes and thus, on the average, the
true percentage errors will be greater than the observed values.
Application of this basic zero point to the clusters containing Cepheids
normally assumes that they have Pleiades (i.e. near solar) metallicity. This
point remains uncertain if only because the Hyades, although older than the
Pleiades, is more metal rich. Boesgaard and Friel (1990) suggest that there
may be a range of ~[Fe/H]NO.2 amongst open clusters near the sun,
independent of age. In the main sequence fitting technique a change of ~
in [Fe/H] affects the derived modulus by a similar amount (VandenBerg and
Poll 1989). Thus some uncertainty must remain until accurate abundances
in the cafibrating clusters are derived. However the mean effect seems
unlikely to be very significant. For instance Luck and Bond (1989) find
that 153 F-M supergiants within 2 kpc of the sun have a mean [Fe/H] =
152

+0.023 :I: 0.018 (s.d. = 0.225, including an estimated observational dispersion


of "'0.2). This shows that in the mean the young objects have near solar
metallicity, although a scatter of ",0.1 is quite possible.
Some while ago there seemed to be a serious discrepancy between the
distances of clusters containing Cepheids as derived from the main sequence
fitting technique and from Stromgren photometry of B stars in the clusters.
This problem was essentially solved by the Balona-Shobbrook (1984)
recalibration of the Stromgren scale (cf. FW87). However a problem still
remains with the important cluster NGC 7790 which contains three Cepheids.
The latest main sequence fit (Romeo et al. 1989) gives a true modulus of
12.68 ± 0.15 for the cluster (Pleiades 5.60) whereas Stromgren photometry
(Schmidt 1981 adjusted to the Balona-Sho'bbrook calibration with Pleiades =
5.60) gives the distinctly smaller value of 12.25 :I: 0.10. Both solutions use
the same absorption (Av = 0.43 :I: 0.18). Further Stromgren photometry in
this cluster would be very desirable.
The results on the cluster Cepheids led FW87 to adopt a true modulus
for the LMC of 18.47 :I: 0.15. The standard error takes into account likely
calibration errors including the uncertainty in the Pleiades modulus. An
update based on slightly revised data and using the S88 rather than the CCII
metallicity corrections gives 18.52 (Caldwell and Laney 1991) or 18.55 if the
Pleiades modulus is taken as 5.60. In the following the LMC modulus is
taken as 18.55 :I: 0.15.
A check on these results can be made using main sequence fits to
young clusters in the Magellanic Clouds. Early results by this method gave
a low LMC modulus (",18.2) (e.g. Schommer et al. 1984, see also Feast 1988).
The new ZAMS discussion of VandenBerg and Poll (1989) resolves this
problem. Thus for NGC 2162, a key LMC cluster for the application of
this method, their results give a modulus of "'18.5 if [Fe/H] = -0.15 (the
likely value for young LMC objects d. Feast 1989).
Binaries containing Cepheids can also to be used to give a Cepheid zero
point. The most recent discussion (Evans 1990) gives results consistent with
our adopted zero point but the uncertainty in the individual estimates of
absolute magnitudes (",om.3) is too great for this method to give a stringent
test at present.
4. OTHER CEPHEID ZERO POINTS
A Cepheid zero point can be obtained from Baade-Wesselink determinations
of absolute magnitude. It remains uncertain how to realistically quantify
possible systematic errors in the various versions of this method (for either
Cepheids or RR Lyrae variables). Early work is briefly summarized by
FW87. Recently, Barnes et al. (1990) have used Baade-Wesselink
determinations for 101 Cepheids to derive a zero point for the PL(V) relation
which is brighter by ",Orn.2 than that given by clusters (the exact difference
varies with period). The internal standard error of their result is ±Orn.12.
Some uncertainty arises from the work of Sasselov et al. (1989) who observed
line asymmetries and doubling in the infrared spectra of Cepheids. This
leads them to believe that the true velocity amplitudes of these stars may
have been systematically underestimated. They suggest that if this effect is
taken into account the Baade-Wesselink absolute magnitudes will be made
153

brighter by between OmA and Om.7. If this is really the case then the
Baade-Wesselink scale is grossly different from that of the clusters.
Theoretical PLC relations can be obtained by combining computed
evolutionary tracks with pulsation theory (Iben and Tuggle 1975, Becker et al
1977, lben and Renzini 1984). PL relations can be obtained provided the
position of the edges of the instability strip in the HR diagram are computed
lor assumed). Zero points derived in this way are ",Om.5 brighter than the
cluster scale (cf. FW87). This is really just another manifestation of the
discrepancy between evolutionary and pulsational masses which may possibly
be solved by including convective overshooting in the models (e.g. Bertelli et
aL 1985).
Most Baade-Wesselink determinations proceed entirely empirically. A
different approach has been taken by Hindsley and Bell (1989) who combine
model atmosphere calculations with the Baade-Wesselink method to get
absolute magnitudes. Their PL zero point is ",Om.5 brighter than the cluster
one (and would be much brighter still if we accepted the Sasselov et al.
conclusion). One of the sources of uncertainty in their work is the need to
adopt a Doppler broadening velocity (DBV) in the atmospheric models. The
results are las they show) quite sensitive to this quantity. They adopt
DBV = 5 km S-l. A change of 2 km S-l changes the derived absolute
magnitude by Om.25.
Generally it has been considered that the random scatter in the
Baade-Wesselink absolute magnitudes will be too great to make it worthwhile
examining them directly for the existence of a PLC relation. Hindsley and
Bell attempt this but find no evidence for a PLC within their errors. Most
of the stars in their sample have periods less than 10 days where the
instability strip is known to narrow (CCIl, Walker 1987) making detection of
a PLC relation rather difficult (though one has apparently been found
amongst the 3-day Cepheids in NGC 1866 (LMC) (Welch et al. 1990)).
5. RR LYRAES AS CALIBRATORS
The RR Lyrae variables can be used to give an independent zero point for
the Cepheids via the distance to the Magellanic Clouds. They are of course
important also as distance indicators in their own right and as calibrators of
other old population indicators (e.g. the luminosity of the red giant tip of an
old population; the integrated iuminosity of globular clusters).
Baade-Wesselink luminosities indicate rather clearly that RR Lyraes show an
[Fe/H) - Mv(RR) relation, though different forms of the method~'ve rather
different resUlts (e.g. Liu and Janes (1990); Fernley et al. (1990) .
Recently Sandage and Cacciari (1990) compiled a list of Baade- esselink
results and derived
<Mv(RR» = 0.19 [Fe/H) + 1.03, (1)
The internal scatter about this line is quite small (om.06).
A relation based on horizontal branch theory (Lee et aL 1990),
<Mv(RR» = 0.17 rFe/H] +0.82, (2)
is about Om.2 brighter. Table 2 shows moduli deduced from RR Lyrae stars
with good photometry in the LMC using equations (1) and (2). The results
of equation (2) are in close agreement with the adopted Cepheid modulus of
154

the LMC. It may be noted that whilst Baade-Wesselink solutions give RR


Lyrae luminosities too faint (according to the Cepheid cluster scale) by ",Om.2,
the same method applied to Cepheids gives luminosities "'Om.2 (or perhaps
more) too bright (see section 4).

TABLE 2. RR Lyraes in the LMC

[Fe/H] (Mod)* (Mod)** No of Ref.


1 2
RRs

NGC 2210 -1.9 18.35 18.52 9 1


NGC 1786 -2.3 18.45 18.61 6 2
NGC 2257 -1.8 18.22 18.40 30 3
NGC 2257 (Field) -1.8 18.39 18.57 9 3

18.35±.05 18.52±.04
Cepheid Modulus = 18.55 (see text)

*
**
Modulus from equation 1 using data from reference
Modulus from equation 2 using data from reference
(1) Walker 1985 (2) Walker and Mack 1988 (3) Walker 1989.

The results from statistical parallaxes of RR Lyraes are dependent on


(uncertain) reddening corrections to the programme stars (cf. Feast 1987 and
references there). With Sturch reddenings the result is <Mv(RR» = 0.68
± 0.14 (Barnes and Hawley 1986). The mean [Fe/H] of these RR Lyraes is
-1. Thus adopting an [Fe/H] coefficient of 0.17 one finds
<Mv(RR» = 0.17 [Fe/H] + 0.85 (±0.14), (3)
very similar to equation (2). However the results with HI reddenings are
",Om. 1 fainter so that this is not a strong discriminant between equation (1)
and (2).
The discussion of Sections 3, 4 and 5 suggest that the adopted LMC
modulus of 18.55 ± 0.15 is secure although various remaining uncertainties,
mainly connected with the Baade-Wesselink method, need to be investigated.
6. TESTS OF METALLICITY EFFECTS
In calibrating Cepheids using the Magellanic Clouds and in applying these
results to more distant galaxies, the sensitivity of the various Cepheid
relations to chemical composition has to be taken into account (cf. Table 1).
Although theory is only used differentially in calculating these corrections,
various approximations and assumptions are necessary and the corrections are
only, at best, linearized estimates. Amongst the uncertainties is the need to
adopt a helium abundance at each metallicity (d. Section 2 and S88). In
155

view of this it is important to test the metallicity corrections empirically.


This is particularly necessary at the present time since Freedman and Madore
(1990 = FM90) have suggested that their results in M31 are consistent with
the absence of metallicity effects in BVRI and that such effects must be at
least a factor four smaller than the predictions of S88. This would be a
very surprising result since the main predicted effect is at B and is primarily
due to the change of (B-V) with metallicity at a fixed temperature (Le. it is
due to changes in line blanketing). Thus if the predictions are grossly in
error it suggests that models such as those of Kurucz grossly overestimate the
sensitivity of line blanketing to metallicity changes. Such a conclusion
would have wide repercussions (and not only for Cepheids).
Tests of metallicity effects are implicit in much of the previous work on
the moduli etc. of the Magellanic Clouds (CCII, Laney and Stobie 1986,
FW87, Feast 1988, 1989, Caldwell and Laney 1991) e.g. the corrections
necessary to make the difference in moduli between the two Clouds the same
as determined by PLC (V, B-V), PL(V) or PL(H). In fact using the
observation together with the theoretical predictions allows one to estimate
the metallicities of LMC and SMC Cepheids. These are found to be
consistent with metallicities derived from (non-Cepheid) young objects (e.g.
[Fe/H] -{l.2 (LMC) and -{l.5 (SMC) (c!. Feast 1989 for summary)). A
simple way to examine metallicity effects is to consider PL relations. If the
PL relations in both Clouds are the same at all wavelengths then so must
the PC relations be. Thus if we attribute any difference in the observed P,
(V-I) relations (say) in the two Clouds to differential reddening we can
correct for this and any remaining difference in P, (B-V)o relations (say) will
be a metallicity effect. Dr C D Laney has kindly carried out this
calculation usin~ the available BVI data in the Clouds (88 stars in all) and
finds 6(B-V)o lLMC-SMC) = +0.12 ± 0.03. This is clearly different from
zero though somewhat greater than the S88 prediction (0.08).
The test of FM90 uses Cepheids in three fields at different radial
distances from the centre of M3l. Since there is a radial metallicity
gradient in M31 comparison of data in these three fields allows, in principle,
a test of metallicity effects. The main conclusion of FM90 is shown in their
figure 6 which shows Cepheid moduli for the three fields derived without any
metallicity correction and a line showing the expected effects of metallicity
based on S88. Applying these corrections would lead to large (Nom.6) and
significant differences between the moduli of the different fields. However
the theoretical line in FM90 figure 6 is not relevant to distance moduli
calculated in the manner that FM90 use for the points in this figure (using
data at BVRI they fit curves through plots of (m-M) App'). against ).-1 to
derive (m-M)o and Av). This is clearly seen in Figure 1 which shows the
results obtained in this way with and without the theoretical corrections
(6M).) of S88. The only changes made in the analysis werej (a) to use
effective wavelengths at R and I appropriate to the Cousins system (e.g.
Bessell 1979) which are believed to apply to these data rather than the
Johnson effective wavelengths FM90 use andj (b) to use R)./Rv (R = ratio
of total to selective absorption) as the abscissa so that the relation in
(m-M) App). is linear, and a slightly different law of reddening (the standard
.56

van der Hulst curve 15). liMy and liMB are direct from S88 eqs. (24) and
(30). liM. was calculated by Dr JAR Caldwell from S88 (cf. Table 1)
and since liM. ~ liMy it is assumed that liMa is also the same. The
results are all adjusted to a true modulus of 18.55 and E(B-V) = 0.074
(CCI) for the LMC. There is no strong trend of modulus with metallicity
when the corrected values are used contrary to the conclusion of FM90 from
their figure 6. In fact the moduli with corrections are somewhat more
mutually consistent than without (mean modulus with corrections 24.51 ±
0.03, without corrections 24.55 ± 0.08) but clearly the data do not provide a
precise test of the metallicity corrections.
One can attempt a test in M31 restricting the data to B and Valone.
In this way FM90 obtain again the large discrepancy of Om.6 between theory
and observations (comparing their inner and outer fields). However this
result is very sensitive to errors in B and V (which multiply by Ry in the
analysis) and following their error discussion one finds the standard error of
their result to be ±Om.6 and again the test fails from lack of sensitivity. A
more complete discussion of the result of FM90 will be possible when their
observations are published in full.
The results on the Magellanic Clouds discussed above suggest that S88
may slightly underestimate the metallicity corrections. Some of this
difference might be due to errors in the adopted metallicities of the LMC and
SMC Cepheids. Better agreement between theory and observations could be
obtained if N ~ 0 (rather than ::::3.56Z) (cf. the comment at the foot of
Table 1). Further tests in nearby gafaxies would be of obvious importance
especially if the empirical results can be extended to supermetal-rich Cepheids
(some galaxies e.g. M51, M83 have metallicities ",0.5 dex above solar in their
inner regions (Pagel and Edmunds 1981».
11ll
I
24·6 )(


24·5

2H
• -
)(
)(

I I
24'3
FIELD -. I III
[Fe/H) -0'23 -0'52
Ay Itorrl 0'28 0'64 0·06
Ay luntorr, 0'53 0-72 -0·04
HEAN HODULUS
CORR 24·5110·03
UNCDRR 24'4510·08

Figure 1. The Modulus of M31 Cepheids as a function of radial distance


from the Centre. Crosses are uncorrected for metallicity effects; Filled
circles are corrected for metallicity effects as described in the text.
157

7. SO HOW RELIABLE IS THE DISTANCE SCALE?


FW87 suggested that the uncertainty of the scale out to the Magellanic
Clouds was :1:0.15 mag or perhaps slightly less. It would appear that this
remains a realistic upper limit to the uncertainty. Additional errors
accumulate in extending the scale to more distant galaxies. The current
situation can be illustrated by the cases of M31 and M33. In their
extensive review articles van den Bergh (1989) and Rowan-Robinson (1988)
adopt 24.3 and 24.25 respectively for the mean true modulus of M31.
However the new Cepheid modulus just discussed ~ave 24.51. It also
indicated that the reddening in the outer field (IV is low (Av = 0.06).
This is much lower than that sometimes adopted or foreground reddening
(e.g. Sandage and Tammann (1988) adopt Av (foreground) = 0.33). If we
take Av = 0.06 as the foreground reddening of the whole galaxy then the RR
Lyrae data of Pritchet and van den Bergh (1987) would give (m-M)o =
24.71 (with an RR Lyrae zero point fixed by the Cepheid distance of the
LMC i.e. equation (2 J above with the constant term changed to 0.79 from
the results of Table 2) or 24.56 retaining the absolute ma~tude calibration
used by Pritchet and van den Bergh. Van den Bergh's {l989) value was
(m-M)o = 24.33 :I: 0.15 with Av = 0.23. Evidently the newer data suggests
a modulus Om.2 or Om.3 greater than that adopted by van den Bergh and
Rowan-Robinson. Some uncertainty still arises from the basic observations
despite great technical improvements. Thus if we place the results of
Metcalfe and Shanks (1990) and FM90 for Cepheids in M31 field IV on the
same zero point (and metallicity correction) we obtain (m-M)A pp B = 24.76
(Metcalfe-Shanks) or 24.61 (FM90). Note that the Metcalfe-Shanks result is
close to the value (24.77) used by Sandage and Tammann (1988, 1990) for
this field as part ot their calibration of the nova scale (and the Virgo
distance). The FM90 value is distinctly lower.
In the case of M33, relatively recent Cepheid work gave a rather low
true modulus (24.05 (Christian and Schommer 1987); 24.1 (Freedman 1985);
24.25 (Feast 1988)). The values adopted in the reviews of van den Bergh
and Rowan-Robinson were 24.5 :I: 0.2 and 24.43 respectively. The most
recent Cepheid work (Freedman et al. 1990), reduced in the way described for
M31 (with LMC at 18.55) gives (m-M)o = 24.72.
Thus through a variety of circumstances the most likely modulus of the
LMC from Cepheids is now 0.08 greater than previously adopted and the
moduli of M31 and M33 are both increased by Om.2 or Om.3. These results
for the nearest galaxies suggest that the uncertainty in the distance scale
within the local g!OUP even averaged over several galaxies may well be at the
Om.2 level i.e. 10% in the distance.
Finally, if the current results on M31 and M33 are confirmed they will
not affect all more general distance indicators in the same way. For
instance van den Bergh (1989) bases his calibration of the Tully-Fisher
relation entirely on these two galaxies so that his distance modulus for Virgo
will increase from 31.61 to 31.83 (if all Virgo galaxies are treated as if at the
same distance). Freedman (1990) has recently derived a zero point for the
infrared Tully-Fisher relation based on Cepheid distances from five galaxies =
M31, M33, NGC 300, NGC 2403 and M81. Of these distances those of
158

NGC 2403 and M81 are the least secure and those of M31 and M33 the
most secure. Freedman's calibration in fact passes through the points for
these last two galaxies. Judging by M31 and M33 the distance scale adopted
here is 8% higher than that of Freedman. However although the true
modulus adopted for M91 is {)n.25 greater than that adopted by Sandage and
Tammann (1988) and Tammann (1988) in their calibration of the Do-a-
relation for spirals the lower proposed foreground absorption more than
compensates for this and leads to a Do-a- zero point (from M31) Om.33 fainter
than that derived by Tammann (1988).
ACKNOWLEDGEMENTS
I am grateful to Dr JAR Caldwell and Dr C D Laney for information and
helpful discussion.
REFERENCES
Barnes, T.G. and Hawley, S.L. 1986. Ap.J. 307, L9.
Barnes, T.G., Gieren, W.P. and Moffett, T.J. 1990 in "Confrontation between
Stellar Evolution and Pulsation" in press.
Balona, L.A. and Shobbrook, R.R. 1984. M.N. 211, 375.
Becker, S.A., Iben, I. and Tuggle, R.S. 1977. Ap.J. 218, 633.
Bell, R.A. and Gustafsson, B. 1978. A.A. Suppl. 34, 229.
Bertelli, G., Bressan, A.G. and Chiosi, C. 1985. A.A. 150, 33.
Bessell, M.S. 1979. P.A.S.P. 91, 589.
Boesgaard, A.M. and Friel, E.D. 1990. Ap.J. 351, 467.
Caldwell, J.A.R. and Coulson, I.M. 1985. M.N. 212, 879 (CCI).
Caldwell, J.A.R. and Coulson, I.M. 1986. M.N. 218, 223 (CCII).
Caldwell, J.A.R. and Laney, C.D. 1991. IAU Symposium 148 in press.
Christian, C.A. and Schommer, R.A. 1987. A.J. 93, 557.
Dean, J.F., Warren, P.R. and Cousins, A.W.J. 1978. M.N. 183, 569.
Evans, N.R. 1990. Ap.J. in press.
Feast, M.W. 1987 in The Galaxy ed. G. Gilmore and B. Carswell (Reidel:
Dordrecht) p 1.
Feast, M.W. 1988. Obs. 108, 119.
Feast, M.W. and Walker, A.R. 1987. Ann. Rev. A.A. 25, 345 (FW87).
Feast, M.W. 1988 in The Extragalactic Distance Scale ed. S. van den Bergh
and C.J. Pritchet (Ast. Soc. Pacif) p 9.
Feast, M.W. 1989 in The World of Galaxies ed. H.G. Corwin and L.
Bottinelli (Springer-Verlag) p 118.
Fernley, J.A., Skillen, I., Jameson, R.F., Barnes, T.G., Kilkenny, D. and Hill,
G. 1990. M.N. in press.
Freedman, W.L. 1985 in Cepheids: Theory and Observations ed. B.F.
Madore (Cambridge University Press) p 225.
Freedman, W.L. 1990. Ap.J. 355, L35.
Freedman, W.L. and Madore, B.F. 1990. Preprint (Ap.J. in press).
Freedman, W.L., Wilson, C.D. and Madore, B.F. 1990. IPAC preprint 54.
Gliese, W. 1969. Ver. Astr. Rech.-Inst. Heidelberg Nr. 22.
Hanson, R.B. 1980 in Star Clusters (IAU Symposium 85) ed. J.E. Hesser
(Reidel Dordrecht) p 71.
Hindsley, R.B. and Bell, R.A. 1989. Ap.J. 341, 1004.
159

Iben, I. and Tuggle, R.S. 1975. Ap.J. 197, 39.


Iben, I. and Renzini, A. 1984. Phys. Rep. 105, 329.
Jenkins, L.F. 1952. General Catalogue of Trigonometrica1 Stellar Parallaxes
(Yale).
Kurucz, R.L. 1979. Ap.J. Suppl. 40, 1.
Laney, C.D. and Stobie, R.S. 1986. M.N. 222, 449.
Lee, Y-W, Demarque, P. and Zinn, R. 1990. Ap.J. 350, 155.
Liu, T. and Janes, K.A. 1990. Ap.J. in press.
Luck, R.E. and Bond, H.E. 1989. Ap.J. Suppl. 71, 559.
Lutz, T.E. and Kelker, D.H. 1973. P.A.S.P. 85, 573.
Martin, W.L., Warren, P.R. and Feast, M.W. 1979. M.N. 188, 139.
Metcalfe, N. and Shanks, T. 1990. Preprint.
Pagel, B.E.J. and Edmunds, M.G. 1981. Ann. Rev. Ast. Astrophys. 19, 77.
Pritchet, C.J. and van den Bergh S. 1987. Ap.J. 316, 517.
Romeo, G., Bonifazi, A., Fusi Pecci, F. and Tosi, M. 1989. M.N. 240, 459.
Rowan-Robinson, M. 1988. Sp. Science Reviews 48, 1.
Sandage, A.R. and Tammann G.A. 1990. Basel Preprint 40.
Sandage, A.R. and Cacciari, C. 1990. Ap.J. 350, 645.
Sandage, A.R. and Tammann, G.A. 1988. Ap.J. 328, 1.
Sasselov, D.D., Lester, J.B. and Fieldus, M.S. 1989. Ap.J. 337, L29.
Schmidt, E.G. 1981. A.J. 86, 242.
Schommer, R.A., Olszewski, E.W. and Aaronson, M. 1984. Ap.J. 285, L53.
Schwan, H. 1990. A.A. 228, 69.
Sonneborn, G., Kuzma, T.J. and Collins, G.W. 1979. Ap.J. 232, 807.
Stothers, R.B. 1988. Ap.J. 329, 712 (S88).
Strand, K.Aa. 1963 in Basic Astronomical Data Ed. K.Aa Strand (University
of Chicago Press) p 55.
Tammann, G.A. 1988 in The Extragalactic Distance Scale ed. S. van den
Bergh and C.J. Pritchet (Ast. Soc. Pacif.) p 282.
VandenBerg, D.A. and Poll, H.E. 1989. A.J. 98, 1451.
van den Bergh, S. 1989. Astr. Ap. Reviews 1, 111.
van Leeuwen, F. 1983. Ph.D. Thesis, Leiden.
Walker, A.R. 1985. M.N. 212, 343.
Walker, A.R. 1987. M.N. 225, 627.
Walker, A.R. and Laney, C.D. 1987. M.N. 224, 61.
Walker, A.R. 1989, A.J. 98, 2086.
Walker, A.R. and Mack, P. 1988. A.J. 96, 1362.
Welch, D.L., Mateo, M., Cote, P., Fischer, P. and Madore, B.F. 1990.
Preprint.
160

DISCUSSION:

Cannon: Do you assume that all of the Cepheid variables within each of the Magellanic
Clouds have the same metal abundance, and is this assumption justified?
Feast: Supergiants and HIT regions suggest that the metallicity of young objects is rather
uniform within each Cloud. The scatter of Cepheids in each Cloud about a mean line in
the B-V /V-I diagram is quite small. Some scatter is expected due to differential reddening
between Cepheids, thus abundance variations which would also produce a scatter must be
small.
Pierce: Given the uncertainty of 10% in the local distance scale, how may Cepheids do
we need to observe in an external galaxy in order to reduce the uncertainty to that of the
local distance scale?
Feast: In principle high quality BVI data on quite a small number of Cepheids should
be sufficient. However, for faint stars in crowded fields the main current uncertainty is
probably the observational errors which may well still be large (including the difficulty
of obtaining enough observations to produce really well defined BVI light curves for each
variable.)
DISTANCES TO VIRGO AND BEYOND

Michael Rowan-Robinson
School of Mathematical Sciences
Queen Mary and Westfield College
Mile End Rd. London E1 4NS

ABSTRACT: New distance measurements to Virgo and beyond are reviewed. The
discrepancy between the Type la supernova and infrared Tully-Fisher methods remains
unresolved. Two new distance methods based on the luminosity function for planetary
nebulae and on surface brightness fluctuations in ellipticals and lenticulars are briefly
discussed. Analysis of IRAS maps of 100 l1m emission from interstellar dust appear to
resolve the controversy about extinction towards the Galactic poles. A new model for the
local flow based on the aDOT IRAS galaxy redshift survey is applied to the analysis of the
Hubble diagram. The implications of the current best estimates of Ho and Q o for
inflation are discussed.

1. INTRODUCTION

Those of you who have been lucky enough to find time to dip into Gibbon's hilarious book
"The Decline and Fall of the Holy Roman Empire" will know that the history of the early
church was characterized by ferocious schisms. The Great Schism of 1378-1414,
during which there were three separate Popes, was not over the distance of the Virgo
cluster, but its ferocity was of comparable intensity. In this review I will try to pick
out some of the salient features of recent work on the distance scale, sifting the
pronouncements of the visionaries and schismatics who make up our field.

In "The Cosmological Distance Ladder" (Rowan-Robinson 1985, RR85) I reviewed work


on the distance scale up to the end of 1982. A supplementary review (Rowan-Robinson
1988. RR88) covers the period January1983- December 1987. The present review
does not attempt to give the level of detail of those reviews. Several other reviews on the
distance scale and the value of the Hubble constant have appeared in the period 1988-
90. Dressler (1988) reviewed the values of Ho , qo, Q o • Ao and to' Tully (1989)
reviewed the distance to Virgo, comparing distance estimates by Tammann. Rowan-
Robinson and van den Bergh. and arriving at a value for Ho. van den Bergh (1989)
reviewed distances out to Virgo and the value of Ho. Sandage and Tammann (1990)
reviewed the distance to Virgo, and of more distant clusters relative to Virgo, arriving at
a value of Ho "freed from all local velocity anomalies". Table 1 summarises the values of
the distance of Virgo and of Ho adopted in these reviews. There is a good concensus
161
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 161-171.
@ 1991" Kluwer Academic Publishers.
162

Table 1: Comparison of estimates of distance to Virgo and of Ho

authors dVirgo Ho
RR 85 18.4 ± 1.6 67 ± 15
RR 88 19.0 ± 1.2 66 ± 10
Dressler 88 60-80
Tully 89 15.2 ± 1 87 ± 10
vdB 89 20 ± 2 67 ± 8
ST 90 21.9 ± 0.9 52 ± 2

between Dressler, van den Bergh and myself about the value and uncertainty of Ho. with
Tully. and Sandage and Tammann representing values at the extreme ends of that
concensus range. There is reasonably good agreement between different methods and
different authors about relative distances between Virgo and more distant clusters. a
point emphasized by RR88 and Sandage and Tammann (1990).

One of the major current problems with the distance scale is the 1 magnitude systematic
error in one or both of the Type la supernova and infrared Tully-Fisher methods
(RR88). Figure 1 shows the range of absolute magnitude of Type la supernovae in the
Virgo cluster, compared with the range of values predicted by theoretical models. It
appears that the predicted range of masses of 56Ni generated. 0.4 -1.0 Mo. is too small
to account for the observed range of absolute magnitudes. Rowan-Robinson (1988)
proposed that the lower end of the mass-range for 56Ni production should be 0.2 Mo.
comparable to that required for Type Ib supernova models. The mean absolute blue
magnitude for Type la supernovae at maximum light should then be revised from -19.6
to -19.3, bringing supernova distance estimates into line with other methods. Infrared
Tully-Fisher distance estimates could be brought into line with other methods by
assuming that there is a Malmquist-type bias involved in comparing galaxies in a rich
cluster like Virgo with those in poor groups like the Local and M81 Groups, of about 0.5
magnitudes in the distance modulus.
163

2. NEW DISTANCE MEASUREMENTS TO VIRGO AND BEYOND

2.1 The Dn-O' method

Faber et al (1989) have given distances relative to Coma for a number of groups and
clusters from the Dn-O' method, of which 17 have at least 4 galaxies measured. Lucey
and Carter (1988) have given distances to 5 southern clusters and Lucey et al (1990)
have given distances to 2 northern clusters. These may be converted to absolute
distances using Dressler's (1987) Dn-O' measurement of the distance modulus of Virgo
by comparing Virgo ellipticals and lenticulars with the bulges of M31 and M81, 110 =
31.5, and the value given by Dressler et al (1987) for the difference in distance
modulus between Virgo and Coma by this method, 1l0(Coma) - Ilo(Virgo) = 3.6. A
rather different calibration is given by Pierce (1990) based on planetary nebulae
distances to Leo. However this results in a value for Ilo(Virgo)-llo(Leo) = 0.9, whereas
the value found by RR88 is 0.17 mag.

2.2 Tully-Fisher method (B band)

The major controversies have been over the true dispersion about the calibration
relation and over line-of-sight structure in Virgo. RR88 used 0' = 0.7 mag., based on
Richter and Huchtmeier (1984: 0' = 0.74), van den Bergh (1984: 0' = 0.78), Kraan-
Korteweg et al (1988: 0' = 0.7). Pierce and Tully (1988), on the other hand, found 0' =
0.50 in Virgo and 0' = 0.35 in Ursa Major. Fouque et al (1990) found 0' = 0.46 for 18
calibrating galaxies and 0' = 0.76 for Virgo. Sandage (1988) found 0' = 1.0 for a
distance-limited sample of galaxies. Burstein and Raychaudhuri (1989) found 0' = 0.4
for Virgo, the difference from Kraan-Kerteweg et al being attributed to line-of-sight
effects and pooorly determined magnitudes. The resulting values for the Virgo distance
modulus are 110 = 31.60 (Kraan-Korteweg et al), 31.40 (Fouque et al), 31.03 (Pierce
and Tully). Burstein and Raychaudhuri argue that the distance modulus is so dependent
on the choice of calibrators as to be indterminate at present.

2.3 Type la supernovae

Cadonau and Leibundgut (1990) have made a compilation of data on Type 1a supernovae
to 1985. This has been used by Tammann and Leibundgut (1990) to derive a value for
Ho = 46 ± 10, compared with the value of 58 derived by RR88. This is partly due to a
less restrictive criterion for inclusion in the sample by Tammann and Leibundgut and to
their somewhat extreme value for MB(max), -19.79.

Barbon et al (1989) have published an updated edition of the Asiago Catalogue of


Supernovae, giving details of all supernovae discovered up to the end of 1988. There are
three of these with good quality magnitudes at maximum light which do not appear in
Table 3.9 of RR88: 1986a in N3367, 1986g in N5128 and 1987d in M+00-32-01.
That in N5128 poses a problem, because it is 4 magnitudes fainter than 1972e in
N5253, which is supposed to be in the same group of galaxies as N5128.
164

2.4 Planetary nebulae

Jacoby (1989) and Jacoby et al (1989, 1990) have advocated the use of the bright end
of the luminosity function for planetary nebulae in galaxies as a distance estimator.
Ciardullo et al (1989) use the method to derive a distance of 10.0 Mpc to the Leo group,
whereas the value found by RR88 from other methods is 17.6 Mpc. Jacoby et al (1989,
1990) derive a value for d(Virgo)/d(Leo) = 1.5, compared to the value 1.08 found by
RR88. Some possible problems with the method are discussed by Sandage and Tammann
(1990).

2.5 Dwarf ellipticals

Bothun et al (1989) use dwarf elliptical galaxies to derive d(Fornax)/d(Virgo) = 0.93


± 0.07, consistent with the value of 1.08 found by RR88. However they give
d(Centaurus)/d(Virgo) = 1.91 ± 0.07, compared with 2.70 by RR88.

2.6 Surface brightness fluctuations in ellipticals and lenticulars

Tonry and Schneider (1988) have developed a distance estimate based on the surface
brightness fluctuations in ellipticals and lenticulars due to unresolved stars. Using M32
to calibrate the method, Tonry et al (1989) derive a distance to Virgo of 14.0 Mpc.

The last 3 methods all give distances significantly lower than those adopted by RR88.
Although each of these methods is of interest, they do not yet appear to be sufficiently
well developed to give reliable distances. Other methods discussed recently include HII
region luminosities (Melnick et al 1988) and the luminosity of Sci galaxies (Sandage
1988).

3. INTERSTELLAR EXTINCTION ESTIMATES DERIVED FROM IRAS 100 ~m


MAPS

0.2 magnitudes of the 1.5 mag. difference in the distance scales of de Vaucouleurs and of
Sandage and Tammann was due to disagreement about the extinction towards the Galactic
poles (RR85).

Rowan-Robinson et al (1990a) have modelled the IRAS zodiacal emission at 12, 25, 60
and 100 ~m and subtracted it out from the total IRAS extended emission, to generate
maps of the emission from interstellar dust (Rowan-Robinson et al 1991). The
correlation of the intensity of the 100 ~m emission from interstellar dust with HI
column-density is excellent, with a relationship

1100 (MJy/sr) = 0.85 { N(HI)/10 20 cm- 2 }

holding over most of the sky away from the Galactic plane. From this Rowan-Robinson et
al (1991) deduce that AV = 0.05 at the North Galactic pole, and 0.08 at the South
165

Galactic pole. For comparison Sandage (1975) argued that AV = 0 at the Galactic poles,
while de Vaucouleurs et al (1976) derived AV = 0.14 at the NGP and 0.16 at the SGP.
The impact of the new estimates on the distances given in Rowan-Robinson
(1985,1988) is < 1%.

4. A NEW MODEL FOR THE LOCAL FLOW

One of the problems which has bedevilled the search for Ho has been ignorance of the
local flow. Estimates of our infall velocity to Virgo have usually been in the range 100-
300 km/s, for example (see refs in RR85, RR88). This range can already account for a
20% variation in the derived value of Ho.

Rowan-Robinson et al (1990b) have described how the aDOT IRAS galaxy redshift
survey can be used to map the local galaxy density (see also Saunders et al 1991, Kaiser
et al 1991, Frenk et al 1991) and hence map the local velocity flow field. To obtain
consistency with the observed motion of the Local group towards the microwave
background radiation, Rowan-Robinson et al derive, for A = 0,

b no-0.6 = 1.23 ± 0.23 (1)

consistent with no = 1 and a modest bias b = 1.23 ± 0.23. or, if there is no bias. with
no = 0.7 (+0.3.-0.2). .. no -0.6 " is in fact Peebles' (1980) function f(n o )' If A "*
0, but k = 0, it is easy to verify that f(no ) is well approximated by no -0.55 • so that
eqn (1) can be satisfied by b = 1 (no bias). ni = 1 (Le. k = 0), no = 0.7. Thus the
aDOT data do not permit a k = 0 universe with low no'

Rowan-Robinson et al (1990b) model the flow derived from the IRAS galaxy map in
terms of a simple model involving 12 superclusters. almost all of them well-known
objects. For 9 of them distances were taken from RRBB. while for the remainder
distances were derived from the recession velocity assuming Ho = 66 , the adopted value
in RR88.

To incorporate the flow model into a solution for Ho it is necessary to proceed


iteratively. adopting a value for Ho. deriving the flow model, and then correcting the
velocities of the full sample of groups and clusters for which distances are known to find
the appropriate value of X2 for that value of Ho.

I give here the preliminary result of a new calculation of this type (Rowan-Robinson,
1991. in preparation). Firstly the flow model of Rowan-Robinson et al (1990b) has
been improved by including additional clusters visible on the density maps of Saunders
et al (1991) and Kaiser et al (1991). and on the cone diagrams of Fairall et al
(1990). The threshold for acceptance of a cluster as real has been reduced from 4-cr to
2-cr. This results in a flow model with 25 clusters (Table 2). Secondly new cluster
distance estimates have been included in the solution, of which the most significant are
166

Table 2: Parameters for 25 clusters used in flow model

name b dist. amplitude


(Mpc)

VIRGO 287 70 19.1 2219


ERI-FOR 207 -52 20.8 655
UMal(N) 145 59 22.7 973
CENT 302 22 53.8 2077
N1600 200 -33 60.0 1753
HYDRA 270 26 65.5 1897
CANCER 203 29 72.4 1330
N.CEN 310 29 69.4 2319
PAVO 332 -24 79.1 2301
N194 117 -60 75.4 1304
PER-PISC 139 -22 86.9 1544
187+55 187 55 87.4 1293
S4 281 0 87.4 2084
COMA 58 88 110.5 1701
A2666(S5) 107 -34 113.6 2319
S3 273 -44 119.3 2829
A496(KB) 210 -30 128.6 2104
A2197 65 44 139.9 4218
A260 140 -33 132.4 5163
S6 202 -56 137.8 3403
I-ERC 32 45 148.6 2922
S894(S7) 19 -35 197.4 7355
SHAPLEY 310 32 209.1 3308
KC 165 15 214.3 4647
KF 30 -50 214.3 5858

where distances not known, Ho = 70 used


167

M(56NillM0 =0·2 0·4 1·0

I
-17 -18
I o ••

MS(max)
I
-19
0 00.'
-20
I
I
Fig 1: The absolute magnitude of Type I supernovae in the Virgo cluster for an assumed
distance modulus of 31.40, excluding known Type Ib or Type Ipec: filled circles, known
Type la; open circles, subtype unknown. The values predicted for models involving the
deflagration of different masses of 56Ni are indicated. Most theoreticians argue that this
mass should be in the range 0.4-1.0 Mo (from RR88).



10000
• • • •

V (km/s)
.~.~.: •

..:- ...
./.,.
. .~<
5000

/:..
~


o 50 100 150
d (Mpe)

Fig 2: Hubble diagram for groups and clusters with reliable distances (weight ~ 3.5).
The velocities are corrected for the local flow model described in the text, based on the
25 clusters of Table 2.
168

the new On-a distances reviewed in section 2.1 above. For each trial value of Ho , X2
has been calculated from

X2 = 1: [( 10g(Vld) - log Ho }/{ 0.2 a(~o)}]2 ,

where V is the recession velocity corrected for the flow and d is the distance, for 84
clusters having distances with weight ~ 3.5 (see RR85 for definition of weight).
Distances are calculated incorporating the prescription of RR88 to remove the Type la
sUBernovalTully-Fisher systematic errors (ie ~~o = -0.3 for Type la s.n. distances,
and +0.5 for Tully-Fisher distances). The minimum X2 of 106.8 is found for Ho = 70.
The corresponding Hubble diagram is shown in Fig 2. The slight increase from the value
of 66 given by RR88 is almost entirely due to the new flow model. The formal statistical
uncertainty in Ho is ± 2 km/s/Mpc, but including the effect of systematic errors, the
uncertainty should be increased to 10 km/s/Mpc (RR88). Equation (1) becomes

b no -0.6 = 1.12 ± 0.23 (2) .

An interesting point is that if velocities are corrected only for the sun's velocity with
respect to the Local group, corresponding to the assumption that the motion of the Local
group with respect to the microwave background is generated at V > 15000 km/s, X2
increases to 183.4. If velocities are corrected also for the motion of the Local Group
with respect to the microwave background, corresponding to the assumption that this is
an entirely local motion, then X2 = 184.2. Thus the flow model does appear to
significantly sharpen the Hubble diagram. This in turn reinforces the claim that the
motion of the Local Group with respect to the microwave background has been correctly
explained by the ODOT galaxy density maps and that eqn (2) is therefore correct.

5. IMPLICATIONS FOR INFLATION

Pulling together the evidence on no and Ho , there are several possible options as far as
inflationary cosmology are concerned:

(A) Assume that IRAS gives a correct picture of the matter distribution (with a modest
amount of bias) and that the above derivation of Ho is correct. Then

no = 1 , Ho = 70, nb = 0.03, and for A = 0, to = 9.4x10 9 years. (3)

This is fine for inflation, but the age of the universe is much too short for the globular
clusters, for which to = 13.5 ± 1 b.years.

(8) As (A) but assume that Tully-Fisher method underestimates distance moduli by 1
mag. (and that the On-a method underesimates them by 0.5 mag.), so that Ho = 50. Then
for A = 0, to = 13.1 b.y., which is consistent with globular cluster ages.
169

(C) Assume that IRAS galaxies are biassed towards regions of low density, so that

no = nb = 0.03, b = 0.14 (from eqn (2) ), Ho = 70 (4)

and then to = 14.4 b.y. if A = 0 ( which is fine, but would be inconsistent with
inflation). If ni = 1 (k = 0), so that inflation is allowed, then to =23.1 b.y., which
seems a bit on the long side of most estimates of the age of our Galaxy.

(0) Assume that the motion of the Local Group with respect to the microwave
background radiation is generated at V > 15,000 km/s, that the agreement of the C8R
and IRAS dipole directions is fortuitous, and that the sharpening of the Hubble diagram
when the IRAS flow model is incorporated is also fortuitous. In this case we are again
free to adopt eqn (4) above.

None of these solutions is satisfactory. I find solution (0) the most unsatisfactory.
Solution (A) requires there to be substantial errors in several well-established areas of
astrophysics, so is almost equally unsatisfactory. The choice between (8) and (C) will
depend on individual taste, but perhaps (8) is the least contrived.

To conclude: the assumption that IRAS galaxies trace the total matter distribution with
only a small amount of bias results in a value for no close to the unity value predicted by
inflation. However the current best estimate of Ho , 70, then gives an age of the universe
too short to be consistent with current estimates, particularly that from globular
clusters. This can be resolved by assuming that there is a substantial systematic error
in the Tully-Fisher distance method (and perhaps other methods), so that the true Ho =
50. An inflationary model with a low value for no (and hence no requirement of non-
baryonic dark matter), positive cosmological constant, and a rather long age for the
universe, is consistent with Ho = 70 provided IRAS galaxies are strongly biassed
towards regions of low matter-density.
170

References

Bothun, G.D., Caldwell, N., and Schombert, J.M., 1989, A.J. 98, 1542
Barbon, R., Cappellaro, E., and Turatto, M., 1989, AA Supp 81, 421
Burstein, D., and Raychaudhuri, S., 1989, Ap.J. 343, 18
Cadonau, R., and Leibendgut, B., 1990, AA Supp 82, 145
Ciardullo, R., Jacoby, G.H., and Ford, H.C., 1989, Ap.J. 344, 715
de Vaucouleurs, G., et ai, 1976, 2nd Reference Catalogue of Galaxies (Univ. of Texas)
Dressler, A., 1987, Ap.J. 317, 1
Dressler, A., et ai, 1987, Ap.J. 313, 42
Dressler, A., 1988, 3rd ESO-CERN Symposium
Faber,S., et al 1989, Ap.J.Supp. 69, 763
Frenk, C., et ai, 1991, in preparation
Fouque, P., Bottinelli, L., Gouguenheim, L., and Paturel, G., 1990 349, 1
Jacoby, G.H., 1989, Ap.J. 339, 39
Jacoby, G.H., Ciardullo, R., Ford, H.C., and Booth, J., 1989, Ap.J. 344, 704
Jacoby, G.H., Ciardullo, R., and Ford, H.C., 1990, Ap.J.356, 332
Kaiser, N., et ai, 1991, MNRAS (in press)
Kraan-Korteweg, R.G., Sandage, A., and Tammann, GA, 1984, Ap.J. 283, 24
Lucey, J.R., and Carter,D., 1988, MNRAS 235, 1177
Lucey, J.R., Gray, P.M., Careter, D., and Terlevich, R.J., 1991, MNRAS (in press)
Melnick, J., Terlevich, R., and Moles, M.,1988, MNRAS 235, 297
Pierce, M.J., 1990, Ap.J. 344, L57
Pierce, M.J., and Tully, R.B., 1988, Ap.J.
Richter, O.-G., and Huchtmeier, W.K., 1984, A.A. 132, 253
Rowan-Robinson, M., 1985, The Cosmological Distance Ladder, W.H.Freeman & Co
Rowan-Robinson, M., 1988, Space Science Reviews 48, 1
Rowan-Robinson, M., Hughes, J., Vedi, D., and Walker, S.W., 1990, MNRAS 246, 273
Rowan-Robinson, M., et ai, 1990b, MNRAS 247, 1
Rowan-Robinson, M., Hughes, J., Jones, M., Leech, K., Vedi, K., and Walker, D.W.,
1991, MNRAS (in press)
Sandage, A., 1973, Ap.J. 183, 711
Sandage, A., 1988a, Ap.J. 331, 605
Sandage, A., 1988b, Ap.J. 331, 583
Sandage, A., and Tammann, G.A., 1990, Ap.J. 365, 1
Saunders, W. et ai, 1991, Nature 349, 32
Tammann, GA, and Uebundgut, B., 1990, AA 236, 9
Tonry, J.L., and Schneider, D., 1988, A.J. 96, 807
Tonry, J.L., Ajhar, EA, and Luppino, GA, 1989, Ap.J. 346, L57
Tully, R.B., 1989, 5th lAP Meeting Astrophysical Ages and Dating Methods
van den Bergh, S., 1989, Astron.Astrophys.Reviews 1,111
171

DISCUSSION:

Lahav: The dynamical tests which give no = 0.2 - 0.3 are based on optical galaxies, so it
is not surprising that one can get different no's (or biasing parameters) for different tracers.

Rowan-Robinson: I agree. Obviously the interesting question is which perspective,


optical or infrared, gives the least biased view of the matter distribution? My prejudice is
that the infrared view does, so that no ,.... 1. Our models of cluster halos suggest that only
a few percent of the cluster mass resides in the virialized 1Mpc-radius core, so the IRAS
neglect of ellipticals should not be too serious.
Tammann: It is amazing for me that the IRAS data give values of no as low as ,.... 1. A
priori one would have expected randomly high values, because IRAS galaxies are spirals,
that avoid the high density regions (remember the Dressler relation between galaxy type
and density) and therefore map only part of the true dumpiness of the visibly matter.
Decreasing the density contrast must in general lead to an overestimate of no

Pierce: Why are the weights of S.N. Ia and Novae so high? In such a "weighted mean"
approach don't you just get back the result you put in, i.e. isn't it all in the relative weights
of the distance indicators?
Rowan-Robinson: The detailed justification of the weights I use is given in my 1988
review (and in 'Cosmological Distance Ladder'). It is true that the weight a method has in
my solution for Ho depends on the assumed values of (J'zp, the zeropoint error, and of (J', the
dispersion about the calibration relation. The total weight of the Tully- Fisher method is in
fact higher than that of Type Ia supernovae. Novae have very little weight in the estimate
of Ho since they have been measured only in Virgo.

Coles: Why does your definition of X2 contain (log~ -logHo)2, rather than just ((~)­
HO)2?
Rowan-Robinson: Because the distance errors are Gaussian in log-d.
White: Could you comment a little further on the planetary nebula and surface brightness
fluctuation methods? I've been very impressed by the work done on these so far, and their
proponents daim much higher precision distance measures to Virgo (with considerable
observational backup) than the supernova or Tully-Fisher methods you concentrated on
(though there are still problems with calibrators).
Rowan-Robinson: The surface brightness fluctuation method, with a calibration based
on theoretical isochrones, has a lot of potential. I don't understand why the planetary
nebula distances to Leo and Virgo appear to be anomalous.
THE LUMINOSITY-LINE-WIDTH RELATIONS AND THE VALUE OF
Ho

M.J. PIERCE
Dominion Astrophysical Observatory
5071 W. Saanich Rd.
Victoria,
Be vax 4M6

ABSTRACT. An extensive examination of the luminosity-line-width relations as an indicator of


extragalactic distances is described. Evidence for significant background contamination in the Virgo
Cluster is presented. It is shown that this contamination is responsible for much of the controversy
regarding the distance to the Virgo Cluster. The absolute calibration is examined and it appears
that the technique is now limited by the uncertainty in the Galactic distance scale. The relation is
applied to samples from the Virgo and Ursa Major clusters and the value of H. is determined. The
cosmological time-scale test is used to examine the inflationary hypothesis.

1 Introduction

One of the more powerful tests of the inflationary hypothesis is the comparison of cosmolog-
ical time scales, H;l and To. The historical discrepancy ofa factor of two in estimates ofHo
has prevented a meaningful application of the test but recent technical developments have
enabled significant progress to be made. This report describes the results of an extensive
examination of the luminosity-line-width relation (Tully and Fisher 1977; hereafter TFR)
as a distance indicator. The data are taken from a photometric survey of several hundred
nearby galaxies currently underway. Three samples will be discussed: i) a complete sam-
ple of nearby galaxies suitable for providing the absolute calibration of the TFR., and ii)
complete samples from the Ursa Major and Virgo clusters.

2 Absolute Calibration

The calibration sample includes all galaxies suitable for the TFR within the Local, Sculptor,
and M81 Groups. The sample includes 15 galaxies, but unfortunately only 6 currently have
individual distance estimates from Cepheids, RR Lyraes or planetary nebulae. Nevertheless,
the data are sufficient to fully calibrate the TFR. The I-band data are shown in figure 1a.
The dispersion for those systems with known distance (the solid points) is uncertain, but
formally 0.20 mag. A X2 test indicates that the probability· that the true dispersion is as
large as 0.7 mag is :5 1%. More extensive cluster samples have dispersions of'" 0.30 mag
or 15% in distance. Consequently, the absolute calibration of the TFR is probably known
to 0.1 mag or only 5% in distance. However, at this point the uncertainty in the Galactic
distance scale becomes significant (e.g., Feast and Walker 1987).
173
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 173-178.
@ 1991 Kluwer Academic Publishers.
174

3 Virgo and Ursa Major Clusters

Magnitude limited samples suitable for application of the TFR were selected from catalogs
of the Ursa Major (BT :$ 13.5) and Virgo (BT :$ 15.0) Clusters. The I-band TFRs for
the two samples are illustrated in figures 1b,c. The former has a dispersion of only 0.30
mag (0.25 mag after correction for the expected line-of-sight depth), while the latter has a
dispersion of NO.7 mag. Kraan-Korteweg et al. (1988) and Sandage (1988) have suggested
that the Virgo sample is representative of the true TFR dispersion and that the Ursa Major
sample is "biased". On the other hand, Pierce and Tully (1988) have suggested on the basis
of a more limited sample that there may be significant background contamination to the
Virgo Cluster samples, with the larger dispersion being an artifact. This issue is crucial to
the determination of Ho since the relative distance between Virgo and more distant clusters
is in good agreement using a variety of distance indicators. Consequently, the determination
of Ho hinges upon the distance to the Virgo Cluster.

Figure 2 illustrates the individual TFR distances vs. the line-of-sight velocities. The top
panel illustrates the Kraan-Korteweg et al. (1988) data with their calibration. The lower
panel illustrates the new data of Pierce and Tully (1991) with the calibration described
above. The vertical dashed lines correspond to the ± 20' limits if the dispersion of the TFR
is the same as that found for the Ursa Major data and assuming a mean distance given
by the brighter spirals (BT :$ 12.0). The two curves illustrate the line-of-sight velocities
expected for a simple Virgocentric retardation of the Hubble flow due to the mass of the
cluster. It is clear from both panels that the vertical lines encompas the distance range with
the highest velocity dispersion and hence likely indicate the cluster core. While the systems
with TFR distances beyond 30 Mpc are generally agreed to be in the background (they were
not considered by Kraan-Korteweg et al.) the issue of background contamination hinges
upon the systems with D 2: 20 Mpc and V :$ 1600 km S-I. With no additional information
the case for significant background contamination would be rather weak, with the distance
of the cluster being 15.5 to 20 Mpc, depending on whether these systems are included or not.
However, figure 3a illustrates the location of these "background candidates" projected upon
the sky. They are clearly isolated both spacially and kinematically. Figure 3b illustrates
the projected density distribution of galaxies with 500 :::; V :::; 1500 km s-1 from Huchra
(1985). It is clear that the systems in question form a distinct subcluster. This group is
known as the S' or B Cloud and there has been speculation for sometime as to whether it
is at the same distance as the core of the cluster (e.g., de Vaucouleurs and Corwin 1986).
Apparently, the true dispersion in the TFR is rather small and that significant line-of-sight
depth effects are present in the spiral population of the Virgo Cluster. The data presented
here clearly demonstrate that the B Cloud is not at the same distance of the "Virgo Core" ,
and that taking a mean distance for the entire sample (e.g., Kraan-Korteweg et ai.) would
significantly overestimate the distance to the cluster and hence underestimate the value of
Ho. It is also clear that the dispersion in the TFR must be small, on the order of 0.3 mag.
175

4 The Value of Ho

Eliminating the background members of the Virgo Cluster yields a mean distance of 15.6
± 1 Mpc, with the primary uncertainty being that of the Galactic distance scale. If we
take the relative distance between the Virgo and Coma Clusters at om = 3.68 ± 0.10 mag
(Aaronson et al. 1986; Dressler et al. 1987) we can obtain an estimate for Ho which avoids
most of the uncertainty in the local velocity field, the result being:
Ho = (7175 ± 150) km s-1/(85 ± 10 Mpc) = 84 ± 10 km S-1 Mpct,
where the primary uncertainty is again in the Galactic distance scale calibration.

5 The HoTo Time-scale Test

Figure 4 illustrates a convenient depiction of the time-scale test. The product HoTo is
shown against A (the cosmological constant) with various cosmological models specified by
n. If we take To = 16 ± 3.5 Gyrs as representative of current estimates from globular
clusters and with a 2<1 uncertainty for the stellar evolutionary models (see Demarque in
this volume), then the hatshed region corresponds to the 2<1 allowed range for HoTo (1.4 ±
0.3). It is obvious from the diagram that even the n = 0 model (Ho To = 1.0) is in difficulty
with A = O. The classical inflationary model (A =0, n = 1.0, and Ho To = 2/3) is excluded
to ~ 4<1. These results clearly favor a non-zero cosmological constant. In this case, the
inflationary model is given by the k = 0 curve. It is interesting to note the estimated value
for HoTo and the k = 0 model imply a value of n rv 0.1, just what is inferred from the
peculiar motions of galaxies. In other words, a non-zero cosmological constant allows an
inflationary model that is consistant with the estimated cosmological parameters, without
appealing to an additional large-scale distribution of dark matter.

Acknowledgements. Much of this work was done in collaboration with B. Thlly.


176

References

Aaronson, M., Bothun, G., Mould, J., Huchra, J., Schommer, R.A., & Cornell, M.E., 1986. Astro-
phys. J., 302, 536.
de Vaucouleurs, G., & Corwin, H.G., 1986. Astr. J., 92, 722.
Dressler, A., Lynden-Bell, D., Burstein, D., Davies, R.L., Faber, S.M., Terlevich, R.J., & Wegner,
G., 1987. Astrophys. J., 313, 47.
Feast, M.W., & Walker, A.R., 1987. Ann. Rev. Astr. Astrophys., 25, 345.
Huchra, J., 1985. in The Viryo Cluster, p. 181, ed. O. Richter & B. Binggeli, Garching:ESO.
Kraan-Korteweg, R.C., Cameron, L.M., & Tammann, G.A., 1988. Astrophys. J., 331, 620.
Pierce, M.J., 1991. in preparation.
Pierce, M.J., & Tully, R.B., 1988. Astrophys. J., 330, 579.
Pierce, M.J., & Tully, R.B., 1991. Astrophys. J., submitted.
Sandage, A., 1988. Astrophys. J., 331, 605.
Tully, R.B., & Fisher, R., 1977. Astr. Astrophys., 54, 661.
177

. -...-
..
'. :
.'
....,.: ....:: 10

.- :..
:~ ~! ..6C1o
r· . :
. . :;'; 12

..: .....
14

-16

16

1.6 2.0 2.2 2.4 2.6 2.6 3.0 1.6 2.0 2.2 2.4 2.6 2.6 3.0 2.0 2.2 2.4 2.6 2.6 3.0

Figure 1 log W~

Figure 2

Distance (Mpc)

Figure 3 Figure 4

15·
2 fr-O.60
0 .. 0.30
0-0.15
1 ~:g:~:=!~97~~~
~ 10'
C

-1

5· -2

~::::::,:dGJoa . p
-3

(20.0 < I),. < 3O.0.lfpc end V." 1800 1<", s")

o· LL..L..L.L..L..u....w-"-,--LL..LLL.w...w
13 h 12.8 12_6 12.4 12_2 12~ -'~~~~~~~~~~__L-~
RA 1.2 1.0 0.8 0.6 0.4 0.2
Hole
178

DISCUSSION:

Hanes: (comment to Pierce and Rowan-Robinson) I have strong reservations about the
weighting of various estimates of distance scales (or Ho) as described by Mike Rowan-
Robinson.

(1) The errors (0') quoted by various authors are not formal statistical estimates in gen-
eral, and have not the same meaning from author to author. In particular, they
represent merely the precision with which repeated calculations on hypothetical new
data would reproduce the derived results, but they usually provide no estimate of
the sensitivity of the results to underlying systematic working assumptions. In that
sense, the errors are almost invariably optimistic, some more so than others.

(2) Moreover, if there are systematic differences - as we all agree there are - from one
technique to another, then the weighting is not only irrelevant but also dangerous
because it will overweight certain techniques of high claimed precision which may be
systematically in error.

(3) Finally, the 0' estimates are, in any event, not independent. Many of the techniques
are inter-related at some very deep level. For instance, a change in the Hyades mod-
ulus would be felt immediately (but perhaps not with the same power) by many
different techniques. One really needs to consider every source of uncertainty in cal-
ibrations to see how the results are variously affected. The quoted 0' values do not
reflect this.
OBSERVATIONAL STATUS OF Ho

G.A.TAMMANN
Astronomisches Institut
Venusstrasse 7, CH-4I02 Binningen
and
European Southern Observatory
Karl-S chwarzschild-Str.2, D-8046 Garching

ABSTRACf. The distance of the Virgo cluster (21.2 ± 0.6 Mpc) from seven independent
methods is presently the best known extragalactic distance. On the other hand the cosmic
recession velocity of the cluster (v = 1144 ± 18 Ian sol) within a nearly Machian frame can
be determined using relative distances to more distant clusters. The resulting value of the
Hubble constant - freed from all velocity perturbations - is H., = 52 Ian solMpc ol . The for-
mal error of this determination is very small.

1. Introduction
Progress in cosmology comes slowly and quietly. Until recently the nature of galaxian
redshifts could only be deduced from good physical intuition. However, there was no
proof of their being caused by the expansion of the universe. But now there is direct evi-
dence to this effect. The surface brightness of elliptical galaxies within a metric diameter
does apparently decrease with (1 + Z)4 as predicted for an expanding universe, and con-
trary to tired-light and other non-physical models (Sandage and Perelmuter 1991). Also
the typical time delation of expanding models is at the verge of being detected in the light
curves of supernovae of type Ia (Leibundgut 1990).
In the light of such progress the exact value ofH., seems like a minor problem. But the
distinction between H., = 50 [Ian solMpcol ] and - as occasionally still propagated - 100 has
yet important implications for cosmology. If H., "" 100, it is not possible that the age of the
oldest globular clusters be > 14 gigayears (Sandage and Cacciari 1990) and that A = O.
Also no (baryon) < 0.3 (50/Ho)2 from the primordial synthesis of the light elements
(Kurki-Suonio et al. 1990) would become so low with H., = 100, that it would not suffice
to provide the virial mass to bind clusters of galaxies, i.e. no'" 0.14 independent of Ho
(Tammann 1982). They would therefore have to be bound, at least in part, by non-
baryonic matter. Ho = 100 therefore requires non-baryonic matter.
The determination of H., is governed by two effects: 1) streaming motions of galaxies,
and 2) selection effects of flux-limited catalogues containing nearby (calibrating!) galaxies
that are necessarily less luminous than their more distant "counterparts". The former effect
could be remedied at large distances where peculiar motions must be negligible as com-
pared to the Hubble flow, but the latter effect provides at best unbiased distances for near-
by objects.

179
T. Shanks et al. (eds.), Observational Tests oj Cosmological Inflation, 179-185.
© 1991 Kluwer Academic Publishers.
180

The importance of the selection effect depends on the intrinsic magnitude scatter of the
distance indicators used. Optimists take this scatter and hence the selection effect to be
negligible; they obtain necessarily high values of H... It is vice versa for pessimists. They
can argue correctly that the true scatter is always larger than the observed one, because
flux-limited catalogues of field galaxies conceal an important fraction of the underlumi-
nous objects, giving the false impression of small luminosity scatter.
Once one has realized that most available distance indicators (that can be calibrated with
nearby galaxies) have considerable intrinsic luminosity scatter, the determination of H..
from field galaxies becomes a highly complex problem, because the importance of the se-
lection effects depends on the variable space density of galaxies. This is because one
samples into different depths of the luminosity function depending on where high-density
regions are distributed along the line of sight. But the regional densities are not known
without reliable distances. The best way out is to compare the results from complete cata-
logues with increasingly fainter flux limits (Sandage 1988), but this route is very demand-
ing on the observations.
For these reasons it is at present much safer to rely on clusters for the determination of
Ro. At least in principle it is possible to consider complete subsets of the cluster
population. In any case one can sample into different clusters to a constant depth by
including all galaxies brighter than n magnitudes below the first-ranked member. Thus
one obtains comparable galaxy samples and at least reliable relative distances to clusters.

2. Ho from the Virgo Cluster

If selection effects can be controlled particularly well in clusters, it is clear that the distance
of the nearby Virgo cluster must be the ideal mile stone beyond the Local Group. The
problem for H.. is, however, that the cluster is so near that its observed recession velocity
is affected by local streaming motions. The observed cluster velocity is therefore an
unreliable measure of the cosmic expansion.
The problem of the Virgo velocity begins with the observed mean velocity of the cluster.
While Binggeli et al. (1987) determined Vo =976 ± 45 km S·1 (corrected to the barycenter
of the Local Group) from 354 morphological confmned cluster members, the 362 galaxies
in the cluster area with v < 3000 km S·1 gave 1033 ± 38 km S·1 (Huchra 1985). Even if the
latter value is probably still affected by background objects, the uncertainty of the
observed value - including the statistical error - is at the 5-10% level.
As the next step the observed cluster velocity must be corrected for the gravitational pull
of the cluster on the Local Group. The resulting Virgocentric "infall" is VYC = 220 ± 50 km
S·1 (Yahil et al. 1980; Tammann and Sandage 1984). Higher values up to VYC = 470 km S·1
have been propagated, but they are a typical consequence of a gradually compressed
distance scale as caused by selection effects. On the other hand there is now a tendency to
values as low as VYC = 150 ± 50 km S·1 (Tammann et al. 1980; Faber and Burstein 1988;
Sandage and Tammann 1990).
The uncertainty of the Virgocentric infall brings the error of the cosmic Virgo velocity
into the 20% range. This is further increased by any shear motion between the Local
Group and the Virgo cluster as suggested by the microwave background dipole.
The way out of this dilemma is to infer the cosmic Virgo velocity from a number of
more distant clusters whose distances relative to Virgo are known. The relative cluster
distances carry minimum selection effects, and the fractional contribution of any possible
peculiar motions to the observed cluster velocities is small.
181

3. The Distance of the Virgo Cluster


Many authors have contributed to the distance detennination of the Virgo cluster. The
seven independent methods that yield reliable results are tabulated in Table 1.

Table 1. Seven Routes to the Virgo Cluster Distance

Method (m-M)O Gal.type Calibrators Ref.


(1) (2) (3) (4) (5)

Globular clusters 31.62 ± 0.17 E RRLyr 1


Novae 31.57 ± 0.43 E M31 2
Supernovae 31.63 ± 0.25 E, S astron. +physical 3
Dn-o relation 31.85 ± 0.19 SO, S Galaxy, M31, M81 4
21 em line widths 31.60 ± 0.15 S 13 nearby gal. 5
Size of our Galaxy 31.50 ± 0.20 S scale length of Gal. 6
Surface brightn.fluct. 31.61 ± 0.15 E theory of stellar evol. 7

The seven methods in column 1 of Table 1 are self-explanatory. Their resulting distance
moduli and their mean (internal) errors are listed in column 2. The type of galaxy that is
involved in a given method is specified in column 3. The zero-point calibration of the in-
dividual methods is widely different as shown in column 4. It is particularly noteworthy
that four methods out of seven make no use of local calibrating galaxies and their Cepheid
distances. Therefore, if the Cepheid distance to LMC were changed by 0.1 mag. and with
it more or less automatically all other Cepheid distances, the mean cluster distance would
be affected by only 0.04 mag.
Column 5 of Table 1 lists the references to the individual methods:

1) The adopted globular cluster distance is the mean of Harris (1988) and Harris et al.
(1990).
2) The novae in the Virgo cluster are 6.8 ± 0.4 mag. fainter than their counterparts in M31
(Pritchet and van den Bergh 1987). From this Sandage and Tammann (1988) have
found a Virgo modulus as shown in the Table.
3) Various astronomical and physical ways to detennine the distance of supernovae of
type Ia and type II in the Virgo cluster are discussed by Sandage and Tammann (1990);
their adopted value is listed in the Table. The implied relatively high maximum lumi-
nosity of supernovae is in perfect agreement with all workers in the field (cf. Branch
1988).
4) The Dn-o relation was extended to the bulges of SO and spiral galaxies of the Virgo
cluster by Dressler (1987). The zero-point detennination was somewhat improved
(Sandage and Tammann 1988; Tammann 1988).
5) If applied to a complete sample of inclined Virgo spirals the cluster modulus follows
unambiguously from the Tully-Fisher relation; although there are considerable dif-
ferences in the details of the reduction procedures, the results of Kraan-Korteweg et al.
(1988) and Fouque et al. (1990) agree perfectly, except for a small zero-point dif-
182

ference within the stated error. Lower Virgo distances based on "arbitrary" (i.e.lumi-
nosity-biased) subsamples (e.g. Pierce and Tully 1988) should be given zero weight.
Infrared H-magnitudes of some Virgo spiral gave originally the hope that the 21cm-line
widthlluminosity relation had significantly less scatter in this wavelength than in B
(Aaronson et al. 1986); this hope is not borne out as the sample increases (Peletier and
Willner 1991).
6) van der Kruit (1986) has shown that if the distance of the Virgo cluster were less than
20 Mpc, our Galaxy and M31 would have scale lengths larger than any Virgo spiral.
This inadmissible conclusion provides, strictly speaking, only a lower limit to the
Virgo distance.
7) The elegant method of measuring the surface brightness fluctuations in elliptical gala-
xies yielded, when calibrated with the peculiar E galaxy M32, a low Virgo cluster
distance (Tonry et al. 1989). However, M32 revealed peculiarities also in its surface
brightness behavior. If instead a semiempirical calibration, based on the revised Yale
isochrones (Green et al. 1987), is used, it follows the listed value (Tonry et al. 1990).

An additional distance of the Virgo cluster has been derived from the luminosity function
(in the light of the emission line at A. 5007) of planetary nebulae in six E or SO galaxies
(Jacoby et al. 1990). The very low result of (m-M)o = 30.84 ± 0.15 could be averaged
into the seven determinations of Table 1 with only a 7% effect on the mean. However, the
discrepancy is much larger than the confidence intervals. There must be a systematic er-
ror. The problem may lie in the fact that the luminosity function of planetary nebulae was
not calibrated in E/SO galaxies but in the bulges of nearby spirals. If this difference is ac-
companied by small metallicity changes, the masses of the central stars would not be the
same, and this would have a profound effect on the A. 5007 luminosity (Jacoby 1989;
Sandage and Tammann 1990). The latter authors also point out that the bulges of spirals
are filled with dust. Finally the ElSO galaxies in Virgo have a bright tail of the luminosity
function that is missing in the calibrating spirals. Jacoby et al. (1990) explain this as the
effect of overlapping planetaries. This cannot be the case for the brightest object that is
more than 1 mag. brighter than the brightest accepted planetary. The sample of planetary
nebulae must therefore be contaminated by some unknown objects, like e.g. H II regions
(Wagner 1991).
It has also been proposed to use the correlation of the H~ luminosity and the velocity
dispersion of H II regions as a distance indicator, - again with a very small ensuing Virgo
distance (Melnick 1988). The tight correlation is, however, almost certainly an artefact of
objects that are selected with a flux limit. In any case the calibrating nearby H II regions
fall into a different part of the correlation than the giant H II regions in Virgo.
Returning to the seven distance determinations in Table 1, one should note that they all
lie within their quoted error range. In view of the completely different methods, different
zero-point calibration, and different Hubble types involved, this agreement speaks for it-
self. Tests for a foregroundlbackground effect, which would be expected to be largest
between the two main subclusters A and B, reveal no extension in depth beyond the naive
assumption of sphericity (Binggeli et al. 1987). For these reasons the best cluster distance
should be obtained from the values in Table 1.
A weighted mean of the seven distance moduli in Table 1 gives (m-M)o = 31.62 ± 0.06,
where zero front absorption is assumed. Any possible absorption must be smaller than
0.1 mag., which would lower the distance by 5% at most. The distance modulus corre-
sponds to a linear distance of 21.2 ± 0.6 Mpc. The formal error of this value is fortu-
itously small, but it is probable to mean that the distance error is less than 10%.
183

4. The Unperturbed Velocity of the Virgo Cluster


The difficulty of observing the mean Virgo cluster velocity has been discussed in Section
2. The way out is to infer the unperturbed Virgo velocity from more distant clusters
whose relative distance to the Virgo cluster is known. If some cluster has the mean veloc-
ity Vi> which is composed of the Hubble velocity of the cluster VH, the peculiar velocity of
the cluster vPOC' and an observational error Vern and if this cluster is known to be fi times
more distant than Virgo, then the Virgo velocity is vvugo =vgffi + v~fi + verlfi =vjfi. The
pleasant feature of the inferred Virgo velocity is that its spurious terms Vpoc and Ven are
also reduced by the factor f;. Moreover, if the procedure is repeated for many clusters, the
terms Vpoc and Verr will average out. For a cluster with Vi - 6000 kIn S·1 the contribution
from vPOC' which is from present evidence < 500 km S·I, amounts anyhow to less than 8
percent, and the contribution of Verr is considerably less for a well observed cluster.
Following this strategy, Sandage and Tammann (1990) have inferred a cosmic Virgo
velocity of Vy' o(cosmic) = 1180 ± 22 kIn S·1 from 17 clusters with particularly good rela-
tive distances (Fig. 1). Using analogously fIrst-ranked cluster galaxies as relative distance
indicators they found Vy~?(cosmic) = 1107 ± 55 kIn S·I. If instead the 51 supernovae of
type Ia with known B anotor pg maxima, nine of which lie in the Virgo cluster, are used
as standard candles, these authors obtained vYirgo(cosmic) = 1037 ± 45 km S·I. (For details
cf. the original reference).
The mean of the three independent methods is vYirgo(cosmic) = 1144 ± 18 kIn S·I. The
error of this inferred velocity is considerably smaller than only the statistical error of the

4.0 ../
.4:'
••
N •• y
o
./
oCJ 3.5 • •
--.J

3.0

Fig.I. - The Hubble diagram of 17 clusters with good relative distances to the Virgo
cluster. The abscissa is the relative distance modulus from the Virgo cluster. The ordinate
is the redshift corrected for a Virgocentric infall of the Local Group of VYC = 220 kIn S·I.
The latter value has negligible effect on the result. The line has slope 0.2 corresponding to
linear expansion. The least-squares line at the Virgo distance .6 (m-M) = 0 yields vYirgo
(cosmic) = 1180 ± 22 kIn S·I.
184

observed cluster velocity, - leave alone the problem of the Virgocentric infall. The inferred
velocity is tied to a more distant frame, which lies in the case of the supernovae typically
at 2500 km S·I and in the case of the 17 clusters at 6000 km S·I. The brightest cluster
galaxies bring their velocity information from an effective distance of even 10000 km S·I.
The inferred Virgo velocity is therefore tied to what must be very nearly the Machian
frame.

5. The Value of Ho
The best value of H.. is obtained by combining the cosmic Virgo velocity of 1144 ± 18 km
S·I with the cluster distance of 21.2 Mpc. The result is H.. = 54 ± 2 km s·IMpc·l.
The systematic error of this determination is determined by the systematic errors of the
input parameters. It is very unlikely that the error of VVirgo(cosmic) is larger than 100 km
S·I. The true Virgo distance lies almost certainly between 16 and 24 Mpc. Therefore the
confidence limits of H.. are Ho = 44 and H.. = 78. For all practical purposes Ho = 50 is
still the recommended approximation.
This value is in perfect agreement with the evidence fromJield galaxies if the selection
effects are accounted for as well as possible (Tammann 1987).

Acknowledgement: Support from the Swiss National Science Foundation is gratefully


acknowledged.

References

Aaronson, M., Bothun, G., Mould, 1, Huchra, J., Schommer, R.A., and Cornell,
M.E. (1986) ApJ. 302,536.
Binggeli, B., Tammann, G.A., and Sandage, A (1987) AJ. 94,251.
Branch, D. (1988) in: The Extragalactic Distance Scale, eds. S. van den Bergh and
C.l Pritchet, p. 146.
Dressler, A (1987) Ap.J. 317,1.
Faber, S.M., and Burstein, D. (1988) in: Large-Scale Motions in the Universe, ed.
V.C. Rubin and G.V. Coyne, p. 115.
Fouque, P., Bottinelli, L., Gouguenheim, L., and Paturel, G. (1990) Ap.J. 349, 1.
Green, E.M., Demarque, P., and King, C.R. (1987) The Revised Yale Isochrones
and Luminosity Functions (New Haven: Yale University Observatory).
Harris, W.E. (1988) in: The Extragalactic Distance Scale, eds. S. van den Bergh
and CJ. Pritchet, p. 231.
Harris, W.E., Allwright, J.W.B., Pritchet, CJ., and van den Bergh, S. (1990),
preprint.
Huchra, J. (1985) in: The Virgo Cluster of Galaxies, eds. O.-G. Richter and M.
Tarenghi, p. 181.
Jacoby, G.H. (1989) Ap.l 339,39.
Jacoby, G.H., Ciardullo, R., and Ford, H.C. (1990) Ap.J. 356, 332.
Kraan-Korteweg, R.C., Cameron, L.M., and Tammann, G.A (1988) Ap.J. 331.
620.
Kurki-Suonio, H., Matzner, R.A., Olive, K.A., and Schramm, D.N. (1990) ApJ.
353,406.
Leibundgut, B. (1990) A.A. 229, 1.
185

Melnick, J. (1988) in: The Extragalactic Distance Scale, eds. S. van den Bergh and
C.J. Pritchet, p. 157.
Peletier, R.F., and Willner, S.P. (1991) Ap.J., submitted.
Pierce, M.J., and Tully, R.B. (1988) Ap.J. 330, 579.
Pritchet, C.J., and van den Bergh, S. (1987) Ap.J. 318, 507.
Sandage, A., and Tammann, G.A. (1988) Ap.J. 328, 1.
Sandage, A. (1988) ApJ. 331, 605.
Sandage, A., and Cacciari, C. (1990) Ap.J. 350, 645.
Sandage, A., and Tammann, G.A. (1990) Ap.J. 365, 1.
Sandage, A., and Perelmuter, J.-M. (1991) Ap.J. 370, 455.
Tammann, G.A. (1982) in: Landolt-Bornstein, Astronomy and Astrophysics, vol.
2c (Berlin: Springer), p. 346.
Tammann, G.A. (1987) in: Observational Cosmology, I.A.U.Symp. No. 124, eds.
A. Hewitt et aI., p. 151.
Tammann, G.A. (1988) in: The Extragalactic Distance Scale, eds. S. van den Bergh
and CJ. Pritchet, p. 282.
Tammann, G.A., and Sandage, A. (1984) Ap.J. 294, 81.
Tammann, G.A., Sandage, A., and Yahil, A. (1980) Physica Scripta 21,630.
Tonry, J.L., Ajhar, E.A., and Luppino, G.A. (1989) Ap.J.(Letters) 346, L57.
Tonry, J.L., Ajhar, E.A., and Luppino, G.A. (1990) preprint.
van der Kruit, P. (1986) A.A. 157,230.
Wagner, S. (1991) private communication.
Yahil, A., Sandage, A., and Tammann, G.A. (1980) in: Physical Cosmology, eds.
R. Balian et al., p. 130.

Discussion

M.J. Pierce: Can you explain why you dismiss the calibration of the luminosity function
of planetary nebulae, i.e. calibrated with the bulge sample from M31 and applied to ellipti-
cal galaxies, but essentially use the same technique applied to the calibration of the Dn-a
relation and of the globular cluster luminosity function, i.e. calibrated with the bulge of
spirals and applied to ellipticals?
GA. Tammann: I cannot offer a definite answer why the planetary nebulae do not work
yet. I have mentioned several problems that they raise. The final answer requires still
more observations, that are on the way. - The Dn-a distance, by the way, involves no E's.
Moreover, Dressler (1987) has tested the hypothesis that spirals and SO's follow the same
relation.

W.J. Sutherland: You showed a Hubble diagram of cluster distances with very small
scatter. Is this the MWB or LG frame, and have you looked for a cos 9MWB dependence?

GA. Tammann: No. The cluster velocities are in the LG frame (corrected for Virgocentric
infall only). Following your suggestion, H. Jerjen and I have looked into the cos 9MWB
dependence after this conference, and indeed we find from the 17 clusters and the super-
novae a - 500 km S-1 signal in the desired direction. This strengthens, of course, our as-
sumption that these objects constitute the Machian frame and that vYirgo(cosmic) reflects
nearly what a fundamental observer would observe.
CALIBRATING CEPHEID SEQUENCES IN NEARBY GALAXIES

N. METCALFE & T. SHANKS


Department of Physics,
University of Durham,
South Road,
Durham DHl 9LE.

ABSTRACT. Using the CCD camera at prime focus of the 2.5m Isaac Newton telescope on La
Palma, we have checked the stellar photographic sequences on which Cepheid photometry was
originally based in four local Tully-Fisher calibrating galaxies, M31, M33, NGC2403 and M81.
Correcting their distance moduli on the basis of our CCD recalibration we find an infra-red Tully-
Fisher relation of MH = -10.3(logV21 - 2.5) - 21.4, with a scatter of O'!'44.

1 Introduction

The absolute calibration of the Tully-Fisher luminosity-linewidth relation relies heavily


upon the Cepheid distances to local galaxies. Most of these distances have depended, until
recently, upon photographic photometry which, in most cases, is several decades old. Here
we have taken the opportunity to use modern CCD detectors to check this photometry, and
hence the reliability of these distances, in four of the most important calibrating galaxies,
M31, M33, NGC2403 and M81 (together with its companion dwarf galaxy Holmberg IX).
Rather than making multi phaseobservations of the Cepheids themselves we have taken the
approach of reobserving the secondary sequences upon which the Cepheid magnitude scales
were based. We then correct these magnitudes on the assumption that the errors in the
secondary sequences also apply to the Cepheids. Our observations were made using the
RCA and GEC CCD detectors at the prime focus of the 2.5m Isaac Newton telescope on
La Palma. Details of the reduction procedure and photometric calibration can be found
in Metcalfe & Shanks {1991}. In general, magnitudes were measured interactively off the
frames using 4" diameter apertures, with an appropriate correction to total magnitudes.
The fields are very crowded, and, except for the very brightest stars, the major source of
error comes from uncertainties in determining the local background level. Because of this
our limiting magnitudes are not dictated by the exposure time, and for errors not to exceed
±O'!'25 we estimate limits of B '" 23'!'5 and V '" 22'!'5. Except where noted, all our distance
moduli are based on simple least-squares fits to the comprehensive set of period-luminosity
relations given by Feast (1988). It should be noted that O'!'2 needs to be added to our
moduli to place them on the scale of Sandage and co-workers. This is entirely due to the
fact that they adopt a different apparent modulus for the LMC from that assumed by Feast.

187
T. Shanks et al. (eds.J, ObservatioTUlI Tests of Cosmological Inflation, 187-192.
© 1991 Kluwer Academic Publishers.
188

2 Results

2.1 M31

The original work on M31 was done by Baade & Swope (1963), who reported JJo =
24.2 on the basis of B and V photometry. Fig.1 shows a comparison between our data
and theirs for'" 100 stars on two CCD frames in their Field IV. The V data shows a

rn .5
M31 Field IV
B-band
.
Se. • rn .5
M31 Field IV
V-band

,.,Ji.
~ ~

.... ..
Ol Ol
I
0 0
Q
u
u .. eu
Iii'
<l -.5 ~ -.5

-1 -1 ~~~~~~~~~~~~L-~
18 18 20 22
B(Baade & Swope) V(Baade & Swope)
Fig. 1 Comparison between our CCD data and Baade & Swope's photoelectric (.)
and photographic data (0) in M31.

slight scale error, which amounts to '" 0'!'2 at V = 22 m • The comparison in B is quite good
for B < 22 m but faintward of this Baade & Swope's magnitudes become much brighter than
ours. That this is a problem with their magnitudes is confirmed by the independent CCD
dataset of Hodge et al. (1988), who have also observed part of Field IV. Because this error
is so large we feel it would not be appropriate to attempt to correct the B magnitudes, and
so we reject all Cepheids with a mean magnitude of B > 22m as unreliable. This reduces
Baade & Swope's original sample to 10. We then correct the V magnitudes of these for the
scale error implied in Fig.l. Fitting to the Feast P-LB, P-Lv and P-L-CBV relations
then gives apparent moduli of 24.S6 ± 0.12,24.62 ± 0.10 and 24.71 ± 0.09 respectively. Un-
like Baade & Swope's results, these are consistent with zero reddening, which would imply
/-Lo = 24.6. However, there is certainly some foreground reddening within our own galaxy
(the IRAS 100/-Lm flux suggests EB-V = 0.04), and so we adopt /-Lo = 24.S ± O.lS. This
value compares well with Freedman (1990) who reports J.lo = 24.4 ± 0.1, and the H-band
data of Welch et al. (1986), which on the Feast scale gives J.lo = 24.4 ± 0.11.

2.2 M33

The Cepheids in M33 were originally measured in B by Hubble (1926). Since then his
magnitude scale has undergone several major reforms, leading to, on the Feast scale,
/-LAB '" 24.9 ± 0.1. We have reobserved sequences around 6 Hubble Cepheids on a field
in common with both Sandage (1983) and Christian & Schommer (1987). Although our
photometry agrees, in the mean, to within O'!'OS with that of Christian & Schommer, and
hence confirms the 0'!'6 error in the Sandage 1983 data, there is a problem. The correc-
tion which we find must be applied to Hubble's data is 0'!'3 different from that quoted
by Christian & Schommer. Fig.2 shows that this is mainly due to a large shift between
189

the correction required on this field and that on the two others they observed in M33.

2.5
2.5
~
:0
2 ..... /"
/ ~
:0
0
go .. ' /
/

• ..... ..
.";"t' /

..
~ 0 0 ~ •. , /
~
~
1.5 ~ 8
:r: ... ~ :r: 1.5 0... /

··~l.:'"
I
Q
U
I
; III
old 0
u U .o.··~{,.
.5 / / 0
IE'
<I
/ IE'
<I .5 /
/
O.
16 17 18 19 20 21 17 18 19 20 21
m p1 (Hubble) m p1 (Hubble)
Fig.2 Hubble's magnitudes in M33 compared with our CCD data and that of
Christian & Schommer: '.' data on our field, '0' data from elsewhere in
M33. The dashed line is our transform, the dotted line is C&S's.

Given our agreement with Christian & Schommer on the one field in common, the most
likely explanation is that variations in Hubble's photometry of a few tenths of a magnitude
occur over the face of M33. We therefore restrict our analysis to those Cepheids which lie
on fields with CCD photometry. The result, for these 11 Cepheids, correcting each field
individually, is P.o = 24.8 ± 0.1. There are two sets of I-band Cepheid observations in M33,
that of Mould (1987) and that of Christian & Schommer. We have combined both sets
and placed the result on the Feast system. This gives P.Al = 24.79 ± 0.09. Combining
the I and B-band moduli suggests that, as with M31, the reddening is very low. If we
account for the foreground reddening (EB-V = 0.03 from the IRAS 100p.m flux), the result
is P.o = 24.7 ± 0.15.

2.3 NGC2403

The Cepheid distance to NGC2403 was first determined by Tammann & Sandage (1968)

NGC2403 Quadrant. II & III

e~
{i}
old
E- ·0
I
Q 0 0
0
~ 0 0
U 0
U
~o?'
0 >'
<I
0
0
0
-1

14 16 18 20 22 14 16 18 20 22
B(Tammann & Sandage) V(Tammann & Sandage)
Fig.3 Comparison between our CCD data and Tammann & Sandage's photoelectric
(0) and photographic (0) data in two quadrants in NGC2403.

from both B and V observations at maximum light. We have reobserved the sequence stars
in their quadrants II and III. The comparison is shown in Fig.3. The scatter is high, presum-
ably reflecting the very crowded nature of NGC2403, but the B-band is brought reasonably
190

into agreement by a 0'?'2 shift. The V data, however, appears to suffer from a much more
severe problem and we therefore discard the V data and concentrate only on the B-band
Cepheid measurements. As Tammann & Sandage only give maximum light data we fit to
the Bmaz ridge-line P-L relation given in Sandage & Tammann (1968), corrected to give
the Feast distance to the LMC. The result, allowing for our correction to the magnitudes,
is J-tAB = 27.71 ± 0.14. The I-band result of Freedman & Madore can be used to estimate
the reddening. On the Feast system, their data give J-tAl = 27.43 ± 0.14. If we believe these
two results then the reddening implied is EB-V '" 0.1, and hence J-to = 27.3 ± 0.2. As with
M31 and M33 the foreground reddening is very low (EB-V = 0.015 from the lRAS 100J-tm
flux), so most of the reddening would have to be internal to NGC2403. This leaves the
problem noted by Tammann & Sandage themselves, namely that their B - V colours were,
on average, 0'!'4 redder than those of comparable galactic Cepheids. Madore (1976) took as
evidence of very high reddening in NGC2403. However, the colours of the faint Tammann
& Sandage sequence stars are much redder than ours, due to the severe scale error in their
V magnitudes. Dividing their Cepheids into bright and faint subsets we find mean colours
of B- V = 0.66±0.12 for B < 22'!'15 and B- V = 1.05±0.04for B > 22'!'15. The formeris
reasonably consistent with galactic Cepheids of similar periods, so most of the discrepancy
in their colours can be explained by the scale error in their faint V magnitudes.

2.4 M81 and Holmberg IX

The distance of M81 is crucial to the Tully-Fisher calibration, but unfortunately there is

HolX & MBI B band Q) HolX & Mal v band


t>D
m 0
~ 0
~
1-" o~
0 0 0
0
«J 0 a 0
~I 0~----------------~~~---4

eu -t'+co ~o
8

~ -1

14 16 18 20 22 14 16 18 20 22
B(Sandage) V(Sandage)
FigA Comparison with Sandage's HoIX and M8l magnitudes: '0' - HolX,
'+' - M8l blue stars, '.' - M8l red stars, '.' - photoelectric.

much dispute as to whether it is at the same distance as NGC2403 or nearly a factor of


two more distant. Only two Cepheids have been measured in M81, by Sandage (1984) in
B, and by Freedman & Madore (1988) in I. Neither has a published sequence, and so
we have checked Sandage's magnitude scale by re-observing some of his brightest blue and
red stars in M81, and by re-observing his sequence on the dwarf companion galaxy, Holm-
berg IX. Both these sets of photometry are on the same scale as his Cepheids magnitudes.
Fig.4 shows the comparison between our measurements and his. No major scale errors
appear to be present. Transforming to the Feast LMC scale, and accounting for a small
offset present in Fig.4, reduces Sandage's distance modulus to J-tAB = 28.5. Freedman &
Madore's 1- band data give j1.AI = 27.64, and at face value might be taken to imply very high
191

reddening. However, these two values cannot be compared directly, as Sandage's modulus
comes from assuming that the two Cepheids are the brightest in M81, and hence lie on
the upper envelope of the Bmax distribution, whereas the Freedman & Madore value comes
from a simple fit to the mean P-LI relation. The true value of the reddening is therefore
almost certainly less than that implied by the above results. Clearly two Cepheids are not
enough to resolve this question, and we correct Sandage's apparent modulus only for the
foreground reddening of EB-V = 0.03 deduced from the IllS measurements. This leads
to 1'0 f"V 28.4.

3 Conclusions

We have identified problems with previous photometry in three of the four galaxies we have
observed, but with our new data we have now established the photometric scales to better
than 15% . It is clear that the main problem with the Cepheids in all four galaxies is in the
value of the reddening, although there are still uncertainties in NGC2403 and M81 due to
the highly crowded nature ofthe fields. Taking 21cm linewidths and H-band magnitudes as
summarized by Sandage & Tammann (1984), our estimates ofthe true distance moduli for
these four galaxies lead to MH = -10.3(logV21 - 2.5) - 21.4, with a dispersion around this
line of ±0'!'44. However, adopting a distance modulus for M81 similar to NGC2403 would
alter the zero-point (for a fixed slope) by f"V 0'!'3, and would reduce the scatter substantially.
As an uncertainty in the distance of one galaxy can affect the local Tully-Fisher calibration
so much it is clear that this calibration is by no means finalised, and claims for intrinsic
scatter < 0'!'25 based on this calibration are somewhat premature.

References

Baade, W. & Swope, H.H., 1963. Astron. J., 68, 437.


Christian, C.A. & Schommer, R.A., 1987. Astron. J., 93, 557.
Feast, M.W., 1988. Observatory, 108, 119.
Freedman, W.L., 1990. Astrophys. J., 335, L35.
Freedman, W.L. & Madore, B.F., 1988. Astrophys. J., 332, L63.
Hodge, H., Lee, M.G. & Mateo, M., 1988. Astrophys. J., 324, 172.
Hubble, E., 1926. AstrophYII. J., 63, 236.
Madore, B.F., 1976. Mon. Not. R. astr. Soc, 177, 157.
Metcalfe,N. & Shanks,T. 1991. Mon. Not. R. astr. Soc, in press.
Mould, J.R., 1987. Pubis astr. Soc. Pacif., 99, 1127.
Sandage, A., 1983. Astron. J., 88, 1108.
Sandage, A., 1984. Astron. J., 89, 621.
Sandage, A. & Tammann, G.A., 1968. Astrophys. J., 151, 531.
Tammann, G.A. & Sandage, A., 1968. AstrophYII. J., 151, 825.
Welch, D.L., McAlary, C.W., McLaren, R.A. & Madore, B.F., 1986. AstrophYII. J., 305, 583.
192

DISCUSSION:

Pierce: Can you comment on why you just recalibrated the old photographic sequences
instead of doing the Cepheid photometry yourself?
Metcalfe: Unfortunately telescope time is limited! By checking the sequences we can
identify where problems with the Cepheid photometry are most likely to cause trouble and
concentrate on these areas. We also now have random phase I-band observations on all of
our fields.

Pierce: I think it is is premature to dismiss the distances and hence the apparent low
dispersion of the Tully-Fisher relation implied from CCD photometry of the full light curves
or random phase I-band photometry until you have photometry of the full light curves
yourself.

Metcalfe: CCDs may be linear but they still suffer from the same crowded field problems
as did the photographic work, and the same systematic biases will be present in the detection
of the Cepheids. As an example of the uncertainties still inherent in the CCD data, the
two published random phase I-band datasets in M33 give apparent moduli which disagree
by O~3.
Rowan-Robinson: You remark that the foreground interstellar extinction is small. There
is evidence in all four galaxies that there is substantial internal extinction (AB > O~5).
Metcalfe: None of the evidence is very convincing. We have repeated the photometry of
some of Humphreys' supergiants in M3l Field IV and find no significant difference between
the colours of the foreground and M3l stars. Also, McCall's HII data in NGC2403 favours
fairly low reddening. It should be remembered as well that there may be a detection bias
towards the least reddened Cepheids.
Peebles: Will the K-band panoramic detectors help with the problems with crowding and
obscuration?
Metcalfe: Yes, conventional aperture photometry simply cannot give reliable results in
such crowded fields.
NEW D - 0' RESULTS FOR COMA ELLIPTICALS

J.R. LUCEyl, R. GUZMAN!, D. CARTER2 & R.J. TERLEVICH2


1 Department of Physics, University of Durham, South Road, Durham DHt SLE.
2 Royal Greenwich Observatory, Madingley Road, Cambridge CBS OEZ.

ABSTRACT. We discuss D - (T results for a new sample of ellipticals in the Coma cluster. Unlike
previous studies, our sample includes many ellipticals that lie well outside the cluster core. These
data have been used to estimate the environmental dependence of the D - (T relation. The zero-point
of the D - (T relation is found to vary by only -8% ± 10% from Coma's core to a projected distance
of ~4h-l Mpc. The projected galaxian density from the core to our outermost ellipticals decreases
by a factor of over 150. Thus over a very wide range of galaxian densities we find no evidence for a
significant zero-point variation in the D - (T relation.

1 Introduction

In the last few years considerable observational work has been undertaken to quantify
the peculiar motions of galaxies in the local universe. For elliptical galaxies most of the
work has been based on the diameter versus velocity dispersion correlation, i.e. the D - 0'
relation (Dressler et al. 1987). This empirical relation allows distances to be estimated
with an uncertainty of ....,20% and has been used to map the nearby non-Hubble motions
(Lynden-Bell et al. 1988, Lucey & Carter 1988, Dressler & Faber 1990, Lucey et al. 1991).
In a rich cluster, gravitational encounters are believed to alter significantly the structure
of the galaxies. Strom and Strom (1978) found that the elliptical galaxies located near the
cores of dense, rich clusters had smaller sizes than those located in both the cluster haloes
and low density clusters. The D - 0' relation may also be affected by environmental factors
(see Kaiser 1988, Silk 1989, Djorgovski, de Carvalho & Han 1988).
Recently Burstein et al. (1990) have presented strong arguments supporting the univer-
sality of the D - 0' distance indicator. In particular, these authors have emphasised that
the distance determinations from the D - 0' and Tully-Fisher relations are in good agree-
ment. Nevertheless some rich clusters do show large discrepancies between their D - 0' and
Tully-Fisher distances, e.g. for the Abell 2634 cluster the D - 0' relation gives a distance
that is 41 ± 9% greater than that derived from the Tully-Fisher relation (see Lucey et al.).
In order to quantify the level of any environmental dependence in the D - 0' relation
we have compared the ellipticals that lie in the halo and core of the Coma cluster. By
studying only one cluster we avoid the difficulty of decoupling peculiar motions from real
environmental differences.

193
T. Shanks et al. (ells.), Observational Tests of Cosmological ltiflation, 193-197.
C) 1991 Kluwer Academic Publishers.
194

2 Observations

We have made new D - (1 observations of Coma cluster ellipticals. As previous work on


Coma's D - (1 relation (Dressler et al. 1987) covered mainly ellipticals that lie near the
cluster centre, our observations concentrated on ellipticals in the outer part of the cluster.
Spectroscopic and photometric observations were made with the 2.5-m Isaac Newton Tele-
scope on La Palma in May/June 1990. (1 and D measurements were derived using standard
techniques and have average errors of 0.024dex and 0.005dex respectively.

3 Results

The D - (1 relation has the form :


logD[arcsec] = slope X log(1[kms- 1] + constant
Using data for several clusters Lynden-Bell et al. derived a value for the slope of 1.2 ± 0.1.
The D - (1 relation for our sample of Coma ellipticals is shown in Fig. 1. MiniInizing the
scatter in log D, we derive a slope of 1.17 ± 0.09. The rms scatter, in log D, about this slope
is 0.078 which is similar to that found for other well-studied clusters.
We have sub-divided our sample into five radial zones of approximately equal numbers
of galaxies to investigate the dependence of the D - (1 relation on distance from the cluster
centre. The D - (1 diagrams for the five zones are also shown in Fig.1. For all radial
zones the form and the zero-point of the D - (1 relations are very siInilar. The innermost
and outermost zones have average distances from Coma's centre of 0.12°(0.14h-1 Mpc) and
3.3°( 4.1h-l Mpc) respectively (h is the Hubble constant in units of 100km s-1 Mpc 1 ).
Whereas the surface density of galaxies decreases by over a factor of 150 between these two
zones, the observed difference in the D - (1 zero-point, in terms of logD, varies by only
-0.032 ± 0.039, i.e. -8% ± 10%. An alternative approach to illustrate this result is shown in
Fig. 2 where we plot the D - (1 zero-point derived from each elliptical against the logarithm
of projected distance from the cluster centre. The best-fitting relationship between these
two variables is :
Zero-point = 1.679(±0.01l) - 0.026(±0.021) logR[degree]
The derived gradient is not statistically significantly different from zero. This conclusion is
robust. Adopting slopes for the D-(1 relation of 1.1 and 1.3 gives gradients of -0.029 ± 0.020
and -0.022 ± 0.020 respectively. Including Dressler et al. 's data for core ellipticals revises
the gradient to -0.024 ± 0.015. Hence there is no statistically significant correlation between
projected distance from the cluster centre and the D - (1 zero-point.
Over the large radial distances considered in our study, Coma is not a well-Inixed system.
The zero-velocity radius of the cluster is 5.2°(Kent & Gunn 1982). TiIning arguments
suggest that galaxies that lie outside half this radius have yet to traverse the cluster core.
In addition, simple dynaInical considerations indicate that galaxies which currently lie in
the cluster core are unlikely to escape out to radii greater than 1°. Therefore in our study
we have compared ellipticals that have undergone very different levels of galaxy-galaxy
interactions. The D - (1 relation appears not to be affected by such interactions and hence
the D - (1 relation is a reliable distance indicator.
195

Our complete somple


109 0., - 1.2 109 a - 1.684 R < 0.28°
Disp - 0.079
I
en

o
b
01
o
o 0 0

0.28" < R < 0.51" 0.51" < R < 0.85"


,......,
I
en o

b
01
o
o

0.85" < R < 2" 2° < R < 5°


,......, ,,
,
I
en , ,,
o ,
oQ/ "
(],'
b
, ,, o
01 ,,
o ,,
,,

0.8 1.2 1.6 0.8 1.2 1.6

log DV [arcsec] log DV [arcsec]

Fig. 1. The first panel shows the D - u relation for our sample of Coma ellipticals. The
remaining panels show the D - u relation for five different radial zones. For each zone the
mean line of slope 1.2 is shown (solid line). For the outer zones, the mean relation found
for the innermost zone is also drawn (dashed line).
196

2
Dressler et at. doto
>
C)

Ol o " 0
.2 1.8 " 00
0 II 0 0 0 0
0" x
§$ I..T

e------
------11 11 -----_________ 0 0 0 0
b _l<_ - - - ill. - -'t::O~ --:~:~: - X- -d:r::~:: 8= ===d: ==::! ==: =.
Ol
0 1.6 x" x" x 0 0 0 0
ox" "q 0 0 0
N
o
1.4

-1.5 -1 -0.5 o 0.5 1


log R [degrees]
Fig. 2. The variation of zero-point, i.e. 1.2 log u - log D, derived from each elliptical as a
function of projected distance from the cluster centre. At the distance of the Coma cluster,
5° is equivalent to 6.3h- 1 Mpc.

References

Burstein, D., Faber, S.M. & Dressler, A., 1990. AlltrophYIl. J., 354, 18.
Djorgovski, S., de Carvalho R. & Han, M.-S., 1988. in Eztragalactic Dilltance Scale, p. 329, ed. van
den Bergh, S. & Pritchet, C.J., A.S.P. Conference Series.
Dressler, A. & Faber, S.M., 1990. AlltrophYIl. J., 354, 13.
Dressler, A., Lynden-Bell, D., Burstein, D., Davies, R.L., Faber, S.M., Terlevich, R.J. & Wegner,
G., 1987. AlltrophYIl. J., 313, 42.
Kaiser, N., 1988., in Large-Scale Strocture and Motions in the Universe, ICTP preprint.
Kent, S.M. & Gunn, J.E., 1982. Alltron. J., 87, 945.
Lucey, J.R. & Carter, D., 1988. Mon. Not. R. Mtr. Soc, 235, 1177.
Lucey, J.R., Gray, P.M, Carter, D. & Terlevich, R.J., 1991. Mon. Not. R. astr. Soc, 248, 804.
Lynden-Bell, D., Faber, S.M., Burstein, D., Davies, R.L., Dressler, A., Terlevich, R.J. & Wegner,
G., 1988. Astrophys. J., 326, 19.
Silk, J., 1989. AstrophYIJ. J., 345, L1.
Strom, S.E. & Strom, K.M., 1978. Astrophys. J. Lett., 225, L99.
197

DISCUSSION:

Peebles: The insensitivity of the properties of elliptical galaxies to environment is striking!


It would be of great interest to see whether A2634 shows a systematical variation in the
Dn - 0' anomaly with distance from the centre of the cluster.
Lucey: I totally agree. A2634 has plenty of ellipticals outside the cluster core waiting to
be observed. All we need is a sympathetic Time Assignment Group.
NOVAE AND THE DISTANCE SCALE

C.J. PRlTCHET
Department of Phync, and A,tronomy,
Univer,ity of Victoria,
PO Boz 3055,
Victoria, BC vaw 3P6,
Canada.

ABSTRACT. Observations of novae in the Virgo cluster by Pritchet and van den Bergh yield a
Hubble constant Ho = 69 ± 14 km s-l Mpc 1 • Some of the problems in this estimate of Ho are
discussed. In the future it should prove possible to push distance determinations using novae out
beyond the Local Supercluster, and hence determine a value of Ho less affected by local velocity
anomalies.

1 Introduction

Novae are excellent distance indicators. They are bright, visible to large distances, and easy
to recognize. They are found predominantly in ellipticals and the bulges of spirals (Ciar-
dullo et ai. 1987); such envirorunents are relatively dust-free and photometrically smooth.
Most importantly, the calibration of novae as standard candles possesses relatively low
intrinsic scatter (van den Bergh and Pritchet 1986), and is well-understood theoretically
(Shara 1981a,b). Furthermore, the observed properties of novae are not strongly affected
by metallicity effects (van den Bergh and Pritchet 1986).
The early history of observations of novae beyond the Milky Way is summarized by
Pfau (1976) and van den Bergh (1988). Pritchet and van den Bergh (1985, 1987a) have
observed novae in Virgo cluster ellipticals; these observations represent the only systematic
studies of novae at distances large enough to determine Ho.

2 Distance Determinations Using Novae

The relationship between the maxiImlm luminosity of a nova and its rate of decline in
(MMRD), first noted by Zwicky (1936), is the usual starting point for deriving extragalactic
distances. The most complete recent study of the MMRD relation for Galactic novae is by
Cohen and Rosenthal (1983) and Cohen (1985), who have derived expansion parallaxes for
a number of (previously undetected) spatially resolved shells around old Galactic novae.
The mean scatter about a linear fit to the high quality subset of Cohen's data is ±O.52 mag
(117).
An alternate calibration of the MMRD relation is obtained by studying novae in the
nearby spiral galaxy M31 (Ciardullo et ai. 1987, Capaccioli et ai. 1989). The 117 scatter
199
T. Shanks et al. (eds.). Observational Tests o/Cosmological Inflation. 199-203.
© 1991 Kluwer Academic Publishers.
200

around the mean M31 relation is in the range 0.20 - 0.28 mag (c/. Capaccioli et al. 1989,
van den Bergh and Pritchet 1986).
A comparison of the M31 and Galactic MMRD relations shows that: (i) the fastest
novae in M31 are less luminous than their Galactic counterparts (possibly because of poor
sampling of the light curves); and (ii) there appears to be a systematic offset of about
0.3 mag between the two MMRD relations (possibly because of assumptions of spherical
symmetry in the expansion parallax technique - e.g. Ford and Ciardullo 1987). The scatter
in the Galactic MMRD is considerably worse, probably due to large and uncertain Galactic
absorption. In view of all of the above, it seems somewhat safer to employ the M31 MMRD
relation as the calibrator for the extragalactic distance scale. This makes the distance scale
dependent on the distance to M31, which is, however, quite well-determined (e.g. van den
Bergh 1989).
One problem with the M31 calibration is the existence of so-called "anomalous" novae.
These novae, which represent roughly 10 per cent of all novae in M31, are unusually bright
(.:lm 2: 1 mag) for their observed decay rate in; at least one such object has also been ob-
served in the Virgo cluster (Pritchet and van den Bergh 1987a). Luminous variable objects
of unknown type, such as the red "nova-like" variable discovered by Mould et al. (1990) in
the bulge of M31, may also be confused with normal novae in MMRD diagrams of distant
galaxies.

3 Novae in the Virgo Cluster

3.1 THE HUBBLE CONSTANT

The first observations of novae in the Virgo cluster were made photographically by Hubble.
More recently Pritchet and van den Bergh (1985, 1987a) have discovered and followed a
number of novae in giant elliptical galaxies in the Virgo cluster. For 7 novae with reasonably
well-determined in and B_ values, they found that the difference in distance modulus
between M31 and Virgo was 7.0±OA mag. Assuming an apparent distance modulus to M31
of 24.65±0.15 (Pritchet and van den Bergh 1987b), and allowing for a small bias (evaluated
from Monte Carlo simulations), they deduced a distance modulus (m- M)o = 31.45± 0.44
for the Virgo cluster. When combined with a !Lean velocity for Virgo (l1!'i.. ) = 1033 ± 38
km s-1 (Huchra 1985), and a Virgocentric infall of 303±39 km s-1 (Aaronson et al. 1982),
this led to a Hubble constant of 69 ± 14 km s-1 MpC 1.

3.2 PROBLEMS WITH A HUBBLE CONSTANT DERIVED FROM NOVAE

There are a number of problems with these observations, and with the final value of the
Hubble constant that is derived. (These problems apply to other observing programmes
designed to study novae in the Virgo cluster and beyond.) (i) Scheduling. Because bright
novae are so fast, repeated observations are required each night to estimate fflmoz and in.
But the nova rate per CCD field is low, so that large amounts of observing time are re-
quired. (ii) Light Curve IfTe!JUlarity. Nova light curves do not always follow a simple linear
decrease in magnitude. As a result, secondary outbursts can be confused with primary
201

maxima if only a limited time series of observations is available. (iii) Calibration Uncer-
tainties. A variety of uncertainties in the calibration of the MMRD have been discussed
above. Of particular concern for distance determinations in the Virgo cluster and beyond
is the existence and nature of the so-called "anomalous novae". (iv) Velocity Uncertainties.
Sandage and Tammann (1988) have pointed out the effects of uncertainties in the velocity
of the Virgo cluster on the determination of Ho from the Pritchet and van den Bergh data.
This problem can be avoided by observing novae in galaxies that are more distant than the
Virgo cluster.

4 The Future of Novae as Distance Indicators

We describe several large projects below that would result in an improvement in the value
of the Hubble constant derived from novae. The first two would result in an improved
calibration of the MMRD relation, whereas the latter two concern observations of novae at
large distances from the Milky Way.
(i) A New Galactic Calibration. The systematic difference between the Galactic and
M31 MMRD relations needs to be understood better. Following Ford and Ciardullo's (1988)
suggestion, it would be extremely helpful to have a new Galactic calibration based solely
on novae whose shell geometry is well determined.
(ii) A New M31 Calibration. Problems with the existing M31 nova database include:
possible transformation errors from one set of observations to another, systematic photo-
metric errors for novae observed photographically against the bright M31 bulge, sensitivity
to reddening, and poor temporal sampling of nova light curves. A new CCD survey of the
M31 bulge would solve many of these problems. Using the R or I bandpasses would allow
observing at all lunar phases (thus improving the continuity of the observations), lower the
effect of internal reddening in M31 (if any), and possibly reduce the effect of the "reddening
pulse" near maximum light (van den Bergh and Younger 1987).
(iii) Ha Observations. The Ha emission from novae remains bright for a much longer
time than does continuum emission (e.g. Fig. 1 of Ciardullo et al. 1990a), and the contrast
of a nova against the underlying background light is excellent when observed through a
narrowband filter. Unfortunatel y, there does not appear to be an MMRD relation in Ha
light (Ciardullo et al. 1990a). It might nevertheless be possible to use the nova rate as a
luminosity (and hence distance) indicator, since the nova rate per unit per unit luminosity
appears to be constant over a wide range of galaxy types (Ciardullo et al. 1990b). Calcula-
tions show that observations of almost complete samples of novae in the Virgo cluster are
feasible in the 0'!5-0'!6 seeing conditions that are not uncommon on Mauna Kea. For this
technique to be applied, more Ha data on M31 novae would be required.
(iv) More Distant Galazies. Our ultimate goal is to obtain observations of novae well
beyond the Local Supercluster, so as to minimize the effects of local distortions in the
Hubble flow on the determination of Ho. The brightest novae can already be detected in
galaxies as distant as the Coma cluster (B"",,,, ~ +25 if (m - M) ~ 35). To obtain in
values for such novae may be within reach of the next generation of 8m telescopes, or 4m
telescopes with adaptive optics.
202

Acknowledgements. Much of this work was carried out in collaboration with Sidney van den
Bergh, to whom the author is grateful for stimulating discussions. This work was supported in part
by the Natural Sciences and Engineering Research Council of Canada.

References

Aaronson, M., Huchra, J., Mould, J., Schechter, P.L., &; 'fully, R.B. , 1982. Astrophys. J., 258, 64.
Capaccioli, M., Della Valle, M., D'Onofrio, M., &; Rosino, L. , 1989. Astron. J., 97, 1622.
Ciardullo, R., Ford, H.C., Neill, J.D., Jacoby, G.H., &; Shafter, A.W., 1987. AstrophYIJ. J., 318,
520.
Ciardullo, R., Shafter, A.W., Ford, H.C., Neill, J.D., Shara, M.M., &; Tomaney, A.B. , 1990a. AIJ-
trophys. J., 356, 472.
Ciardullo, R., Ford, H.C., Williams, R.E., Tamblyn, P., &; Jacoby, G.H. , 1990b. AlJtron. J., 99,
1079.
Cohen, J.G. , 1985. Astrophys. J., 292, 90.
Cohen, J.G., &; Rosenthal, A.J. , 1983. Astrophys. J., 268, 689.
Ford, H.C., &; Ciardullo, R. , 1988. in The Eztragalactic Distance Scale, p. 128, ed. S. van den
Bergh &; C.J. Pritchet, BYU Press, Provo.
Huchra, J.P., 1985. in The Virgo ClulJter, p. 181, ed. O.G. Richter &; B. Binggeli, ESO, Garching.
Mould, l.R., Cohen, l., Graham, l.R., Hamilton, D., Matthews, K., Picard, A., Reid, N., Schmidt,
M., Soifer, T., Wilson, C., Rich, R.M., &; Gunn, J.E. , 1990. Astrophys. J., 353, L35.
Pfau, W., 1976. Astr. AlJtrophys., 50, 113.
Pritchet, C.l., &; van den Bergh, S., 1985. Astrophys. J., 288, L41.
Pritchet, C.J., &; van den Bergh, S., 1987a. AstrophYIJ. J., 318, 507.
Pritchet, C.J., &; van den Bergh, S., 1987b. A6trophys. J., 316, 517.
Sandage, A.R., &; Tammann, G.A. , 1988. AstrophYIJ. J., 328, 1.
Shara, M.M., 1981a. AstrophYIJ. J., 243, 268.
Shara, M.M., 1981b. Astrophys. J., 243, 926.
van den Bergh, S., 1988. Publ. A8tron. Soc. Pac., 100, 8.
van den Bergh, S., &; Pritchet, C.J. , 1986. Publ. Astron. Soc. Pac., 98, 110.
van den Bergh, S., &; Younger, P.F. , 1987. A3tr. Astrophy•. Suppl., 70, 125.
Zwicky, F., 1936. Publ. Astron. Soc. Pac., 48, 191.
203

DISCUSSION:

Pierce: What size telescope would you need to monitor and recalibrate the M3l novae?
Pritchett: This could be done with a very small telescope - say 0.4 m - equipped with a
widefield CCD.
Hanes: What fraction of the luminosity of M87 is sampled by a CCD at the CFHT prime
focus?
Prichett: About 25% for the 2' x 3' field that we had available.
A HIGH RESOLUTION, GROUND BASED OBSERVATION OF A
VIRGO GALAXY.

T.SHANKS, N. TANVIR, P.DOEL, C.DUNLOP, R.MYERS, J.MAJOR


Department of Physics,
University of Durham,
South Road,
Durham DH1 SLE.

M.REDFERN, N.DEVANEY & P. O'KANE


Department of Ezperimental Physics,
University College Galway,
Ireland.

ABSTRACT. We argue that a very simple inflationary model with nbar!!on = 1 would be allowed, if
Hubble's constant lay below its presently accepted range. As part of a programme to check the value
of Hubble's constant we present new, high resolution observations of the Virgo galaxy, IC3583, using
the technique of image sharpening from the William Herschel Telescope. Although our observed
image of this galaxy has 0."65 seeing the galaxy is not resolved into stars. We suggest that this
preliminary result may imply a lower limit to the Virgo distance of 14Mpc and suggest that future
observations will allow further improvements on this limit.

1 Introduction.

The immediate aim of the work described here is to determine the value of the Hubble
constant. But the overall motivation for this work arises from the attractiveness of the
inflationary cosmological model which removes, in particular, the horizon and flatness prob-
lems in the standard Big Bang Model (see Guth's review, this volume). However, inflation
causes several other problems for observational cosmology because, in its simplest version,
the prediction is that the value of the density parameter should be very close to n = 1.
Shanks (1985) suggested that if the value of the Hubble constant, Ho, lay lower than the
currently accepted range then three particular problems for the n = 1 model would be
alleviated, as summarised below.

205
T. Shanks et al. (eds.), Observational Tests o/Cosmologicallnf/ation, 205-210.
© 1991 Kluwer Academic Publishers.
206

2 Arguments for a low value of Ho.

The first observational problem for the inflationary, n = 1 model ifHo (= 100hkms- 1 Mpc- 1 )
lies in the usual range with 0.5 :s; h :s; 1.0 is that the implied age of the Universe then lies
between 6.5 :s; to :s; 13GYT and this is lower than the currently accepted age of globular
clusters at 16 ± 2Gyr, (see reviews by Demarque and Renzini in this volume). If this age for
globular clusters is correct then an upper limit for no = 1 models is Ho < 40kms- 1 Mpc 1 j
the improved agreement between the universal and stellar ages therefore constitutes the
first motivation for proposing a low Hubble constant in an inflation model.
Secondly, the inflation model's n = 1 prediction is apparently at odds with the limit
on the baryon density from primordial nucleosynthesis of Helium-4. This suggests that
0.01 < n o h 2 < 0.06 (eg Turner 1984); if h > 0.5 then this implies that nbaryon < 0.24 and
this is often taken to imply that the density of the Universe is dominated by non-baryonic
particles. However, as noted by Peebles(1984), this produces a new fine-tuning problem
in the inflation model because the baryonic and exotic particle densities are, to order of
magnitude, equal. Very few exotic particle models leave this as any more than an awkward
coincidence. In this view, the inflation model has removed the flatness problem only for it
to be replaced by another fine-tuning problem, the 'baryon flatness' problem. However, if
h '" 0.25 then values of nbaryon "" 1 are allowed and the original simplicity of the inflation
model is restored, with no further fine-tuning problem introduced.
Thirdly, a low value of Ho also helps the missing mass problem in the place where it
is most severe, in the centre of the Coma galaxy cluster. The ratio of this cluster's virial
mass to its X-ray gas mass, under the thermal bremsstrahlung model, within a radius of
1h- 1 Mpc , is Mvirial/Mgas = 15h-1.5. Thus if 0.5 :s; h:S; 1.0 then the virial mass is a factor
of 5 - 15x bigger than the gas mass. But if h "" 0.25 then the X-ray gas mass is within
a factor of 2 of the virial mass and, considering the errors, in good agreement. So with a
low value of Ho, not only is n = 1 in baryons allowed but there is a very plausible baryonic
candidate for the Coma missing mass in terms of ionised gas.
A further advantage is that with n = 1, baryonic models with adiabatic perturbations
have much less of a problem with microwave background isotropy than similar, open models
(eg Wilson & Silk 1981). Adiabatic perturbations arise more naturally from inflation than
any other type of perturbation. However, it must be conceded that the smoothness of
the microwave background is still the point where the model is most vulnerable to future
observations. The scale invariant model predicts 6T IT = 2 - 4 X 10- 5 on scales of 7' for the
OVRO experiment of Readhead et al (1989) compared to the observed 6T IT :s; 2.1 X 10- 5 .
Here the predicted range is taken from Holtzman(1989) and Suto et al(1990) with both
results scaled to a correlation function scale length of TO = 2.5h- 1 Mpc, thus assuming a
'bias' factor of b '" 2 so that the n = 1 model does not contradict cosmic virial theorem
results (e.g. Hale-Sutton et al, 1989). A similar procedure predicts 6T IT "" 1 X 10- 4 for
the South Pole experiment of Meinhold & Lubin(1991) compared to the observed 6T IT :s;
4 X 10- 5 • Further, Holtzman(1989) shows that this baryonic model predicts a quadrupole
anisotropy consistent with the observed upper limit of 6T IT :s; 3 X 10-5 from COBE
(Smoot, this volume). Thus, given the uncertainties in normalising these predictions, the
n = 1, n=1, baryonic model cannot be very confidently ruled out by the current limits on
the microwave background anisotropy. If there is no detection of anisotropy at small and
207

intermediate scales in more sensitive experiments, then the baryonic model might still be
rescued by appealing to an n > 1 primordial spectrum which lowers the predicted 6T/T,
although requiring a more complicated inflationary scenario.
A low value of the Ho also means that the protocluster scale length produced by the
adiabatic model falls within the observed galaxy clustering coherence length, making galaxy
formation in collapsing clusters more plausible than previously. Finally, the cluster collapse
provides a more natural 'biasing' mechanism than is available in 'bottom-up' models of
galaxy formation; if galaxies are more easily formed at the centre of the forming cluster
than elsewhere then the mass may be left more widely distributed than the galaxies on
scales of several Megaparsecs.
Therefore as part of a programme to check the value of Ho we next describe an attempt
to resolve a Virgo cluster galaxy into stars and thereby provide a new estimate of this
cluster's distance.

3 High Resolution, Ground Based Observations of a Virgo Cluster Galaxy.

Virgo remains a crucial step on the extragalactic distance ladder. Sandage & Tammann(1990)
among others have shown that a reliable Virgo-Coma distance modulus is now available from
a variety of distance estimation methods applicable to galaxies in clusters. This route to Ho
does not use the uncertain Hubble velocity of the Virgo cluster. Sandage & Tammann(1990)
find Ho = 52(±2)(21.9/ DVirgo)kms- 1 Mpc- 1 • However, the distance to Virgo, DVirgo, is
very uncertain. Distance estimates have been given in the range 11 ~ DVirgo ~ 23Mpc to
give values of Ho in the usual 50 ~ Ho ~ 100kms- 1 Mpc- 1 range.
The original aim of the Hubble Space Telescope was to resolve Virgo galaxies and
use Cepheids and the Hertzsprung-Russell diagram to provide powerful estimates of this
cluster's distance. Over the past 2 years, we have been investigating the potential of
ground-based 'image sharpening' as an alternative approach to the Virgo problem. The
technique, due to Fried(1978), works by using telescope sub-apertures which are matched
to the 'seeing cell' size in the atmosphere. Essentially this means that much of the time
it is dominated by only a few 'speckles'. The movement of the image in the individual
apertures is monitored using a bright foreground star; a bright star is required so that the
motion can be tracked down to timescales of < 10ms. The upper bound to the resolution
of the final image is set by the diffraction limit of the sub-aperture. In initial 1" seeing,
the typical matched sub aperture size is O.4m giving a diffraction limit of 0."25 in the B
band. This technique forms the basis for the MARTINI instrument of Doel et al(1990).
This uses 6 sub-apertures with on-line corrections for atmospheric motions and images to
a CCD detector. We have also used a complementary off-line approach in this work using
the Galway Image Sharpening Camera of Redfern et al(1989).
In bright time observations at the William Herschel Telescope (WHT) the image sharp-
ening technique has produced some excellent test results on bright star systems, regularly
achieving sub - 0."5 images. In GISC observations of the ADS6650 triple star system we
improved the seeing from 0."6 to 0."34 FWHM using the primary star to resolve the 0."6
double at 6" distance (see Fig.l). The best seeing improvement so far achieved by MAR-
TINI in an off-line experiment was from 0."4 to 0."24; in on-line experiments it has achieved
208

250 250

200 :e
. . ZOO

150 150

• e
100 100

50 50

0 0
0 50 160 I~ 260 2~ 0 50 160 I~ 260 250

Figure 1. The triple star ADS6650 before and after image sharpening.
The single star is used as the monitor star to sharpen the o. "6 double
star at 6" distance. The seeing improvement is from o. "6 to O. "34
FWHM.

.
.,
-.
\.,',. '\

Figure 2. The WHT MARTINI image of the Virgo galaxy, IC3583.


The resolution is o. "65 FWHM and the limit for stellar detections is
R=22.3mag.
209

0."36 from an initial 0."54. Of course, Virgo is an ideal target for image sharpening be-
cause there are so many galaxies that it is easy to find examples with an overlapping, bright
foreground monitor star.
We now argue that 0."3 seeing is enough to resolve Virgo galaxies, if the cluster lies in
its currently accepted range. The argument is based on the galaxy M101 which is the most
distant galaxy with identified Cepheids (Cook et al, 1986). The Cepheid distance modulus
implies a distance to M101 of 7.2Mpc. Now M101 is clearly resolved in the Kitt Peak CCD
images of Cook et al, with median 1."2 seeing. Hence in 0."3 seeing, a similarly crowded
Virgo galaxy should be resolved even if Virgo lies at 29Mpc. On the other hand, if Virgo
lies only at llMpc then a similar galaxy to M101 should be resolved in 0."8 seeing. Thus
the Cepheid distance to M101 seems to cause problems for the Ho = 100kms- 1 M pc- 1 even
in relatively low resolution images of Virgo galaxies. This is part of the Sandage-Tammann
case against a short Virgo distance. The conclusion is that in 0."3 seeing Virgo will be
resolvable if its distance is less than 29Mpc.
In February, 1990, we obtained an image of the Virgo galaxy, IC3583, using the above
technique. Unfortunately the weather and seeing were poor during the dark nights. Also
due to an instrumental problem, we lost a factor of 3 in the signal-to-noise of the CCD
image. Thus the resulting 1.5hr exposure has 0."65 seeing and a limit of R = 22.m3 for
stellar detections. The image is shown in Fig. 2. If Virgo was at a distance less than 14Mpc
then on the basis of the M101 observation the galaxy should be resolved. At our detection
limit we should be probing up to '" 1m into the stellar luminosity function but although
the galaxy 'fuzz' is detected the brightest stars are not resolved. Thus our preliminary
conclusion is that this observation is consistent with the view that Virgo is at a larger
distance than 14Mpc. New observations are currently in hand to take us to our 0."3 limit
and produce yet stronger constraints on the Virgo distance.

4 Conclusions.

We have discussed the advantages for cosmology if the true value of the Hubble Constant
was smaller than currently believed. Firstly, the age of the Universe in an n=1 in:H.ation
model would be smaller than the age of the globular clusters. Secondly, a simple baryon
dominated model with n = 1 is then more consistent with nucleosynthesis arguments based
on the observed abundance of Helium. Thirdly, the hot X-ray gas increasingly becomes a
likely solution to the missing mass problem in rich galaxy clusters. As part of a programme
to make a new determination of Ho we have shown that the image sharpening technique
can frequently allow 0."3 resolution images to be obtained from ground-based telescopes.
A MARTINI instrument observation of the Virgo galaxy, IC3583, in 0."65 seeing has failed
to resolve that galaxy into stars. From our preliminary result, we suggest that this is
unexpected if Virgo is at a distance less than 14Mpc, assuming a distance of 7Mpc for
MID!.

Acknowledgements. TS acknowledges the support of a Royal Society 1983 University Fellowship.


210

References

Cook, K.H., Aaronson, M. &; Illingworth,G., 1986. Astrophys.J., 301, L45.


Davies, R.D., Lasenby, A.N., Watson, R.A., Daintree, E.J.,Hopkins, J., Beckman, J., Sanchez-
Almeida, J., &; Rebolo, R., 1987. Nature, 326,462.
Fried, D.L., 1978. J. Opt. Soc. Am., 68, 1651.
Doel, A.P., Dunlop, C.N., Major, J.V., Myers, R.M., Purvis, A. &; Thompson, M.G., 1990. Proc.
S.P.I.E., 19, 1236.
Hale-Sutton, D., Fong, R., Metcalfe, N. &; Shanks,T., 1989. Mon. Not. R. astr. Soc, 231, 569.
Holtzman, J.A., 1989. Astrophys. J. Supp., 11, 1.
Meinhold, P.R.,&; Lubin, P.M., 1991. Astrophys.J., 310, Lll.
Melchiorri, F., Melchiorri, B.O., Ceccaralli, &; Pietranera, L., 1981. Astrophys. J., 250, L1.
Peebles, P.J.E., 1984. Astrophys.J., 284, 439.
Readhead, A.C.S., Lawrence, C.R., Myers, S.T.,Sargent, W.L.W, Hardebeck, H.E.,&; Moffet, A.T.,
1989. Astrophys. J., 346, 566.
Redfern, R.M., Devaney, M.N., O'Kane, P., Ramirez, E.B., Renasco, R.G., &; Rosa, F., 1989. Mon.
Not. R. astr. Soc, 238, 791.
Sandage, A.R., &; Tammann, G.A., 1990. Astrophys.J., 365, 1.
Shanks, T., 1985. Vistas in Astr., 28, 595.
Suto, Y., Gouda, N., &; Sugiyama, N., Astrophys. J. Supp .. 1990,14,665.
Turner, M.S., 1984. in Eleventh Tezas Symposium on Relativistic Astrophysics, p. 106, ed. Evans,
D.S., New York Academy of Sciences.
Wilson, M.L., &; Silk, J., 1981. Astrophys. J., 243, 14.

DISCUSSION:

W.J. Sutherland: Can't you avoid the crowding problem in finding Cepheids by using
difference frames at many epochs?
T. Shanks: This may help in finding them but not in measuring their magnitudes and
light curves.

M.J. Pierce: If Virgo is really at 40Mpc (Ho '" 25 - 30) then don't you run into more
fundamental problems like LMC-type objects being comparable in size to the Milky Way
and M3l?
T. Shanks: No. As van der Kruit has pointed out, small Virgo distances imply that
the Milky Way and M3l are brighter than the brightest Virgo spirals. Increasing the Virgo
distance by a factor of 2-3 will leave these galaxies only a magnitude fainter than the Virgo
M* and therefore comparable to typical Virgo spirals.
GLOBULAR CLUSTERS AS EXTRAGALACTIC DISTANCE
INDICATORS

DAVID A. HANES
Department of Physics, Queen's University
Kingston, Canada K7L 3N6

ABSTRACT. In this brief review, I will describe some recent developments in the use of
globular clusters as extragalactic distance indicators. My principal message will be by way
of a warning: globular cluster systems (GCSs) differ in perplexing ways in various galaxies
and environments, and caution may be needed in their use as distance indicators, despite
their so far exemplary behaviour. They must, however, be telling us exciting things about
the formation, evolution and interaction of galaxies, and the time may be nearly ripe for
the development of a clear interpretation.

1. Introduction

In giving talks on this subject some ten years ago (Hanes 1980), I enthusiastically endorsed
the use of globular clusters as extragalactic distance indicators. Among their attractions
were the following features:
(i) globular clusters are intrinsically very luminous and consequently far-reaching;
indeed they have even been detected in the galaxies of the Coma cluster (Harris 1987;
Thompson and Valdes 1987);
(ii) they are non-variable, and their study in external galaxies requires little more than
"snapshots" - there is no need for the long-term photometric monitoring required for
Cepheids, for example;
(iii) they are found in the halos of galaxies, in photometrically smooth backgrounds
free of much extinction;
(iv) they provide salutary checks on the more usual distance scales which rely on
indicators of Population I;
and
(v) they are numerous in many galaxies, where they have lately become well-studied,
thanks in part to the advent of telescopes such as the CFHT in sites offering superb
seeing.
To be weighed against those benefits, of course, are the obvious problems:
(i) individual globular clusters span a vast range of luminosity, and their use as dis-
tance indicators requires us to consider the full luminosity functions (LFs) in method-
ologically sound fashion (Hanes and Whittaker 1987);

211
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 211-215.
© 1991 Kluwer Academic Publishers.
212

but
(li) the local (calibrating) samples are those found in our own galaxy and M31 -
both spirals - while the most populous and best-studied external systems are those
in ellipticals. There is no assurance that the LFs in such manifestly different systems
will be closely alike at birth; moreover, even if they are,
(iii) not all globular clusters are very robust, and the mechanisms which may shred
them are unlikely to be identical in galaxies of different types (Fall and Rees 1977).
Even if the LFs were identical at the time of the formation of the GCSs, then, the obvious
danger is that they might have evolved over time into something quite different. Indeed, as
Harris (1991) has emphasized, it is remarkable that to first order the LFs clearly are very
similar: the adoption of any reasonable distance scale implies nearly identical turnover
magnitudes in the GCSs studied so far. But their continued use as far-reaching distance
indicators demands a clear understanding of second-order differences if we are to avoid
potentially troublesome scale errors.

2. Some Recent Developments: Continuing Good News


We have recently carried out analyses of the globular cluster LFs in a couple of galaxies
of special interest: NGC 1399, the central giant elliptical in the Fornax cluster (Bridges,
Hanes and Harris, 1991); and NGC 4594 (M104), the well-known Sa "Sombrero" galaxy
(Bridges and Hanes, 1991). What can we conclude?
(i) NGC 1399. This galaxy is of special interest in that it lies nearly antipodal to
the Virgo cluster, and indeed NGC 1399 is very like M87 in appearance, central
cluster location, and superfluity of globular cluster population. This circumstance
allows a scale-independent determination of the relative distance moduli of the Fornax
and Virgo clusters through a straightforward intercomparison of the GCLFs, and a
consequent estimate of the Virgocentric motion of the Local Group. We find, in
agreement with many others studies, that the Fornax cluster is closer than the Virgo
cluster by about 10%; an infall velocity of -200 - 300 km/sec is implied.
(ii) NGC 4594. In this large-bulged spiral, with a rich GCS (Harris, Harris and Harris
1984), we have for the first time reached the turnover in the GCLF at B-24 (Bridges
and Hanes, 1991). The galaxy lies closer than the Virgo cluster by about 0.7 mag in
distance modulus. The GCLF is not sensibly different from its appearance in Virgo
ellipticals, although this constraint is a mild one since only the bright half of the
function is plumbed.

3. Some Recent Developments: The Worrying News


We have now explored a large part of the "parameter space" - that is, GCSs have been
studied in galaxies of different types in varied environments (see Harris 1991 for the most
recent review) - and we can consider the evidence for important differences among them.
Unfortunately, such differences are disquietingly many. Among them are the following:
(i) Colour gradients are seen in some but not all GCSs. For example, the globular
clusters in NGC 1399 are redder by about 0.1 mag in (B - V) at 1 armin radius than
213

those at 3-4 arcmin. Gradients of comparable amplitude are seen in the Virgo cluster
ellipticals NGC 4472 and 4649, but not in M87 (Couture, Harris and Allwright 1991).
If such gradients reflect progressive enrichment in a dissipative collapse, it would seem
that not all GCSs were formed in identical fashion - obviously worrying news in any
application that relies upon the supposed universality of the GCLF.
(ii) The globular cluster distributions mimic the distribution of underlying halo lu-
minosity in some but not all galaxies. In NGC 1399, for example, the GCS follows
the halo luminosity; in M87 and NGC 4472, the halo luminosity is more centrally
concentrated than the GCS (by about one power in a log-log representation). The
difference is apparently not in the GCSs, which are much the same in projection from
galaxy to galaxy, but rather in the halos themselves. If the GCSs are thought to form
in a protogalactic collapse phase, the disquieting implication is that the formative
processes (or at the very least the epoch of cluster formation) differ from galaxy to
galaxy; the implications for the LF are unclear.
(iii) The specific frequency S (Harris and van den Bergh 1981) of GCSs, a measure
of the relative richness of globular cluster populations, varies by an order of magni-
tude over galaxies in disparate environments: those giant ellipticals centrally placed
in big clusters such as Virgo and Fornax have a superfluity of globular clusters rel-
ative to galaxies in les;-privileged locations. Mergers of ordinary galaxies could not
produce this effect, since clusters and stars would presumably be accreted indiffer-
ently (or, indeed, the accumulated globular clusters might be disrupted, with a net
decrease in the specific frequency for the central cannibal). On the other hand, it has
long been known (Hanes 1977) that the GCLF in the cluster-rich galaxy M87 very
closely matches in shape those in several other more normal Virgo ellipticals, so the
population scalings may have no impact on distance determinations.
(iv) There are suggestions that the characteristic dispersion q of the GCLF depends
upon galaxy type, being somewhat smaller in spirals than in ellipticals (Harris 1991).
Such an effect might lead to distance scale errors in applications which use the local
(spiral-based) calibration in the determination of distances to incompletely-sampled
GCLFs in ellipticals, although the consequences are not in fact clearcut: if the dis-
persions differ, then we likewise have no guarantee of the equivalence of the turnover
magnitudes and the tool loses credibility. On the other hand, as noted above, the
shape of the GCLF in NGC 4594 (M104) seems consistent with that in ellipticals,
and there may be no cause for worry.
Space does not permit a full elucidation of some other problematic features, including
differences in mean metallicities from system to system (with potential implications for
star formation/IMF differences) and so forth. Clearly the situation is confused.

4. The Continuing Story


In various collaborations, we are exploring further aspects of this line of research in the
hope of clarifying some of the imponderables. For example, we (Hanes, Harris, Jacoby,
Pritchet and van den Bergh) are using the CFHT in a study of the luminosity functions
214

of the globular clusters and the planetary nebulae in the galaxies NGC 4494 and 4565,
an E/S pair likely to be at the same cosmic distance. This study will allow a direct
intercomparison of the zero-points in these two extragalactic distance indicators; moreover,
it will permit an important check on the consistency of the methods in galaxies of different
types. Other lines of endeavour include the study, by Hanes and Bridges, of the GCSs
associated with dominant cD galaxies in sparse groups, in an effort to understand whether
central dominance is in itself enough to guarantee a superfluity of globular clusters. Some
of these galaxies have associated cooling flows which, it has been suggested, may be being
deposited into the continuing formation of low-mass stars or globular clusters (Fabian,
Nulsen and Canizares 1984), another aspect to be explored.

5. Conclusions
It is probably sufficient merely to repeat the 'punch line' which I gave in the abstract:
namely, globular cluster systems (GCSs) differ in perplexing ways in various galaxies and
environments, and caution may be needed in their use as distance indicators, despite their
so far exemplary behaviour. They must, however, be telling us exciting things about the
formation, evolution and interaction of galaxies, and the time may be nearly ripe for the
development of a clear interpretation.

Acknowledgements
It is a real pleasure to acknowledge fruitful and long-standing collaborations with a variety
of co-workers, including Bill Harris, Chris Pritchet, Rene Racine, Terry Bridges, and
Sidney van den Bergh. This research has been supported in part by an operating grant
awarded by the Natural Science and Engineering Research Council of Canada.

References
Bridges, T.J. and Hanes, D.A. (1991) A.J. (submitted)
Bridges, T.J., Hanes, D.A. and Harris, W.E. (1991) A.J. 101469.
Couture, J., Harris, W.E. and Allwright, J.W.B. (1991) Ap.J. in press.
Fabian, A.C., Nulsen, P.E.J., and Canizares, C.R. (1984) Nature 310,733.
Fall, S.M. and Rees, M.J. (1977) MNRAS 181, 37P.
Hanes, D.A. (1977) Mem. RAS 84, 45.
Hanes, D.A. (1980) in Globular Clusters, NATO Advanced Study Institute ed. D.A. Hanes
and B.F. Madore (Cambridge University Press), p. 213.
Hanes, D.A. and Whittaker, D.G. (1987) A.J. 94, 906.
Harris, W.E. (1987) Ap.J. Lett. 315, L29.
Harris, W.E. (1991) Ann. Rev. Astron. Astrophys. in press.
Harris, W.E., Harris, H.C. and Harris, G.L.H. (1984) Ap.J. 276,491.
Harris, W.E. and van den Bergh, S. (1981) A.J. 86, 1627.
Thompson, 1. and Valdes, F. (1987) Ap.J. Lett. 315, L35.
215

DISCUSSION:

Cannon: I think your earlier statement, that fitting the cluster luminosity function to
get distances is analogous to fitting main sequences in colour-magnitude diagrams, is a bit
misleading. In the latter case you have one of the axes fixed (by the stellar temperature
scale), but in the former you really need to get round the maximum of the L.F. in order to
have a feature, otherwise you can slide the diagram along both axes.
Hanes: In MS fitting, where one uses colour-magnitude diagrams of V vs. B-V, one can
likewise slide along both axes since one has reddening effects to worry about. So there is
some ambiguity there. The globular cluster LF has sufficient curvature in the logarithmic
representation that it is clearly not self-similar even when sampled to well above the turnoff
(Le. population seatings and distance shifts ~ be distinguished) provided only that one
has a sufficiently rich population, such as is provided by a typical bright elliptical galaxy.
But these days we can almost routinely reach turnovers near B = 26 so the question is
moot in many cases.
HIGH-RED SHIFT TESTS OF no

B. Guiderdoni
In8titut d'A3troph,l3ique de Pam,
9Sbu Boulevard Arago,
F-75014 Pam.

ABSTRACT. This paper briefly reviews the present status of the classical cosmological tests involv-
ing high-redshift galaxies: the Hubble diagram, the Loh-Spillar test, the faint gaJ.axy counts and
the age of distant galaxies. The review focuses on the crucial problem of the correction of galaxy
evolution. As a matter of fact, the simplest analysis of the data leads to a low-density open universe.
Nevertheless the observations could be reconciled with the flat geometry predicted by inflation if
the cosmological constant was non-zero. As an alternative to this desperate solution, we show that
a scenario in which the evolution of galaxies is driven by merging in an no = 1 flat universe could
accomodate most of the data.

1 Introduction

The curvature of the Universe is measurable by means of geometrical tests involving high-
redshift galaxies. Some of these tests have been proposed in the early epoch of modern
cosmology, but the past status of observations forbade any significant conclusion. The
recent developement of sensitive detectors now allows to revisit the question. The interest
of these classical tests is their sensitivity to all matter, baryonic and non-baryonic, on very
large scales comparable to the curvature radius. Here galaxies are simply used as beacons
tracing the space-time structure. In principle, all the tests are straightforward and should
give unambiguous results. Unfortunately, galaxies are not inert beacons. They strongly
evolve in timescales of a few Gyr. The analysis of the tests must take this evolution into
account, by means of a model of spectrophotometric evolution which allows to compute
evolving synthetic spectra of galaxies from a minimum set of assumptions about the star
formation history. By coupling these spectra with classical cosmological models, apparent
magnitudes and colours can be derived, which take into account the entangling of the
cosmological and evolutionary effects in a consistent way. The purpose of this paper is to
describe the state-of-the-art in this rapidly-evolving field, by focussing on the problems
arising from the modelling of galaxy evolution. The reader wishing to know the general
background of the tests is referred to the good review proposed by Sandage (1988).

217
T. Shanks et al. (eds.J, Observational Tests of Cosmological Inflation, 217-231.
© 1991 Kluwer Academic Publishers.
218

2 Principles and fundamental formulae

In Friedman-Lemaitre cosmological models, the curvature radius of the Universe is:

Ro = Ho 1aoC 11/2 with ao = 0 0 + Ao - 1 (1).

Here c is the speed of light, Ho is the Hubble constant, 0 0 is the density parameter, and
=
AO is the reduced cosmological constant. If k = sign(ao) -1, the geometry is hyperbolic.
If k = +1 it is spherical. The case k = 0 corresponds to a flat universe, with "infinite"
curvature radius. This is the case predicted by inflation. We shall hereafter assume AO = 0
and come back to a non-zero AO in section 8.
e
If one defines == 1 + z, the space-time properties of the Universe can be computed by
means of the time element and the comoving radial coordinate (see e.g. Weinberg, 1972):

(2)

(3)

Then one can show that:


(i) The apparent angular size () of an object of rest-frame linear size I is given by

(4).
DA(Z) is the angular distance.
(ii) The apparent bolometric flux / from an object of rest-frame bolometric luminosity
L is given by
/ = L/411'Di(z) with DL(Z) == Ror.(z)e (5).
DL(Z) is the luminosity distance
(iii) The volume element of a spherical shell at redshift z is

dV(z) = -411' D~(z)cdt(z) (6).

Thus any measurement of the variation of one of these quantities with Z allows to constrain
the values of 00 and Ao. For that purpose, several tests have been introduced:
(1) Angular size () versus z. This test is a measure of DA(Z) and assumes that objects
of constant linear size I, the so-called "rigid rods" , are available. It has been operated on
cluster cores, without significant results (see Sandage, 1988).
(2) Apparent magnitude m versus z. This test is a measure of Ddz) and assumes that
objects of constant luminosity L, the so-called "standard candles" , are available.
(3) Sur/ace density 0/ galaxies N versus z. This test, recently called the "Loh-Spillar"
test, is a measure of the volume V (z) of spherical shells. It needs conservation of the comov-
ing number of galaxies. It also uses all galaxies while the previous tests use special objects
219

which can have anomalous behaviours. The problem is to get redshifts for a significant
sample of field galaxies.
(4) Surface density of galaxies N versus apparent magnitude m. This test, the classical
faint galaxy counts, is a measure of V(DL). As the "Loh-Spillar" test, it also needs con-
servation of the comoving number of galaxies. Since V and DL have related behaviours,
it is a priori less sensitive to cosmology. Nevertheless its great advantage is to use only
magnitudes and to probe the universe very deeply.
(5) Age of the galaxies tgal versus z. It gives at each z a lower value of the age of the
universe t(z). The test is mainly sensitive to the Hubble constant Ho. The problem is that
it uses the reddest galaxies, which are likely the oldest ones, but can be strongly biased.
All tests assume constant properties for galaxies, which is far from reality. In fact,
galaxies evolve. Their luminosities and colours vary with redshift. Moreover it is even very
likely that at least some of them can merge, and the assumption of conservation of the
comoving number density is questionable. So it is necessary to correct the tests for the
intrinsic evolution before addressing the value of 0 0 , The computation of these k- end e-
corrections (for red shift of the spectra and intrinsic evolution) need specific models which
are described in the next section. Finally, it can be shown that all tests are independent
of Ho except through the evolutionary time scales in the e-corrections. Hereafter we shall
take Ho = 50 km s-I Mpc- I .

3 The models of spectrophotometric evolution

The principles of this kind of models are the following ones: at each time step, stars form
from the gas content according to the classical parameters (Star Formation Rate and Initial
Mass Function). These stars are placed on the Zero-Age Main Sequence (hereafter MS)
of the HR diagram. The models use compilations of stellar evolutionary tracks taking
into account the MS, Giant Branch (hereafter GB) and late stages of stellar evolution
in order to compute at each time step the distribution of the stellar populations in the
HR diagram. This distribution is combined with a library of stellar spectra (or with stellar
colours) in order to estimate the synthetic spectrum of the stellar population. The apparent
magnitudes and colours at redshift z are derived from the spectra after coupling with the
standard cosmological models and convolving with the response curves of filters. After B.
Tinsley's pioneering works, there are basically three families of models currently available
in the literature, which have been extensively compared to observations.
Bruzual's model described in Bruzual (1981, 1983) and Bruzual & Kron (1980) has been
largely distributed throughout the world. In particular, it has been used by Koo (1986), in
Berkeley for the studies of the optical counterparts of high-redshift radiogalaxies, and in
Durham to analyse the deep surveys. The early version used low-resolution stellar spectra
and OAO UV spectra, and Yale tracks for the MS, with a semi-empirical GB. Then the
model has been upgraded in various ways, which contribute to a rather confusing situation.
A last version has been proposed by Charlot & Bruzual (1990) with stellar tracks by Maeder
& Meynet (1988) and a number of technical improvements.
The Japanese model is an adaptation of the model of Arimoto & Yoshii (1986, 1987) to
high-redshift galaxies, proposed in Yoshii & Takahara (1988). In its published versions, the
220

model has only stellar colours and the galaxy spectra in the UV are borrowed to Bruzual
and connected to the visible spectra in a rather questionable way. It is also based on Yale
tracks and has the advantage of being consistently coupled to a model of chemical evolution.
The lAP model is described in Guiderdoni & Rocca-Volmerange (1987, 1988), and
Rocca-Volmerange & Guiderdoni (1988) on the basis of Rocca-Volmerange et al. (1981).
It is also used in Edinburgh. The early version used Yale tracks, but we recently introduced
tracks from Maeder & Meynet (1988) and VandenBerg (1985), as well as a more refined
computation of the evolution of massive stars. The results presented in this review are
based on this upgraded version which fully confirms the previous results. There is no
coupling yet with chemical evolution but this will be done soon with the model described
in Rocca-Volmerange & Schaeffer (1990). The internal extinction is estimated from the
gas content and simple models of metal enrichment and transfer. The nebular emission is
computed from the flux of ionizing photons. Here we keep a standard IMF with a slope
x = 1.7 for stars more massive than 2M0 (observational value from Scalo, 1986)' x = 1.35
between 1 and 2M0 , and x = 0.25 below. The synthetic spectra satisfactorily reproduce
the colours of the Hubble sequence of spectral types, with a single parameter, namely the
time scale for star formation. The age of galaxies can be a secondary parameter. There is
no need of varying the IMF to reproduce the colour sequence, as in some versions of the
other models.
The quality of a model lies in the consistency of its stellar inputs and it its capability of
accomodating a large amount of constraints at all wavelengthes (UV, visible and near-IR)
with the smallest number of free parameters.

4 The Hubble diagraIn

A "standard candle" with current absolute magnitude M>. (in a filter centered at A) has an
apparent magnitude at red shift z:

(7)
with the bolometric distance modulus (m - Mhol(Z) = 510g(DL(Z)/Mpc) + 25. The status
of the test was renewed by the observations of optical counterparts of radiogalaxies in the
visible and in the near-IR at redshifts Z > 1. The objects are selected on the basis of their
radio properties in catalogues of powerful sources (3C, IJy, Parkes, 4C). The spectroscopic
redshifts of the optical counterparts can be obtained from their strong emission lines.
The observer-frame B or V bands probe the UV in the galaxy rest frame, which is
very sensitive to bursts of star formation and some scatter is expected. Nevertheless, the
near-IR bands could be less sensitive to the history of star formation. So less scatter is
expected and there was some hope to derive a value of 00 from K observations. Spinrad
& Djorgovski (1987) proposed 00 = 0.4 ± 0.6. More recently Weir et al. (1990) showed
that the uncertainty in the history of star formation erases any cosmological information.
Figure 1 shows examples of the V -band and K -band Hubble diagrams and predictions
for various SFR histories with 0 0 = 1 and 0.1. It is difficult to descriminate between the
models. Nevertheless, the optimistic side will consist in remarking that with standard SFR
histories, the data are reproduced within the current range of plausible values for 00.
221

There is some debate about the origin of the scatter in the IR Hubble diagram. It
must be noticed that, even if this diagram is too scattered to give a relevant cosmological
information, the scatter is remarkably less than would be expected if the galaxies had strong
bursts of star formation, as shown in figure 2 (Rocca-Volmerange & Guiderdoni 1990).
The scatter is not consistent with galaxy formation at zfor < 5 and SFR timescales ~0.1
Gyr. That seems to exclude the "young radiogalaxy" hypothesis in which star formation is
induced by the radio jet, unless some fine-tuning in the SFR is introduced.

5 The Loh-Spillar test

If cp(L, z) is the luminosity function (LF) at z, the surface density of galaxies in solid angle
dw with flux (per Hz) brighter than II is:

(8)

with the luminosity LI per Hz:


(9).

The quantities A(z) and B(z) only depend on z and on the values of 0 0 and Ao and can be
computed from eq. (6) and (5). The total flux (per Hz) from galaxies at redshift z is:

dzdw c
F(> !I) = --A(z)B(z)
411" Ho
1 00

Lr
cp(L,z)LdL (10).

Providing (i) the LF has the Schechter form with characteristic luminosity L * and con8tant
slope a, and (ii) all galaxies evolve at the same rate (Le. single-type evolution):

c== F(>/!) = r(a+2,xj) WIt


. h
XI
LI
==- (11).
IIN(> !l) Xjr(a + 1, XI) L*
r(8, X) is the incomplete gamma function. The method consists in measuring N(> /I) and
F(> !I) at fixed z, computing C, and deriving XI from eq. (11). Then putting this value
in the integral of eq. (8) equal to cp*r(a + 1, XI), it is possible to determine cpo A(z)c3 / Hg.
This quantity does not depend on Ho. Providing (iii) there is no number-density evolution
(con8tant cp*), it is easy to derive 00 from the variation of A( z) versus z.
This method has been used in Loh & Spillar (1986) and Loh (1986, 1988). They obtained
narrow-band multicolor photometry of ~ 1000 galaxies in five 7' x 10' fields at m800nm < 22.
The redshifts (up to z ~ 0.75) were determined from the spectrophotometry. Finally the
authors' last analysis yielded 0 0 = 1.1:!:g:~ at the 95 % confidence level (see figure 3).
Formally, the method has two advantages. It avoids the k-corrections by using shifted
passbands for the various redshifts, as well as the e-corrections since luminosity evolution
is directly measured (by the depth x, of the survey at redshift z) under the assumptions
(i) and (ii). Nevertheless, there has been much debate about the effect of dispartures from
the three assumptions (i), (ii) and (iii).
222

• Brlchtat CI...ter lI.m...... • Brlchtal Cluter lI.m......


25 • 3CIt
• 3CR
• IIrIcbt I aDd RG • IIrIcbt I aDd RG

>20

15

• 3CIt • 3CR
20 • 11,. • I ly
• 3C 328.1 B compo • 3C 328.1 B compo

.. 15

, ,
10

0.1 0.1
r.doblll s redoblll s

Figure 1: The V -band and K -band Hubble diagrams for no = 0.1 (left-hand panel) and 1 (right-
hand panel). Various SFR time scales and redshifts of galaxy formation can accomodate the data
within the observed scatter .

• 3CR
20 .• I Jy
11 =1.~
3C 326.1 B compo

10

:.: Ifi •

,
10

0.1
REDSHIFT

Figure 2: The K-band Hubble diagram with no = 1. The SFR time scale is 0.1 Gyr and the
redshift of galaxy formation is specified. The dashes-and-dots give the curve for the average lUass
evolution due to merging (see section 9). From Rocca-Volmerange & Guiderdoni (1990).
223
faint ,alaxy couols 10 BI

10'~~------~--------~----~
20 10
Zf or = (a )
10'
..,
•U
Q.
::E
..,
'0
.
I


~IOOO
10'

4 .n
..,
.
2-
0.2 '"
s: 3 I

<t 0.9 ... u


100 ...• Plterson et al. 79

•-S-
• Jarviloo and Ty"oR 81
E o Jona•• t al. flO
z o• Koo 88
Infante et a1. 88
2.0 10 + ~~'.1~i8 Dumbleto
* lI.tcalf. at a1. 90
0.00 0.25 0.50 0.75 1.00 o• TJaon 88
LIIl,. et a1. flO

0.1
15 20 25
appareot melDltude m

fainl ,alaxy counls In BI redahift cllstrlbuUoD


10·
nO = 0.' ( b)
~zfor
no = (c )
10~
Z
for
= 60 20<81<22.5

.
Broadhuret at a1. 88
10'
I
+ LD8S daap survay
)(II
• (ColI_ at a1.
.!ilooo
S
N
I
.. M
Z (1'0)
~ 100
E
6 Petenon It al. 79
• Jarviloo and TYSOD 81
o Jon••• l a1. 90
• Koo 88
(03')
Z o Intanll .t a1. 88
l~)
10 + ~~'.1~i8Dumblelo
20
* lIetcalf. et a1. flO
o• rIJ:D 88
,.atal.90

0.1
15 20 25 0.2 0.4 0.6 0.8
apparent malDltude m reclsblft 11

Figure S: The result of the Loh-Spillar test in its early version. From Loh & Spillar (1986).

Figure 4: (a) and (b) Sensitivity of the predicted counts to no and the redshift of galaxy formation
Z! or. Dotted lines: no evolution. (c) Predictions for the redshift distribution.
224

First some-type dependent evolution would make the apparent A(z) lower than the true
A(z), and consequently OOapp > 00. Various estimates have been proposed: ~Oo ~ -0.16
(crude estimate in Loh & Spillar early paper); ~Oo ~ -0.4 (Loh 1988; from Tinsley's
model); ~Oo ~ -0.4 (Yoshii & Takahara 1990; from the Japanese model); up to ~Oo ~ -1
in Bahcall & Tremaine (1988), by means of a fit of the data with Tinsley's model. The
basic problem of the present status of the Loh-Spillar test is that the measured evolution
is ~MB == MB(Z = 0) - MB(Z = 0.75) = 0.3 ± 0.2 for red galaxies and 0.2 ± 0.2 for
blue galaxies while all models of spectrophotometric evolution predict ~MB = 0.6 to 1
mag for red galaxies (and much weaker evolution for later galaxies). These estimates of
the systematic error due to type-dependent evolution are based on the models, with no
concern of this measure, except in Bahcall & Tremaine. Of course the "low-Iuminosity-
evolution scheme" favours assumption (ii). But we must admit that we do not understand
the evolution in the Loh-Spillar sample.
Secondly number-density evolution would make the apparent A(z) higher than the
true A(z), and consequently OOapp < 0 0 , The estimates span from ~Oo ~ 0.1 in Loh &
Spillar (1986) and Loh (1988), to ~Oo ~ 2 in Bahcall & Tremaine (1988). The reason of
this discrepancy is that Loh & Spillar use a "low-merging scheme" based on observational
estimates of pairs of bright galaxies while Bahcall & Tremaine give a theoretical estimate of
the accretion of companions due to dynamical friction. Moreover Caditz & Petrosian (1989)
showed that assumption (i) is not consistent with the data, independently of any assumed
value of 0 0 . In particular, the slope a of the LF evolves and can be as flat as a = -0.2
at redshift z > 0.4. Such an evolution is opposite to that expected in any simple merging
scenario. The authors suggest that the most likely value would be 0 0 ...., 0.2. Finally Omote
& Yoshida (1990) claimed that gravitational lensing could severely affect the method. All
these works make the early result of the test very doubtful.

6 The faint galaxy counts

The number of galaxies with type j at redshift [z, z + dz[ contributing to the counts per
magnitude bin around apparent magnitude m~ (through the ~ filter) is:

(12).

dV is given in eq. (6). t/>;(M~) is the comoving LF for type j determined from the analysis
of nearby galaxies. The absolute magnitude M~ is computed from m~ by means of eq. (7).
Predictions for magnitude, redshift and colour distributions are computed by integrat.ing
the elementary contributions.
Tinsley (1980), Bruzual & Kron (1980), Shanks et at. (1984), King & Ellis (1985),
and Yoshii & Takahara (1988) interpreted galaxy counts brighter than m ~ 24 under
the assumption of number-density conservation. In spite of large uncertainties, the general
trend of their conclusions, as reviewed by Koo (1989), is that small values of qo are required
to fit the observed slope of the magnitude distribution. At magnitudes fainter than m ~
24, the influence of space curvature on the counts becomes stronger when evolution is
simultaneously taken into account.
225

Guiderdoni & Rocca-Volmerange (1990) give an analysis of these deep surveys. The
straightforward interpretation is based on the nearby luminosity functions for each spec-
tral type, and on the extrapolation in the past of the standard scenarios reproducing the
properties of the nearby galaxies, under the assumption of conservation of the comoving
number density. Figure 4ab shows the sensitivity of the standard predictions in the N(m)
diagram to no and %Jor. Figure 4c shows the fit of the redshift distribution of the Durham
deep surveys (Broadhurst et al. 1988; Colless et al. 1990). Thus the fit of the number
counts (in all bands from UV to red) and redshift distribution requires a low no (. . . 0.1)
and a high value of %Jor ( ..... 10).
It is not easy to get rid of this conclusion. As a matter a fact, it seems to be unsensitive
to uncertainties in the inputs of the model. Metcalfe et al. (1990) and Yoshii & Peterson
(1991) give the same result on the basis of their models of evolution which have other
stellar data. This conclusion is very robust since it is due to the smaller volume elements
at high redshifts for no = 1. Nevertheless there are two caveats. First the models must be
normalized to twice the local value of 4>* given by Efstathiou et al. (1988). The analysis of
the APM survey by Maddox et al. (1990) suggests either a hole in the local distribution of
galaxies or non-standard evolution. Secondly it might be possible to fit the counts with a
population of dwarfs. Recently Cowie et al. (1991) claimed that the K-band counts can be
reproduced by these dwarfs with no = 1, but failed to reproduce the B-band counts. The
puzzle is to understand why this huge local population is undetectable in all other surveys
(e.g. Binggeli et al. 1990).
• IIrIcbte.t ct....r ......ben
.......
2
• ct 0024+ 1884 ....
2 .........

.

I

3r-----~----~------~----_r----~ 3r-----~----~------r-----_r----_,
• n.ld .....pl. (Koo eel • ".Id nmple (Koo eel
....
2 --_.- .•. _._._.
..
1-I

o o
o 0.2 0•• 0.' 0.11 o 0.2 0 .• 0.' 0.11
ndobilla

Figure 5: Optical colour against redshift for various SFR histories and 0 0 = 0.1 (left-hand panel)
and 1 (right-hand panel). Dashes-and-dots: "reddest" model, with SFR time scale 0.1 Gyr and
redshift of galaxy formation zJr,r = 40. Dotted line: no evolution.
226

7 The ages of distant galaxies

The discovery of optical counterparts of radio-galaxies at redshifts z > 3 lead some authors
to use the age of these distant galaxies as constraints on the values of the cosmological
parameters. The first work by Lilly (1988) based on 0902+34 at z = 3.395 gave an age ~ 1
Gyr. As a matter offact, the age ofthe Universe at redshift 3.395 is 1.42 Gyr with Ho = 50
and no = 1. Then several authors gave very small ages for the distant galaxies (Bithell
& Rees 1990; Chambers & Charlot 1990). Thus the situation appears confused. In fact,
Rocca-Volmerange & Guiderdoni (1990) showed that any age between"" 0.5 Gyr and"" 2.5
Gyr can fit the spectrophotometry within the error bars. Consequently it is impossible to
put constraints on cosmology from these objects in the present state of the data.
May be the galaxies at redshifts "" 0.5 are more interesting. The age of the Universe was
about two thirds of its current age. It turns out that the reddest galaxies at these redshifts
are well reproduced by models without evolution (Colless et al. 1990). Guiderdoni &
Rocca-Volmerange (1988) showed that models in a low-density universe, which have weak
evolutionary corrections, can fit the data. On the contrary, it is difficult to accomodate the
data with an no = 1 universe as shown in figure 5. The situation still degrades if Ho is larger
than 50. Several solutions could be imagined to this problem: (1) A low-value of Ho. (2)
Intrinsic extinction. It is not plausible since cluster galaxies which are devoid of gas show
exactly the same behaviour as field galaxies. (3) Metallicity effect. The reddest galaxies
might be metal rich. Nevertheless Yoshii & Takahara (1988) find the same difficulty with
their model taking into account chemical evolution. (4) The stellar clock is not accurate.
This is related to the question of the ages of globular clusters which are only marginally
consistent with the age of the universe 13 Gyr if Ho = 50 and 00 = 1 (see this conference).

8 Are the tests consistent with inflation ?

If we summarize the results of the tests, it appears that they seem to suggest an open
universe with a low value of 00. The Hubble diagram is not constraining, and is consistent
with this value. The Loh-Spillar test, within the presently-unsolved problems, seems to
give a low 00 in its re-analysis by Caditz & Petrosian. The faint counts suggest a low 00·
Finally there is no difficulty with the colours of mild-redshift galaxies.
In order to save inflation from the faint counts, Fukugita et a1. (1990) introduced a
non-zero cosmological constant. The effect of Ao # 0 is to increase the observable volume
of the universe and consequently the observable number of galaxies. The authors repro-
duce the counts with no = nb = 0.1 (consistent with primeval nucleosynthesis) and the
corresponding Ao = 0.9 for the flat universe. On the other hand, Ao = 0.9 increases the age
of the universe up to 25 Gyr and eliminates the difficulty with the colours of mild-redshift
galaxies. Nevertheless, despite these successes (Efstathiou et at. 1990), cosmologists are
very reluctant to introduce a non-vanishing Ao since it would require the fine-tuning that
the inflationary paradigm precisely tries to avoid.
Thus if we believe in inflation, we must introduce a new assumption in order to rec-
oncile the predicted counts with the data. An appealing possibility is the introduction of
strong number evolution, in addition to the normal, unavoidable luminosity evolution. This
227

assumption is a natural prediction of the CDM flat universe in which galaxies form from
smaller units in a hierarchical scenario (see e.g. Frenk et al. 1988). In Rocca-Volmerange
& Guiderdoni (1990) and Guiderdoni & Rocca-Volmerange (1991) we presented a model in
which the evolution of galaxies is driven by merging. Figure 3 shows that the observed scat-
ter in the Hubble diagram is consistent with this model providing giant radiogalaxies are
objects acccreting companions at a rate higher than the mean. Finally, number evolution
strongly perturbates the Loh-Spillar test and could make it useless.

9 The merging-driven evolution of galaxies in a Hat universe

We model number-density evolution by assuming:


(1) self-similarity. for the comoving mass function at redshift z:

tjJ(M, z)dM = tjJ*(z)( ~z))(>(z) exp - ~z) ~~) (13)

with M* (z) being the characteristic mass at the knee, tjJ* (z) a characteristic density and
o(z) the slope. These three parameters are allowed to vary with redshift.
(2) conservation of the total comoving mass density. With eq. (13), this leads to:
M*(z)tjJ*(z)r(o(z) + 2) = M*(O)tjJ*(O)f(o(O) + 2) (14)
The values M*(O), ,p*(0) and 0(0) are fixed in order to fit the current LF of galaxies, with
the consistent mass-to-light ratios computed in our model.
The actual number evolution probably acts on M*, tjJ* and 0. We introduce comple-
mentary approaches to this evolution, by assuming that only two of the three parameters
change. We parametrize the evolution of one of the factors of eq. (14) as (1 + z)'1, and
we consistently compute the evolution of one of the other factors with keeping the last one
as a constant. Thus, with the condition that galaxies are more numerous and less massive
for increasing z, we get three series of models, namely the so-called M* - tjJ* model (with
tjJ*(z) = tjJ*(O)(1 + z)'1 and constant M*,p*), M* - 0 model (with M*(z) = M*(O)(1 + z)-'1
and constant M*f(o + 2)), and tjJ* - 0 model (with tjJ*(z) = tjJ*(O)(1 + z)-'1 and constant
tjJ*r(o + 2)). The actual evolution is probably a compromise between these three limiting
cases. The parameter 11 is free. Spiralling of companions down to galaxies due to dynamical
friction could be the regulating process of merging and number evolution, following Bahcall
& Tremaine (1988). We show that it leads to values of 11 - 1.5 to 2.
Our continuous, exponentially-decreasing SFR's are interpreted as the average of the
successive bursts of star formation associated with merging. Such a behaviour is at least
suggested by the simulation of Baron and White (1987) who show that in a forming galaxy,
star formation triggered by cloud collisions exponentially decreases with time.
If there is number evolution and mass conservation, the faint galaxies are the so-called
building blocks with absolute magnitudes which can be as faint as those of typical dwarfs.
The question is to know whether a low intrinsic surface brightness combined with the
"cosmological fading" in 1Olog(1 + z) makes them detectable at the level of the currently-
available surveys. In order to solve this problem, we introduced in Guiderdoni & Rocca-
Volmerange (1991) the consistent computation of surface brightness, under a third assump-
tion: (3) the mass profiles of building blocks depend on mass according to relations fitted on
228
faint lalaxy counla In 8, redahlll distribution
10·r-r--------,--------~----_,
I a) 11=1.~ I b)

60 20<8,<22.11

.
Broadhuret et aI. 88
10'
+ LDSS deep au"e),
• (Collen et al. go)
JlIOOO
..
M 40
't Z
~ 100 6 Petenon et al. 78

e
• ~arvl. and TYlon 81
o Jone. et al. go
i' • Koo 88
o In'ante et al. 88
10 + ~rya1~B8Dumbleto
20
* lIetealle .t .1. go
* boon 88
• Ully et at go

20 25 0.2 0.4 0.6 0.8


apparenl malnllude m redshlll z

colour dlatrlbullon faint galaxy counls in K

50 108
Q o"1 22<8,<23
II 11=1.!) lIetoalle et al. 80
:0 105
~ 40

..
0
"0
GO " 10'
0 i
c:i
• 30
CI .§ 1000
e
...
~
...
<3
N
I

• Cowie et al. 91

..
." 100
"- 20
Ii eZ o Glazebrook el al 90

.s 10
1:I
ez
10 -

( C) Cd)
0 0.1
0 I 2 3
apparenl colour 8,-'+ 15 20
apparent magnilude m
25

Figure 6: Predictions for no = 1 and number evolution as (1 + z}". The redshifts of galaxy
formation are Zfor = 5 (EjSO) to 2 (late spirals). Three models are presented with '7 = 1.5 (see
section 9). The range for '1 = 1 and 2 is sketched in fig. 6a. (al Number versus apparent magnitude.
(b) Redshift distribution. (c) Colour distribution. The figures are drawn from Guiderdoni & Rocca-
Volmerange (1991). (d) K-balld counts with no = 0.1 and 1, and a typical scenario of number
evolution (M*"'* with '1 = 1.5) with no = 1. Dotted lines: no evolution. The scenario of number
evolution is consistent with the data.
229

nearby galaxies. For that, we compute their current, rest-frame absolute magnitudes MB
by correcting the absolute magnitudes at z for cosmology and evolution, and we use stan-
dard observational relations of the luminosity profiles with MB. We also take into account
the effect of seeing.
Figure 6abc gives the results for the three limiting cases. Apparently all kinds of number
evolution seem to conspire to reproduce the observational slope, whatever parameters of
the LF are actually affected by evolution. The values Tl = 1 to 2 which fit the data are
also those suggested by a model of dynamical friction. The surface-brightness threshold
of Tyson (1988) at J.tB = 28.7 mag arcsec- 2 is so deep that the building blocks are visible
in the survey. The falling down of the predictions is due to the crude modelling of seeing.
The second important point is that the condition of high zlor is not necessary. The fit
is here given for zlor = 2 to 5, but the predictions are not very sensitive to the precise
values. Figure 6b gives the redshift distribution. Figure 6c shows the colour distribution at
magnitude 22.5. This last fit is more sensitive to the choice of Z lor. In figure 6d, it appears
that the deep K counts (Cowie et al. 1991) are consistent with this scenario of merging.

10 Conclusions

Clearly there is no definitive conclusion yet. The most straightforward interpretation of


the tests taking into account the evolution seems to suggest a low-density open universe,
within the uncertainties. We could be satisfied for that if the inflationary theory did not
exist. But the inflationary paradigm is so attractive that it is worthwhile to revisit the
tests. All problems for reconciling the data with inflation vanish if we introduce a non-zero
cosmological constant. But this seems so difficult to be justified that we could really wonder
whether the cure is not worse than the disease. An alternative possibility is to introduce
number evolution in the galaxy population. Such an evolution is a prediction of bottom-up
hierarchical models of galaxy formation and should be testable. In particular the stability
of disks puts severe constraints on this scenario (Ostriker 1990).
Finally, it is interesting to recall that Hubble (1936) attempted to measure the curvature
of the universe from faint counts. With the value Ho ~ 530, he found no ~ 14. Some way
has been done since this value and we can be reasonably optimistic for the future.

Acknowledgements. I am grateful to Brigitte Rocca-Volmerange with whom much of the work


reviewed here has been carried out. It is also a pleasure to thank the organizers for their hospitality
and the quality of the meeting. Finally thanks to Alain Blanchard for help and comments.

References

Arimoto, N., & Yoshii, Y., 1986. Astr. Astrophys., 164, 260.
Arimoto, N., & Yoshii, Y., 1987. Astr. Astrophys., 173,23.
Bahcall, S.R., & Tremaine, S., 1988. Astrophys. J., 326, L4.
230

Baron, E., & White, S.D.M., 1987. Astrophys. J., 322, 585.
Binggeli, B., Tarenghi, M., & Sandage, A., 1990. Astr. Astrophys., 228, 42.
Bithell, M., & Rees, M.J., 1990. Mon. Not. R. astr. Soc, 242, 570.
Broadhurst, T.J., Ellis, R.S., & Shanks, T., 1988. Mon. Not. R. astr. Soc, 235, 827.
Bruzual, G., 1981. Ph.D. dissertation, University of California Berkeley.
Bruzual, G., 1983. Astrophys. J., 273, 105.
Bruzual, G., & Kron, R.G., 1980. Astrophys. J., 241, 25.
Caditz, D., & Petrosian, V., 1989. Astrophys. J., 337, L65.
Chambers, K.C., & Charlot, S., 1990. Astrophys. J., 348, Ll.
Charlot, S., & Bruzual, G., 1990. preprint.
Colless, M., Ellis, R.S., Taylor, K., & Hook, R.N., 1990. Mon. Not. R. astr. Soc, 244, 408.
Cowie, L.L., Gardner, J.P., Wainscoat, R.J., & Hodapp, K.W., 1991. preprint.
Efstathiou, G., Ellis, R.S., & Peterson, B.A., 1988. Mon. Not. R. astr. Soc, 232, 431.
Efstathiou, G., Sutherland, W.J., & Maddox, S.J., 1990. Nature, 348, 705.
Frenk, C.S., White, S.D.M., Davis, M., & Efstathiou, 1988. Astrophys. J., 327, 507.
Fukugita, M., Takahara, F., Yamashita, K., & Yoshii, Y., 1990. Astrophys. J., 361, L1.
Guiderdoni, B., & Rocca-Volmerange, B., 1987. Astr. Astrophys., 186, 1.
Guiderdoni, B., & Rocca-Volmerange, B., 1988. Astr. Astrophys. Sup pl. Ser., 74, 185.
Guiderdoni, B., & Rocca-Volmerange, B., 1990. Astr. Astrophys., 227, 362.
Guiderdoni, B., & Rocca-Volmerange, B., 1991. submitted.
Hubble, E., 1936. Astrophys. J., 84, 517.
King, C.R., & Ellis, R.S., 1985. Astrophys. J., 288, 456.
Koo, D.C., 1986. Astrophys. J., 311, 651.
Koo, D.C., 1989. in The Epoch of Galaxy Formation, p. 71, ed. C.S. Frenk al., Kluwer.
Lilly, S., 1988. Astrophys. J., 333, 161.
Lilly, S.J., Cowie, L.L., & Gardner, J.P., 1990. preprint.
Loh, E.D., & Spillar, E.J., 1986. Astrophys. J., 307, Ll.
Loh, E.D., 1986. Phys. R. Letters, 57, 2865.
Loh, E.D., 1988. Astrophys. J., 329, 24.
Maddox, S.J., Sutherland, W.J., Efstathiou, G., Loveday, J., & Peterson, B.A., 1990. Mon. Not.
R. astr. Soc, 247, IP.
Maeder, A., & Meynet, G., 1988. Astr. Astrophys. Suppl. Ser., 76, 411 .
Metcalfe, N., Shanks, T., Fong, R., & Jones, L.R., 1990. preprint.
Omote, M., & Yoshida, H., 1990. Astrophys. J., 361, 27.
Ostriker, J., 1990. in Evolution of the Universe of Galaxies, p. 25, ed. R. Kron, Astr. Soc. Pac ..
Rocca-Volmerange, B., Lequeux, J., & Maucherat-Joubert, M., 1981. Astr. Astrophys., 104, 177.
Rocca-Volmerange, B., & Guiderdoni, B., 1988. Astr. Astrophys. Suppl. Ser., 75, 99.
Rocca-Volmerange, B., & Schaeffer, R., 1990. Astr. Astrophys., 223, 427.
Rocca-Volmerange, B., & Guiderdoni, B., 1990. Mon. Not. R. astr. Soc, 247, 166.
Sandage, A., 1988. Ann. Rev. Astron. Astrophys., 26, 561.
Shanks, T., Stevenson, P.R., Fong, R., & MacGillivray, H.T., 1984. Mon. Not. R. astr. Soc, 206,
767.
Scalo, J.M., 1986. Fundam. Cosmic Phys., 11, 1.
Spinrad, H., & Djorgovski, S., 1987. in Observational Cosmology, IA U Symp. 124, p. 129, ed. A.
Hewitt et at., D. Reidel.
231

Tinsley, B.M., 1980. Astrophys. J., 241, 41.


Tyson, J.A., 1988. Astron. J., 96, 1.
VandenBerg, D.A., 1985. Astrophys. J. Suppl. Ser., 58, 711.
Weinberg, S., 1972. Gravitation and Cosmology, Wiley, New York.
Weir, W.N., Djorgovski, S., & Bruzual, G., 1990. in Evolution of the Universe of Galaxies, p. 25,
ed. R. Kron, Astr. Soc. Pac ..
Yoshii, Y., & Takahara, F., 1988. Astrophys. J., 326, 1.
Yoshii, Y., & Takahara, F., 1989. Astrophys. J., 346, 28.
Yoshii, Y., & Peterson, B.A., 1991. preprint.

DISCUSSION:

M. Pierce: How sensitive are your models to chemical evolutionary effects on the giant branch ?
B. Gulderdonl: The answer to this question depends on the wavelength of the observations and
the age range of the galaxy. In the rest-frame UV which is observed in the visible, and for young,
high-redshift galaxies, the contribution of the giant branch is low. Of course, this contribution is
higher for older galaxies in the IR. I would like to recall that the Japanese model, which has chemical
evolution, seems to find constraints for cosmology which are similar to ours.
D. Lynden-Bell: With the new knowledge that IRAS has given us that rapid star formation
can occur in red galaxies, how do you treat absorption ? Do you allow more absorption in young
galaxies?
B. Guiderdoni: Our intrinsic extinction is consistently computed from the gas content of the
galaxy, its metallicity, and a simple model of transfer. In this model, young galaxies have more
extinction than old ones. Nevertheless, we did not try to reproduce the extreme IRAS objects which
are too scarce to be significant in the classical tests.
R. Ellis: You introduced merging to help explain the galaxy counts, but didn't then change your
evolutionary models accordingly. Surely mergers will lead to a major increase in the star formation
rate with implications for both luminosities and colours?
B. Guiderdonl: That is plausible. On the other hand, I would like to emphasize that our merging
scenario is the accretion of numerous small companions by big galaxies and that the numerous bursts
of star formation can average to form the continuous SFR we used.
R. Fong: There are now being produced very good data on number-colour distributions at faint
magnitudes. Won't these help constrain the possible models? If the colours are well-fitted by
standard models, might this not be indicative that merging may not be a significant effect ? Might
expect merging to change the colours and distort the colour distributions given by standard models.
B. Guiderdoni: The colour distributions are very sensitive to the history of star formation. I don't
think one can reject merging on this single basis. I would rather say that once a model of merging
is given, they would help to constrain the influence of merging on the star formation history.
COSMOLOGY WITH GALAXIES AT IDGH REDSHIFTS

SIMON 1. LILLY
Department ofAstroMmy, University of Toronto
60 St George Street
Toronto, Ontario M5S lA7
Canada

ABS1RACT. Following some remarks on the methodology of the classical cosmological tests, recent
observations of galaxies at high redshift are examined to see whether they place any constraints on the
cosmological models. It is concluded that, at least at present, they do not. The principle difficulty is that the
luminosity evolution of galaxies is poorly constrained.

1. Introduction

The past few years have seen a dramatic increase in our observational knowledge of galaxies at
high redshifts, with quasars and radio galaxies (which provide a clearer view of their stellar
populations) being found at ever higher redshifts and with large-scale redshift surveys of faint
galaxies being undertaken. In this context, it is of interest to reexamine the classical cosmological
tests to assess their potential for constraining cosmological models.
The purpose of this paper is two-fold: Firstly, to show the effects of different models on the
observables of objects at high redshift (and in particular the effects of the often ignored A '# 0
models) and, secondly and mindful of the well-known difficulties and assumptions involved, to
see if any models can be rigourously excluded and/or whether a consistent "best-fit" model
emerges. In this regard three representative cosmological models are considered, with [Oo.A] set
equal to: (a) [1,0]; (b) [0.1,0]; and (c) [0.1,0.9]. Models (a) and (c) are of course consistent with
inflation in having flat geometries with k = O. Sandage (1988) has comprehensively reviewed the
theory and practice of these classical tests, while Peebles (1984) has examined the effects of A '#
O. It should be stressed at the outset that it is the view of the author that no such model does
emerge given the enormous uncertainties of galactic evolution, and consequently that the
observations of high redshift objects at present provide no compelling constraint on the geometry
of the Universe.

1.1 ABSOLUTE AND DIFFERENTIAL TESTS

A key distinction should be drawn between absolute and differential tests. The former may be
characterized by the use of input parameters for the local Universe that are derived from quite
independent (and often quite different) observational material or theoretical analyses. In contrast,
the latter involve only the study of differential effects with redshift using observational material
that is, as far as possible, the same at different redshifts. Examples of absolute tests include the

233
T. Shanks et al. (eds.), Observational Tests o/Cosmologicallnf/ation, 233-241.
© 1991 Kluwer Academic Publishers.
234

well-known comparison of the expansion rate of the Universe with globular cluster ages and, of
more relevance to this article, arguments concerning the number of gravitational lenses seen at
high redshifts (e.g. Fukugita et all990a, Turner 1990) and all other tests involving knowledge of
the local galaxy luminosity function (e.g. Cowie et al 1991). Obvious examples of differential
tests include the classic magnitude-redshift and angular size-redshift tests. While the absolute
tests may ultimately provide a stronger confirmation of the Big Bang cosmology since they are
based on independent measurements, it is likely that the differential tests may be more practical
for determining the correct choice of cosmological parameters in the near term.
As an interesting aside it should be noted that the two cosmological models that are
consistent with inflation (i.e. those with k = 0, with and without a cosmological constant) have
the same differential behaviour for redshifts z > 00- 1(3 because they have the same flat geometry
and the same R(t) dependence once the A-model becomes matter dominated. A flat A-model can
be regarded for many purposes at high redshifts as an Einstein-de Sitter Universe with an
effective Ho that is reduced from its presently observed value: co(z) scaling roughly as 00-0·38 and
dro/dz as 00-0.5 for z» 00-1(3. This means that differential tests between these models are best
camed out at z < 00- 1/3. Furthermore, as regards the subject of this conference, it should be noted
that in almost every case the two models consistent with inflation flank the most plausible
alternative, i.e. a low 00, k = -1 model.

1.2 THE FUNDAMENTAL PROBLEM OF GALAXY LUMINOSITY EVOLUTION

Whether dealing with individual galaxies, selected according to criteria that mayor may not be
well understood (e.g. in the m(z) or 9(z) test), or dealing with galaxy populations (e.g. in the
volume-redshift test) a knowledge of the properties that the galaxy or galaxies in question would
have at different redshifts is required. Since almost all galaxian properties locally are related to
luminosity (e.g. comoving number density, size, colour), this amounts to requiring knowledge of
the luminosity the galaxies concerned as a function of time. Galaxies are expected to evolve in
luminosity, and our present poor determination of the parameters controlling this process is the
biggest single impediment to the use of galaxies as cosmological probes.
Even for a co-evally produced "simple" stellar population (SSP), the luminosity evolution
may be large and is definitely uncertain. Unless the IMP is extremely steep, the luminosity of an
SSP at all wavelengths is dominated by stars that are at or have just left the main sequence turn-
off point. The colour of the population will reflect the colours of these stars and will be a function
primarily of age (and also metallicity etc.). The luminosity will mainly reflect the number of such
stars and will therefore depend strongly on both the IMP and the age. The IMP is hardly
constrained at all by the colours of the population and is also very hard to determine by any other
means. The classic Tinsley and Gunn (1978) relation 6M = (1.3 - O.3x) L\!n('t), where x is the
logarithmic slope of the IMP, is still approximately valid (see e.g. Charlot and Bruzual 1991).
More complex stellar systems will have a more complex luminosity evolution.
It is often incorrectly stated that luminosity evolution in the infrared waveband will be
"small" - the foregoing emphasizes that it can be substantial and, for as long as the IMF is poorly
determined outside of the solar neighbourhood, it will be highly uncertain. The ultraviolet/optical
wavebands will be sensitive to the short-lived effects of bursts of star-formation, and the
evolution in the infrared should at least be more gradual. Nevertheless, as seen in the models of
Arimoto and Yoshii (1987), a stellar population that is slowly built up over time (e.g. their Sed
galaxy), may actually have a larger (although negative) evolution in the infrared than in the
optical waveband.
235

This fundamental uncertainty plays havoc with every attempt to use galaxies as cosmological
probes. Even attempts to use the fonn of the luminosity function to locate a characteristic
luminosity (e.g. Loh and Spillar 1986) will fail if the differential luminosity evolution of different
classes of galaxies alters the shape of the luminosity function. The solution may be to use non-
galactic objects such as Type la supernovae on the m-z relation (see Ellis, these proceedings) or
the non-luminous parts of galaxies, e.g. through the effects of gravitational lensing (e.g. Fukugita
1990a, Turner 1990). Both pose fonnidable observational and/or intetpretational difficulties,
however.

2. The classical cosmological tests

2.1 THE TIME-REDSHIFT RELATION '1:(z)

Figure 1 shows the predicted 't(z) for the three models in tenns of the inverse Hubble constant
Ho- 1. There is a factor oftwo difference in the present age of the Universe and a factor of three
difference in the age at high redshifts.
0

N
.....,
-:'"a
2
....
't! 3
Q)
~

0 0.5 1
Epoch (110- 1)
Fig. 1: The age of the Universe as a fez) for [Qo-A] = [1,0] (solid), [0.1,0] (dashed), [0.1,0.9] (dot-<lashed].

Constraints on the absolute ages of radio galaxies at z > 3 may be derived from their colours.
For example, 4C41.17 at z = 3.8 has 't> 0.33 Gyrs (Chambers and Charlot 1990). These ages are
consistent with all of the three models, although they imply early fonnation epochs if Ho is large
or if [000.A] = [I,D]. It should be stressed that the debate (e.g. Lilly 1989, 1990, Chambers and
Charlot 1990, Chambers this volume) over the evolutionary status of these galaxies has little
bearing on this cosmological constraint (since this is ultimately set by the minimum age at the
highest redshift) but is extremely relevant for our ideas of galaxy fonnation.
An interesting differential timescale test is provided by the colour evolution of SSPs. The
differential rest-frame colour evolution over a modest (factor of 2) range of age is independent of
many of the uncertainties introduced in absolute age detenninations by rhings like the metallicity
and reddening, and by the mass-lifetime calibration of the main sequence (see e.g. Lilly 1987).
The models of Charlot and Bruzual (1991) give A(U-VJo = 1.16 [loglO ('t/'to)] while those of
Arimoto and Yoshii (1987) give a 15% shallower slope. The effects of any intennediate age
236

populations and of galaxy fonnation at fmite z will both be to give the appearance of an
accelerated colour evolution, and this anaylsis provides an upper bound to the differential age of
the Universe. The data of Lilly (1987) suggest that the (U-V)o colours of gE galaxies in rich
clusters have changed by only 0.15 ± 0.1 back to z = 0.45 implying a maximum value of 10glO
('tI'tfj) of 0.22-0.25 which may be compared with 0.24 for [1,0], 0.18 for [0.1,0] and 0.14 for
[0.1,0.9] . Note however that the expected value for [1,0] increases to > 0.29 for zF < 3. Even here,
however, there are still some second order corrections required to take account of the effects of
luminosity evolution on the colour-magnitude relation.

2.2 THE MAGNITUDE-REDSHIFr RELATION m(z)

The high z radio galaxies follow a remarkably tight K(z) relation in the infrared waveband. Data
from Lilly (1989) (together with new data obtained with Michael Rigler on very high z radio
galaxies from McCarthy et all991 and Chambers et al 1990) are shown in Figure 2 in the fonn
of a Reduced Absolute Magnitude (calculated in the [1,0] model and taking into account
aperture- and K-corrections). Note in passing that Hill and Lilly (1991) have shown that there is
no dependence of absolute magnitude with radio luminosity across four decades of radio power at
z= 0.5.

16
-4
18

20
-2
22

24
o
26~~~~0~.~1--~~~~1~~~~~~10· 0.1 1
redshift (z) Age (T/To)
Fig. 2 (left): K-z diagram for radio galaxies plotted as an equivalent visual Reduced Absolute Magnitude-
see text Cosmological curves are as in Fig I, and the short dashed curve is a polynomial fit to the data. R 1
and R4 show the average of nearby Abell clusters of Richness Classes 1 and 4.
Fig. 3 (right): The luminosity evolution that is required to match the observed K-z relation to the different
cosmologies assuming that the radio galaxies are evolving standard candles, as a f('r/'tcV assuming zF = 5.
Short dashed lines show predictions of luminosity evolution for x = 0, 1 and 1.35.

The natural interpretation of the small scatter is that the infrared light is dominated by a mature
stellar population at all epochs and support for this interpretation comes from a quantitative
analysis of the infrared morphologies (Rigler et al 1991). Assuming that the trend to brighter
absolute magnitudes at high z is a purely evolutionary effect, the required evolution (as a function
of 'efto assuming zF = 5) for the different cosmologies is shown in Figure 3, together with
predicted evolutionary paths for SSPs as a function of the slope of their lMF. Even the large
evolution required to accommodate the [0.1,0.9] model can be produced by plausibly flat IMFs
(Le. x "" 0), and indeed the straight line fonn of the evolution required for this geometry matches
the fonn of the predicted evolution. Note that rapid evolution of this fonn places stringent
237

constraints on the range of ages of the galaxies at a given z (see Lilly 1989). The K -z relation for
radio galaxies is thus consistent with a wide range of cosmologies even with these restrictive
assumptions about the nature of the radio galaxies (i.e. that they are indeed evolving "standard
candles".

2.3 TIlE ANGULAR SJZE-REDSHIFr RELATION 8(z)

Figure 4 shows the predicted angular size (for Sh-l kpc) for the three cosmologies. There has
been little systematic work on measuring the sizes of high redshift galaxies since the optical work
on radio galaxies of DjorgovsId and Spinrad (1981). The recent introduction of infrared arrays
makes the measurement of infrared sizes feasible and this may eventually give an interesting
constraint, particularly for the reddest objects which will have the smallest contamination from
the blue "aligned" component in the radio galaxies (see e.g. Rigler et al 1991). An attractive
feature of this test is the mimimum in angular size shown by both the flat models. Nevertheless,
since for elliptical galaxies, log re = -{).58 MB + constant (Konnendy 1980), we need to know the
zero-redshift luminosity to within O.S magnitude to tie down the expected size to within a factor
of two. This uncertainty in luminosity is small compared with the effects of luminosity evolution
seen on Figure 3. There are, of course, other evolutionary effects of a dynamical nature that could
alter the true size of a galaxy.
100 10 r----~-....__-.---.- ........"T'",.""I,..,
,.,.,.
...., ,.
o ..... /
~
-'" /

~
II
10

/
/~// /
1 //
;' /
i /
1/
//
0.1 1 10 10
redshift (z) (l+z)

Fig 4 (left): Angular size of a 5h- t kpc rod for the three cosmologies considered (curves as in previous
Figs).
Fig 5 (right): dN/dz for objects of constant physical cross-section (curves as before).

2.4 TIlE LINE-ELEMENT RELATION dN/dz

Figure S shows the predicted dN/dz for objects of constant physical cross-section. This test is
most appropriate for quasar absorption lines, which are outside of the scope of this paper. The
slope of dN/dz (parameterized as (1 +z)1) for MgII systems was initially thought to be r > 2
(Lanzetta et al 1987) but more recent analyses place it closer to the Einstein-de Sitter expectation
(Steidel, private communication). The line-element enters in a more complex way in the
238

calculation of the incidence of gravitational lensing. Turner (1990) has recently proposed that the
relatively low number of quasar gravitational lenses that have been discovered argues against a
A-dominated cosmology, but a revision of some of the critical input parameters and a re-
evaluation of the selection biases in this scheme have led to a substantial weakening of the
strength of this statement (Turner and Fukugita, in preparation).

2.5 THE VOLUME-ELEMENT RELATION dVJdz

The change in comoving volume element with redshifi for the different cosmological models is
shown in Figure 6. There are large effects between the two flat models at quite modest redshifts
(e.g. a factor of 2.7 in dV Jdz at z = 0.5) and at z = 2 they differ by approximately an order of
magnitude. There has recently been discussion as to whether the number-magnitude relations in
the Band K bands (Fukugita et all990b, Cowie et all991, respectively) constrain the geometry,
particularly as redshifi information for B-band selected samples is now available down to B =
24.2 (see Broadhurst et al1989, Colless et all990, Lilly et al 1991).

.,.
":9 0.1
~
......
(.I

~ 0.01

>
og
0.001

0
0.000 .01 0.1 1 10
redshift (z)
Fig 6: dVcldz lor the three cosmologies considered (cwves as before)

The fundamental difficulty is again galaxy evolution and the likelihood that the galaxy
luminosity function will change not only in normalisation (Le. in L *) but also in shape if
different luminosity galaxies of different masses and/or Hubble types have different evolution.
This is illustrated in Figure 6, in which the redshifi distribution for 23 < B < 24 is plotted for (i)
the [1,0] and (ii) [0.1,0.9] cosmological models with no galaxy evolution, and then for the [1,0]
model with various generic alterations to the LF - (iii) an increase in L * with look-back time, (iv)
an increase in C\>* with look-back time (given the limited range of luminosity sampled by a
magnitude limited sample, this is equivalent to evolutionary models that produce a piling up of
galaxies around L*, see e.g. the bursting model of Broadhurst et al1989) and (v) a coupling ofL*
and C\>* (C\>*L* = constant) so as to mimic the effects of merging. The parameters in models (iii)
and (iv) have been adjusted to produce the same total number of galaxies as in (ii). It should be
noted that the merging scenario (v) is by itself extremely inefficient at raising the number of
galaxies (see e.g. Lilly et al 1991 for a discussion). Even with full redshifi information, it is
clearly very difficult to disentangle cosmology from various combinations of evolutionary
alterations to the luminosity function.
239

N
l1li
.2 3000
.s
e....
~ 2000
CD
o
~
-:: 1000
...
{
redshift (z)
Fig 7: The predicted redshift distribution for 23 < B < 24 calculated for an unevolving [1,0] and [0.1,0.9]
models, and for the [1,0] model with evolution in L*, in ~*, and in both ~* and L* (such that ~*L* =
constant) to represent a merging model (see text for details).

In the B-band, various groups have found that the median redshift increases only slowly with
depth and tracks the prediction for an unevolving galaxy population despite the increasingly large
excess of faint galaxies over the predictions of such models (Broadhurst et al1989, Colless et al
1990, Lilly et al 1991). At B = 24, the median <z> = 0.4 (Lilly et al 1991) despite a threefold
enhancement in the number of galaxies. This was interpreted as reflecting either (a) Broadhurst-
type evolution of the luminosity function (we are seeing large numbers of dwarf galaxies boosted
up to roughly L* luminosities thereby effectively increasing ,*) or (b) a A-dominated
cosmological volume element. If maintained over cosmological timescales, as might be required
to explain the counts at B = 26.5, the first explanation would require the dwarfs to have a highly
efficient mechanism for the removal of metals or for them to have not been counted in the
present-day Schecter galaxy luminosity function (see e.g. Lillv et all991)'
10?r-~--r--r--~-.--~~--~~--,-~~'

E'IOWTlON 1YT88) with .. - 10


•• - 0.001011. B·.. - -lUI
au. aaIuJ adz

10

115 20 25
K.. magnitude
Fig 8: The K-band counts of Cowie et al (1991) compared with simple evolving models of the galaxy
population for the three cosmologies considered (curves as in Figs 1-6).
240

In the K-band, the counts are closer to the predictions of non-evolving galaxy populations in
standard cosmologies (Cowie et all990, 1991). The blue colours thus implied are suggestive of
the evolutionary scheme above, and this has led Cowie et al (1991) to rule out A-dominated
cosmologies. However, given the large uncertainties concerning the galaxy luminosity function
both locally (e.g. Efstathiou et al 1988) and at earlier times, this conclusion may be premature.
Figure 7 shows the K-band counts of Cowie et al (1991) and an extremely simple evolving model

,*
(based on the Arimoto and Yoshii evolutionary models) in the three cosmologies in which the
local value of has been reduced to 0.001 Mpc-3 (the value derived from the DARS - Efstathiou
et a1 1988). Clearly it is possible to find scenarios in which 00 1:- 1 cosmologies are consistent

,*
with the K-band counts, and A 1:- 0 will be acceptable if one further changes the shape of the
luminosity function so as to decrease the effective further.

3. Conclusions

The fundamental uncertainties in galactic luminosity evolution make almost all attempts to use
galaxies at high redshift to constrain the cosmological parameters extremely difficult. The most
positive thing that can be said is that all of the plausible models, including those with significant
A-tenos, can be made consistent with the available data.

References

Arimoto, N., and Yoshii, Y.; 1987, Astr.Ap., 173,23.


Broadhurst, TJ., Ellis, R.S., Shanks, T.; 1988, MNRAS, 235,827.
Chambers, K.C., Charlot, S.; 1990, ApJ.Lett., 348, L1.
Chambers, K.C., Miley, G.K., van Breugel, W.; 1990, ApJ., 363, 21.
Charlot, S., Bruzual, G.B.A.; 1991, 367, 126.
Colless, M.M., Ellis, R.S., Taylor, K., Hook, R.N.; 1989, MNRAS, 244,408.
Cowie, L.L., Gardner, I.P., Lilly, S.l, and McLean, I.S.; 1990, ApJ.Lett, 360, Ll.
Cowie, L.L., Gardner, I.P., Wainscoat, R., Hodapp,K-W.; 1991, preprint.
Djorgovski, S., Soinrad, H.; 1981, ApJ., 251, 457.
Efstathiou, G., Ellis, R.S., Peterson, B.; 1988, MNRAS, 232, 431.
Fukugita, M., Futamase, T., Kasai, M.; 1990a,preprint.
Fukugita, M., Takahara, F., Yamashita, K, Yoshii, Y.; 1990b,preprint.
Hill, G.I., Lilly, S.l.; 1991, Ap.l., 367, 1.
Konoendy, I; 1980, In "ESO Workshop on 2-dimensional Photometry" ed. P. Crane and K. Kjar
(Geneva: ESO), p.191.
Lanzetta, K., turnshek, D., Wolfe, A.; 1987, ApJ., 322, 739.
Lilly, SJ.; 1987, MNRAS, 229, 573.
Lilly, S.l; 1989, ApJ., 340, 77.
Lilly, S.l.; 1990, In 'The Evolution of the Universe of Galaxies", (ed. R.G. Kron),
Astron.Soc.Pacific, p. 344.
Lilly, S.l., Cowie, L.L., Gardner, IP.; 1991, ApJ., 369, 79.
Loh, E., Spillar, E.; 1986, Ap.I.Lett., 307, Ll.
McCarthy, PJ., Kapahi, V.K., van Breugel, W., Subrahmanya, C.R.; 1991, ApJ. in press.
Peebles, P.I.E.; 1984, Ap.l., 284,439.
Sandage, A., 1988, Ann.Rev.Astr.Astrophys., 26, 561.
Turner, E.; 1990, preprint.
241

DISCUSSION:

Hoyle: The redshift-magnitude relation with corrected magnitudes, has had a long history
of remaining strictly linear. First, Hubble found it straight to z = 0.2. Then Sandage
extended the line to z = 0.5. Now you have it still straight to z = 2, or more. Isn't there
a case for taking the evidence at its full value, namely that qo = + 1? It seems rather
artificial to correct for evolution or whatever in order to bring the relation to qo = 0.5.
Admittedly, this would be hard for the theoretician to swallow, but perhaps this is the way
it is.
Lilly: An important point. Nevertheless, I think it should be remembered that straight-
forward consideration of the stellar initial mass function suggests that the red-visuallumi-
nosities of early-type galaxies will change with age, and will probably get brighter at earlier
times.
Peacock: Some people have suggested that the morphologies of hi -z 3CR galaxies derive
from lensing. Can you set any constraints on that from your data?
Lilly: In this sample at z '" 1, the infrared morphologies are generally simple, with the
main symmetric galaxy and the aligned component showing weakly through. We do see
faint red companions in many cases but interpret these as conventional "trappers" for the
radio source activity.
I would be surprised if lensing was significant.

Ellis: Older cosmologists would remind us that A f:. 0 has frequently been invoked to
interpret observational data that cannot otherwise be understood. The outcome was usually
a new astrophysical understanding of the source. Do you share my view that this 'A -phase'
may not last very long and the outcome will be exciting anyway?
Lilly: Yes, that is probably right. My aim in this talk has simply been to say that several
rather independent astrophysical affects seen at high redshifts may easily be accommodated
into a A f:. 0 cosmology. I do not take these arguments as strong evidence in favour of
adopting A f:. o.
PROSPECTS FOR MEASURING THE DECELERATION PARAME-
TER

llichard S. Ellis
Department of Physics,
University of Durham,
South Road,
Durham DHl 3LE.

1 Introduction

Tests of inflation, via measurements of the cosmological density parameter, no, have pri-
marily used gravitational methods and local galaxy samples (Davis & Peebles 1983, Bean
et al1983, Rowan-Robinson et al1990). Conflicting results have emerged and the extent
to which such samples may be biased is unclear. Previously, observers attempted to mea-
sure the geometrical deceleration parameter, qo. The strategy changed when it was realised
that the sources traditionally used as probes have probably evolved significantly over the
look-back times concerned (Tinsley 1968, Lilly & Longair 1985).
Here I aim to revive qo tests with a fresh approach. qo is important to measure indepen-
dently of local no estimates for several reasons. Foremost, inflation requires the Universe to
be spatially fiat, i.e. k =0 which can be reconciled with low estimates of the baryonic mass
density, 0.009< nbO h2 <0.04 (Pagel 1990), if the cosmological constant, A, is 3 (I-no) H~
(Peebles 1984). Although somewhat contrived, this might resolve the globular cluster age
conflict, recent galaxy count/redshift data and large scale structure problems, each of which
received prominent discussion at this workshop. Furthermore, regardless of whether A #0,
deceleration tests make no assumptions as to the expected bias in the sources used. The
test is a geometric one involving relative distances and is usually independent of Ho.

2 Deceleration Tests Cannot be Based on Galaxies

The classical qo tests using first-ranked cluster galaxies have been reviewed by Kron (1984),
Tammann (1984) and Sandage (1988). Redshift-dependent biases are minimised if the
sources are 'special objects' of ~constant luminosity irrespective of the detected proportion
of the galaxy luminosity function. Bhavsar (1990) finds ~65% of Hoessel et al's (1980)
photometric sample to be in this category with an intrinsic scatter of ±0.m2. Unfortunately,
that subset cannot be readily identified from the larger set with dispersion ±0.m35, so
photometry of >20 clusters at z ~0.75 is needed to get qo to ±0.1. Currently, 13 examples
are known with z>0.5 from the two objective catalogues of distant clusters (Gunn et al
1986, Couch et al1991).
Models predict that the effects of passive luminosity evolution (backtracking the H-R
diagram of present-day stellar populations) dominates the Hubble diagram at both optical
243
T. Shanks et al. (eds.), Observational Tests ojCosmological Inflation, 243-249.
© 1991 Kluwer Academic Publishers.
244

and infrared wavelengths (Figure 1). Further complications occur if the evolutionary history
varies from cluster to cluster; for example, if there is enhanced star formation in some
clusters (Dressler & Gunn 1990). The increased qo sensitivity for more distant radio galaxies
(:::100 sources to z:::4) is similarly diluted by larger expected evolutionary changes (Lilly
1990). The Hubble diagram is thus an ineffective probe of qo unless the evolutionary changes
are known to high precision and are shared by the entire population under study.

1
I 1_____ - - -q~
--
III
SNla BeG -
GI
"C

---------- ----
.5 qo=O.25
~
l1li
..;::::::::-"::::=--- -
cs
a 0

Sl
~ -.5
--- --- - - - -
'-'
<I
-1

o .2 .4 .6 .8 1 1.2 1.4 1.6 1.8 2


Redshift

Figure 1: Change in the apparent brightness (magnitudes) of a standard candle, nor-


malised to the qo=O.5 model, as a function of qo (A=O). Curves show the ezpected change
in V and K (~J.'m) for a passively-evolving elliptical formed at zF=5 (Ho=50) calculated
using Bruzual's (1983) code.

An alternative approach uses the surface density of faint field galaxies as a volume test
for qo (Yoshii & Takahara 1988, Koo 1990). The local galaxy density (Efstathiou et al
1989) and the abundance offaint (B rv 27) galaxies (Tyson 1988) suggest large volumes and
hence a low qo, almost regardless of the maximum redshift to which viewing is possible.
However, the test does assume galaxies are conserved in number. Statistically-complete
redshift surveys to fixed apparent magnitude limits (Broadhurst et a11988, Colless et al
1990) reveal a gradual increase in the comoving number density of visible galaxies to z:::0.5
suggesting recent mergers and/or bursts of star formation as an alternative explanation.
Similar redshift tredns are seen for !RAS galaxies whose infrared emission is known to
predominantly arise from interactions (Saunders et aI1990).
If the excess faint optical population is largely identified with galaxies showing strong
[0 II] equivalent widths, are those objects preferentially interacting? High resolution NTT
images show some convincing examples but the samples are currently too small for a conclu-
sive statement. However, even a minor amount of merging would compromise the volume
test significantly and imply a luminosity function which evolves in shape, rendering volume
element tests as used by Loh (1988) ineffective, even spectroscopic rather than less-precise
photometric, redshifts are used. Evidently, neither cluster nor field galaxies are effective
tracers of qo.
245

3 Use of Supernovae of Type Ia

Unlike galaxies, supernovae of type Ia (SN Ia) are thermodynamic events not subsets of a
statistical population whose homogeneity depends critically on survey tactics. A SN Ia is
either detected or not. When seen, its peak luminosity is governed by the Ni 56 core mass
and local data demonstrates this to be exceptionally well-defined to better than ±0.m3,
and possibly as small as ±0."'19 (Leibundgut 1991). Although SNe Ia can be seen to high
redshift, their role in qo tests has been minimal because of the effort needed to find them.
This problem can now be overcome with efficient CCDs and modest-aperture telescopes.

22

23

24

25

26U-L-~~~~~~~~~~~~~-J-J-J~~

-20 o 20 40 60 80
REST-FRAME DATE (days)

Figure 2: Rest-frame B light curve of the z=0.31 AC 118 SN Ia (points) uncorrected for
extinction, compared with a standard light curve for a SN Ia (solid line). t=O is chosen at
the estimated epoch of maximum light (after N¢rgaard-Nielsen et a11989)

To demonstrate the feasibility of finding high z SNe, N!1lrgaard-Nielsen, J!1lrgensen,


Hansen, Couch & I undertook a systematic monitoring of 65 clusters with z ~0.3 over
1986-89 using a 1.5-m telescope in Chile. A cluster search maximises the detection rate (for
a fixed field of view), minimises internal reddening problems (since the host galaxies are
mostly dust-free early-types), and ensures the data can be used for other projects e.g. grav-
itationallens searches. In ~200 comparisons capable of detecting excess light to V =23.5,
two SNe were found and studied in detail- a SN II at z =0.28 (Hansen et a11989) and a
SN Ia at z=0.31 (N!1lrgaard-Nielsen et aI1989). A further 1-3 likely SNe were detected but
deemed too marginal to follow up.
The rest-frame B light curve for the z=0.31 SN Ia matches the composite curve derived
from 16 local SNe Ia (Cadonau et al1985) when time-dilation effects are included (Figure
246

2). The event was unfortunately detected after maximum light which limits the precision
with which its peak brightness can be determined. However, in those 40% of SNe to z~0.5
found pre-mazimum in such a survey, the peak brightness could be determined to a precision
limited solely by reddening corrections and photometric errors. Both effects should not add
significantly to the intrinsic scatter if the events occur primarily in early-type galaxies.
To determine qo to ±0.1 then requires 35 SNe Ia at z=0.3, or 10 at z=0.5, each seen
pre-maximum. Taking our own uncertain SNe Ia rate for clusters 0.2 h 2 SNe Ia (10 10 LB0
century)-l, lower than some recent estimates, and allowing for those seen post-maximum,
indicates ~1000-3000 CCD hrs on a 2.5-m telescope is required (depending on z) to find
enough SNe Ia for useful follow-up work. There are enough equatorial clusters with 0.2<
z <0.5 to mount an intensive campaign with a 2.5-m telescope whose dark time programme
be dedicated to this task for 2 years. Such an instrument would be powerful enough to
both survey clusters and construct light curves of detected SNe to determine maximum
light. Spectroscopy would require override time on 4-m telescopes in either hemisphere.
Tammann (1984) estimates >300 nights of 5-m dark time was spent pursuing the cluster
galaxy Hubble diagram and no satisfactory conclusion on qo emerged. The local proof
and physical understanding of the homogeneity of SNe Ia as events, not objects, and the
clear demonstration that technology can now identify SNe Ia at cosmologically significant
distances suggests inflation can be tested directly with a more modest sacrifice of 2-3 m
telescope time provided the organisational difficulties can be overcome.

References

Bean A.J., Efstathiou, G., Ellis, R.S., Peterson, B.A. Shanks, T., 1983. Mon. Not. R. astr. Soc,
205,605.
Bhavsar, S., 1990. in Hubble Centennial Symp., p. 209, ed. Kron, R., ASP Series.
Broadhurst, T.J., Ellis, R.S. &; Shanks, T., 1988. Mon. Not. R. astr. Soc, 235,827.
Bruzual G., 1983. Astrophys. J., 273, 105.
Cadonau, R., Sandage, A. &; Tammann, G.A., 1985. in Supernovae as Distance Indicators, p. 151,
ed. Bartel, N., Springer-Verlag.
Colless, M.M., Ellis, R.S., Taylor, K. &; Hook, R.N., 1990. Mon. Not. R. astr. Soc, 244,408.
Couch, W.J., Ellis, R.S., Malin, D.F. &; MacLaren,!., 1991. Mon. Not. R. astr. Soc, , in press.
Davis, M. &; Peebles, P.J.E., 1983. Astrophys. J., 267, 465.
Dressler, A. &; Gunn, J.E., 1990. in Hubble Centennial Symp., p. 200, ed. Kron, R., ASP Series.
Efstathiou, G., Ellis, R.S. &; Peterson, B.A., 1989. Mon. Not. R. astr. Soc, 232, 431.
Gunn, J.E., Hoessel, J. &; Oke, J.B., 1986. Astrophys. J., 306, 30.
Hansen, L., Nl1lrgaard-Nielsen, H.U., Jl1lrgensen, H.E., Ellis, R.S. &; Couch, W.J., 1989. Astr. As-
trophys., 211, L9.
Hoessel, J., Gunn, J.E. &; Thuan, T.X., 1980. Astrophys. J., 241, 486.
Koo, D.C., 1990. in Hubble Centennial Symp., p. 268, ed. Kron, R., ASP Series.
247

Kron, R., 1984. in Spectral Evolution of Galazies, p. , ed. Gondhalekar, P., RAL Publ.
Leibundgut, B., 1991. in Supernovae, p. 751, ed. Woosley, S., Springer.
Lilly, S.J., 1990. in Hubble Centennial Symp., p. 344, ed. Kron, R., ASP Series.
Lilly, S.J. & Longair, M.S., 1984. Mon. Not. R. astr. Soc, 211, 833.
Loh, E., 1988. Astrophys. J., 329, 24.
Nfllrgaard-Nielsen, H.U., Hansen, L., Jfllrgensen, H.E., Aragon Salamanca, A., Ellis, R.S. & Couch,
W.J., 1989. Nature, 339, 523.
Pagel, B.E.J., 1990. in Baryonic Dark Matter, p. 237, ed. Lynden-Bell, D. & Gilmore, G., Kluwer.
Peebles, P.J.E., 1984. Astrophys. J., 284, 439.
Rowan-Robinson, M., Lawrence, A.J., Saunders, W., Crawford, J., Ellis, R.S., Ftenk, C.S., Parry,
I., Xiaoyang, X., Allington-Smith, J., Efstathiou, G. & Kaiser, N., 1990. Mon. Not. R. astr. Soc,
247, 1.
Sandage, A., 1988. Ann. Rev., 26, 561.
Saunders W., Rowan-Robinson, M., Lawrence, A., Efstathiou, G., Kaiser, N., Ellis, R.S. & Ftenk,
C.S., 1990. Mon. Not. R. astr. Soc, 242, 318.
Tammann, G.A., 1984. in Clusters of Galazies, p. , ed. Mardirossian, F. et al, Reidel.
Tinsley, B.M., 1968. Astrophys. J., 151, 547.
Tyson, A.J., 1988. Astron. J., 96, 1.
Yoshii Y. & Takahara, F., 1988. Astrophys. J., 326, 1.
248

DISCUSSION:

Renzini: Could you comment on the type Ia supernova rate that you infer from your
survey?
Ellis: The formal SN la rate from our distant cluster survey is comparable to that quoted
by Van den Bergh et al (Ap. J. 323, 44, 1987) but a factor of 4 below that of Tammann
(in 'Supernovae' ed Rees, M.J, Stoneham, R.J. p 371, 1982) - However, since both Van
den Bergh et al and ourselves only detect 1 event in EjSO galaxies, as opposed to 19 in
Tammann's survey ( and 8 in Capellaro and Turatto's (Astr. Astrophys. 196, 10, 1988)
compilation) the distant rate remains highly uncertain. Of course, our pilot programme
was planned assuming the Tammann rate would be the minimum applicable - one might
expect SNe to be more common in the past for various reasons - but, as I explained earlier,
I cannot believe we have missed ~ of the SNe la. Larger samples are required!
Peebles: What is your opinion of the astronomical reasonableness of the merging hypoth-
esis for the deep galaxy counts? H the remnants are spirals can we understand the thin
discs? H the remnants are not spirals what might they be?
Ellis: I can only repeat the evidence. At bi = 21.5 we find the 30 - 50 % excess counts
matches well the size of a new population of galaxies with spectacular [0 II] emission.
Coadding representative spectra reveals spectral features indicative of a sudden increase
in star formation. High resolution imagery suggests many of these galaxies have complex
structures as expected for the merging hypothesis. Whether this is sufficient to explain the
fainter counts is unclear. Remember a low mass satellite sparked into life by a burst of star
formation would look quite respectable in the rest-frame UV at z 0.2. One should not
f'J

convert contours of UV light at the burst peak into contours of mass in the quiescent phase.
I think it more profitable to explore this area of astrophysics than shrug my shoulders and
invoke A.
Cannon: With reference to your relatively low rate of discovery of supernovae, is there
a selection effect whereby you find supernovae more easily in isolated small galaxies where
they dominate the total light, and miss those that occur in giant galaxies which are swal-
lowing dwarfs?

Ellis: We looked at this possibility very carefully with simulations and found no problems
(Hansen et al ESO Messenger 47, 46,1987). I should stress that SNe are found by virtue of
the excess light they produce - we do not rely on resolving them within galaxies (although,
astonishingly, with the WHT we managed to do this for the one found at z = 0.31). In fact
we don't even have to detect the host galaxy. The only possible selection effect that would
reduce the number found would be the increased noise associated with brighter galaxies in
the cluster. Here we are talking about a very small surface brightness range and simulations
show the variation in noise to be negligible in clusters with z > 0.2.
Lilly: Do you see the effects of time dilation?
249

Ellis: Yes we do. The decline rate, /3, of the light wave between maximum and the inflec-
tion point for SNe la has a narrow distribution locally. Without time dilation correction
the z = 0.31 SN has a highly atypical /3; with time dilation the value is close to the median
of the local /3 distribution for Type la.
Pierce: Do you take multi-wavelength data in order to estimate colours and hence extinc-
tions for any detected supernovae?
Ellis: The search was conducted in a single passband (V) to maximise the detection of
SNe. Once a SN is found, data is secured in V .!!.llil R using over-rides on other telescopes.
Ideally it would be desirable to get as much colour information as possible but practicalities
of over-rides make this difficult to achieve.
ALIGNED RADIO GALAXIES

K. C. CHAMBERS
Sterrewacht,
Postbus 9519,
2900 RA Leiden,
The Netherlands.

ABSTRACT. The discovery of high redshift radio galaxies at z ~ 4 has unleashed a host of questions
about their nature. Many new objects are being found, and their unusual properties are being
uncovered. They generally exhibit the "alignment effect", wherein their highly elongated lumpy
morphologies are lined up along the axes of their powerful radio sources. Evidence for both large
star formation rates and scattered quasar light are found in the aligned blue continuum. The aligned
infrared continuum is more likely to be dominated by starlight. The K-band Hubble diagram for
radio galaxies is now extended to z ~ 4, but it will be difficult to use it to test different cosmological
world models.

1 Introduction

Our perceptions of high redshift radio galaxies have changed dramatically in the last three
years. In the early eighties, high redshift (1 < z < 2) radio galaxies were viewed as natural
extensions of their low redshift counterparts - generally rather normal giant ellipticals. This
of course made them attractive objects for studies of galaxy evolution and ultimately, it
was hoped, for use in distinguishing between cosmological world models (e.g. Spinrad 1985,
Lilly and Longair 1894). However, in the late eighties a series of new discoveries and new
ideas has forced us to recognize these objects as having extraordinary properties which
are not understood. The first of these was the discovery by Chambers et al. (1987) and
McCarthy et al. (1987) of the "alignment effect": at high redshift, radio galaxies often
have highly elongated optical continuum morphologies, and these optical extensions are
reasonably well lined up with the axis of the powerful radio source. At about the same
time the long standing "z = 2" barrier for radio galaxies was broken with the discovery of
4C40.36 at z = 2.267 by Chambers, Miley and van Breguel (1988) using a new technique of
selecting candidates by their ultra steep radio spectrum. The discovery of higher redshift
sources quickly followed: Lilly (1988) discovered the "1 Jy" source 0902+34 at z = 3.4, and
Chambers, Miley and van Bruegel (1990) found the ultra steep spectrum source 4C41.17
at z = 3.8, which at this time is still the most distant known radio galaxy.
Most work on distant radio galaxies has centered on various complete samples from
flux-limited surveys. However, given the very large number of radio sources which become

251
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 251-256.
© 1991 Kluwer Academic Publishers.
252

avaliable at fainter and fainter flux limits, it is desirable to develop techniques for pre-
selecting the best candidates from such surveys. Our ultra steep radio spectrum method
is based on the observed correlation between the radio spectra and luminosities of radio
sources (Heeschen 1960) and the fact that the fraction of 4C radio sources which had
identifications on the Palomar Sky Survey was a strong function of the radio spectral index
(Blumenthal and Miley 1979).
Based on this, we have begun an extensive radio, optical, and infrared investigation
of these sources (Chambers, Miley, van Breugel 1987, Chambers, 1989). Our ultra steep
spectrum sample comprises 4C sources known to have spectral indices of 0 < -1, between
178 and 5000 MHz, where 5" = kv o. The current status of our 4C ultra-steep spectrum
survey is encouraging. Out of 33 4C ultra-steep spectrum sources we have obtained 31
optical identifications, attempted optical spectroscopy of 32, found 16 with strong emission
lines and z > 0.5, 8 of these have z > 2.
We are currently extending our ultra-steep spectrum survey to fainter flux levels us-
ing sources from the Parkes, Texas, Molonglo, 6C, and 8C surveys (Miley et al. 1989).
Preliminary results on a subsample of 120 objects indicate 76 of these are aligned, 14 are
unresolved, 18 appear to be intermediate redshift rich clusters, 7 are weak identifications,
and 5 are blank fields. Our first spectroscopic run indicates that at least 7 of the aligned
objects are new z > 2 radio galaxies.
Other groups pursuing new high redshift radio galaxies are working from samples from
a variety of surveys, including: the Parkes selected areas, The MIT-Greenbank survey,
The Molonglo survey, the B3 survey, and miscellaneous other samples (e.g. Spinrad 1990,
McCarthy et al. 1990). At the moment, there are presently known about 65 radio galaxies
with 1 < z < 2, 21 with 2 < z < 3 and 6 with z > 3. The numbers are sure to increase
rapidly in the next few years.

2 The Alignment Effect

High red shift radio galaxies generally exhibit the "alignment effect" wherein their elongated
lumpy morphologies are lined up along the axes of their powerful radio sources (Chambers,
Miley, van Bruegel 1987, McCarthy et al. 1987). At redshifts above 1 this phenomena
becomes very common, perhaps extending down to low power radio sources (Windhorst
et al. 1991). Although there are exceptions (objects which are elongated but not lined up
with the radio axis, and objects which are round), in general "high redshift radio galaxies"
are "aligned radio galaxies". There are a few examples at redshifts '" 0.5, but there are no
equivalent objects at low redshift.
The origin of the blue continuum light is not clear. There is marginal evidence that
the UV continuum is mostly starlight (Chambers and McCarthy 1990), but there is also
evidence for scattered quasar light as indicated by the discovery of significant polarization
in 3C368 (di Serego Alighieri et al. 1989, Scarrot, Rolph, and Tadhunter 1990). A scattered
component would be consistent with unified schemes of AGN's (e.g. Barthel 1989). In this
picture the differences between radio loud quasars and radio galaxies are attributed entirely
to their orientation with respect to the observer, i.e. a quasar is a radio galaxy seen near
to the radio axis. In radio galaxies the nucleus is assumed to be obscured along our line of
253

sight but is beamed anisotropically in the plane of the sky and then scattered into our line
of sight by some scattering medium associated with the host galaxy. The scattering could
be due to dust or perhaps more likely, electron scattering from a dense hot ICM (Fabian
1989).
In the near infrared it is much harder to avoid the conclusion that the light is dominated
by starlight. Dust scattering would become very inefficient, and electron scattering would
require that the scattered light have the same color as the quasar. Most radio galaxies are
redder than quasar continua, and the red flux required to make up the difference is consistent
with what one would expect from continuity arguments (unambigous radio galaxies are seen
up to redshifts z 1, they should have counterparts at higher redshifts). Lilly (1989) has
I'V

emphasized that the continuity and low dispersion of the K-band Hubble diagram for radio
galaxies implies that starlight dominates the K-band measurements.
Interestingly, the alignment effect extends down into the infrared (or well below the 4000
A break in the restframe, Chambers, Miley, and Joyce 1988, Eisenhardt and Chokshi 1990).
In our 4C sample of ultra steep spectrum radio galaxies, most show infrared alignments.
Thus for aligned radio galaxies, some reasonable fraction of the light in the K-band Hubble
diagram is aligned with the radio source. This is difficult to understand if the K band light
is dominated by an old stellar population.

3 The Ages of Aligned Radio Galaxies

The spectral energy distributions of aligned radio galaxies have two features which resemble
galaxies and which also suggest that the star formation in these objects was higher in the
past. First, most high red shift radio galaxies appear to have significant 4000A breaks (Lilly
1988, Chambers and Charlot 1990). The magnitude of the 4000A break is a rough measure
of the ratio of current to past star formation. If the galaxy were dominated by the current
star formation, (Le. the hypothetical primeval galaxy) the SED would be roughly flat in
lv, with little or no break at 4000 A . Since they do have a break, then there must have
been a time in the past when the star formation rate was higher. Second, all of the radio
galaxies studied have a significant turndown in the rest frame ultraviolet (fv) at ~ 1500
A (Chambers and McCarthy 1990). For any reasonable initial mass function this feature
suggests the star formation rate is decreasing - otherwise the ultraviolet would be dominated
by hotter stars. (This sharp turndown is also hard to reconcile with a nonthermal origin
for the UV.)
The crucial question is when in past was the star formation higher? Lilly (1988) has
suggested that the bulk of the star formation in 0902+34 was at a much earlier epoch and
unrelated to the current burst. Lilly has also emphasized that in a z 1 sample there are
I'V

some objects which are aligned in the blue but have a rounder morphology in the infrared.
If this is due to a ongoing starburst superimposed on a an underlying older giant elliptical,
then from the turndown in the rest frame ultraviolet one must also accept a highly truncated
IMF (M :5 3M0 ) for the current starburst. And if the infrared light is dominated by an
older stellar population, then at best there should be a slight 'anti' alignment (Le. lined up
with the minor axis) as there is at low redshift (Palimaka et al. 1979, Guthrie 1979).
But many of the aligned radio galaxies are aligned in the infrared (Chambers, Miley,
254

and Joyce 1988, Eisenhardt and Chokshi 1990, Chambers and Miley 1991). Chambers and
Charlot (1990) pointed out that an older population is not required to fit the 4000 A break
of these aligned objects. For a normal IMF, the ultraviolet turndown is well fit by a stellar
population a few times 10 8 years old, and the star formation must have been much higher
in the immediate past 10 8 years. But this is just what is required to have built up a
significant red population and provide a 4000 A break. If the star formation is occuring on
such a relatively short timescale, expanding in a deflagration from the center of the galaxy
outwards with the radio source, then one would expect the outermost regions to be the
bluest and show the strongest alignment. The data are therefore consistent with the idea
that the alignment effect is due (at least in part) to star formation triggered by the radio
source (Rees 1989, Begelman and Cioffi 1989, DeYoung 1989, Daly 1990).
Note that there is not much time available in an inflationary universe for these objects to
= =
have produced an older population. For Ho 50 km seCl Mpc l and no 1, the universe
is only 1.25 Gyr old at the redshift of 4C41.17 (z = 3.8). A burst lasting"" 1 Gyr would
have to have begun at very high redshift, and we would still be seeing its effects. If one
is willing to accept a shorter burst, then one is approaching the young stellar population
scenario anyway.

4 Extending the K-Band Hubble Diagram to z "" 4

With the discovery of 4C41.17 and 0902+34 the infrared K band Hubble diagram has now
been extended to nearly z ~ 4 (Lilly 1989, Chambers, Miley van Breugel1990, Chambers,
Charlot and Miley 1991). Note that this is as far as will be useful, since for z > 4 the
K band samples short ward of the 4000 A break and will be much too sensitive to current
star formation. With a Hubble diagram extending over 3 orders of magnitude in redshift,
is there any hope of using it to measure no? Are radio galaxies, particularly aligned radio
galaxies, useful standard candles?
The advantages of using high redshift radio galaxies as standard candles include: they
can be found at cosmological distances, thus they provide sufficient "leverage" on different
cosmological world models; they sample most of the volume of the visible universe; the
radio source selection finds an otherwise "normal" giant elliptical (at least at low redshift);
they provide a sample of objects selected independently of the their optical properties; and
the K-band Hubble diagram looks so good it must be useful for something!
The disadvantages of using high redshift radio galaxies as standard candles include:
there are presumably strong evolutionary effects at such large lookback times and there is
no unique way to deconvolve luminosity evolution from the effects of different no; radio
selection has its own problems, particulary the correlation between luminosity and redshift
in flux limited samples; we now know there is a relationship between the radio and opti-
cal/infrared properties, but it is not the one we expected (morphology rather than absolute
luminosity); and there is contamination by nonstellar light, ego scattering of quasar light.
Given the disadvantages one wonders if it is worth plotting the Hubble diagram, but
the fact is it looks very nice, with good continuity and low dispersion (Lilly 1989, Cham-
bers, Charlot, and Miley 1991). The linearity is fortuitous, since there are intrinsically
curved effects in the diagram, such as the curve of growth for a given aperature, luminos-
255

ity evolution, dynamical evolution, and the K-correction. Nonetheless all these effects can
be modeled (though not uniquely), and at the moment we find that for a normal (Scalo)
IMF, the high redshift objects fall along reasonable evolutionary tracks centered on the low
redshift objects. This implies a relatively low dispersion in the mass of the objects, or a
selection bias which is very sensitive to mass. When the current crop of high redshift radio
galaxies have measured K-band magnitudes, it should be possible to determine if the dis-
persion is increasing at high redshift. From the present measurments there is no convincing
determination of no.

References

Barthel, P.D., 1989. Astrophys. J., 336, 606.


Begelman, M. C., & Cioffi, D. F., 1989. Astrophys. J. Lett., 345, L21.
Blumenthal, G., & Miley, G.K., 1979. Astr. Astrophys., 80, 13.
Chambers, K. C.,1989. Ph.D. Thesis, The Johns Hopkins University.
Chambers, K. C., & Charlot, S., 1990. Astrophys. J. Lett., 348, L1.
Chambers, K. C., Charlot, S., & Miley, G. K. 1991. in preparation.
Chambers, K. C., & McCarthy, P.J., 1990. Astrophys. J. Lett., 354, L9.
Chambers, K. C., & Miley, G.K., 1989. in The Evolution of the Universe of Galaxies, ed. R.
G. Kron, (San Francisco: Astro. Society of the Pacific).
Chambers, K. C., Miley, G.K., & Joyce, R. R., 1988b. Astrophys. J. Lett., 329, L75.
Chambers, K. C., Miley, G.K., & van Breugel, W., 1987. Nature, 329, 604.
Chambers, K. C., Miley, G.K., & van Breugel, W., 1990. Astrophys. J., 363, 21.
Chambers, K. C., & Miley, G.K., 199Lin preparation.
Daly, R., 1990. Astrophys. J., 355, 416.
DeYoung, D., 1989. Astrophys. J. Lett., 342, L59.
di Serego Alighieri, S., Fosbury, R. A. E., Quinn, P. J., & Tadhunter, C. N., 1989. Nature, 341,
307.
Eisenhardt, P., & Chokshi, A., 1990. Astrophys. J. Lett., 351, L9.
Fabian, A.C., 1989. Mon. Not. R. astr. Soc, 238, 41p.
Guthrie, B. N. G., 1979. Mon. Not. R. astr. Soc, 187, 581.
Heeschen, D., 1960. Publ. Astron. Soc. Pacific, 72, 968.
Lilly, S. J., 1988. Astrophys. J., 333, 161.
Lilly, S. J., 1989. Astrophys. J., 340, 77.
McCarthy, P. J., van Breguel, W., Spinrad, H., & Djorgovski, S., 1987. Astrophys. J. Lett., 321,
L29.
McCarthy, P.J., Kapahi, V. K., van Breugel, W., & Subrahmanya, C. R., 1990, preprint.
Miley, G.K, et al. , 1989. ESO Messenger, 56, 16.
Palimaka, J.J. Bridle, A.H., Fomalont, E. B., &Brandie, G. W., 1979. AlJtrophYIJ. J., 231, L7.
Rees, M. J., 1989. Mon. Not. R. astr. Soc, 239, 1p.
Scarrott, S. M., Rolph, C. D., & Tadhunter, C. N., 1990. Mon. Not. R. astr. Soc, 243, Sp.
Spinrad, H. Djorgovski, S., Marr, J., & Aguilar, L., 1985a. Publ. Astron. Soc. Pacific, 97, 992.
256

Spinrad, R., private communication


Windhorst, R. A. et aI. , 1991. A.trophll" J., in press
K Band Galaxy Counts and the Cosmological Geometry

Lennox L. Cowie
Institute for Astronomy, University of Hawaii

ABSTRACT. The Hawaii deep survey has now established galaxy counts to a K
magnitude of 23. The low number density of these counts at the faint end favors
a qo = 0.5, A = 0 geometry which minimizes the available cosmological volume.
However these models provide no explanation for the large number of faint blue
galaxies. Combining spectroscopic data with K magnitudes for a blue selected
sample shows that the large number of galaxies at B = 23 - 24 appears to be a
population of blue dwarf galaxies lying at typical redshifts of 0.3 to 0.4. These
dwarf galaxies contain as much K light as the normal galaxy population and may
dominate the baryonic and dark matter content of the universe.

1. Introduction

In a K band galaxy sample we are seeing light from near-solar mass stars
whose lifetimes are comparable to the age of the galaxy. This means that evolu-
tionary corrections are much smoother and can be more securely modelled. There
is also the advantage that out to substantial redshifts the K band samples the rela-
tively well understood optical portion of the galaxy spectrum so other uncertainties
in the modelling are removed or minimized. Finally because the optical and near IR
spectrum is generally much flatter than the UV portion the dimming with redshift
in the near IR is generally much smaller than in the optical and we sample to large
redshift at much brighter magnitudes.
Motivated by these considerations the Hawaii deep survey was designed to
obtain deep infrared galaxy samples using the new infrared arrays. The survey
provides galaxy counts to K = 23 and a well-defined sample of K-selected galaxies,
with corresponding optical colors, which can be used as a basis for spectroscopic
studies. The follow up spectroscopic studies are nearly complete for a B < 24
selected sample but are only just begining for the K-selected sample. The optical
data and the spectrosopy of the blue selected sample are described in Lilly,Cowie
and Gardner (1991jLCG) while the K band counts and the properties of the K
band sample are given in Cowie et al. (1991).

257
T. Shanks et al. (eds.), Observational Tests otCosmologicallnjlation, 257-265.
© 1991 Kluwer Academic Publishers.
258

2. Optical and Infrared Galaxy Counts

The current status of the differential galaxy counts in Band K is given


in Figure 1. Consider first the B band counts, which I am going to compare with
number-count models, both with and without galaxy evolution, that I have adapted
from those of Yoshii and Takahara (1988). Luminosity evolution has a much larger
effect on the blue counts than cosmological geometry, as can be seen from Figure 1
where the evolving models are shown as solid lines and nonevolving models as
dashed lines. With the luminosity evolution included, one sees galaxies to much
larger distances and consequently the models predict many more counts, but of
course they predict that the excess counts should be produced by extended tails
or secondary peaks at high redshift. Prior to the advent of spectroscopic data, this
seemed a good explanation for the excess counts but as the faint-object spectra
have become available in the last couple of years it has become clear that it must
be ruled out (cf. LCG). To demonstrate this I show data from a number of complete
spectroscopic samples in Figure 2. To B = 24 there is only one possible object in
these samples which could even conceivably be at z > 1; the majority have z < 0.6.
Until the faint K band counts became available there were two possible ex-
planations for the combined faint blue counts and blue selected spectroscopy, one
cosmological and one astrophysical, with no clear cut way to decide between them.
Either there is a major population of blue dwarf galaxies that was present at
z - 0.3 - 0.4 but that was not included in the assumed present-day luminosity
functions (this is discussed in much more detail below) or alternatively there must
be much more volume at low redshift than even a qo = 0.02 Friedman model would
predict. The latter could be the case, for example, in a zero curvature n = 0.1 uni-
verse with a cosmological constant, a possibility that has many other attractions
such as solving the cosmological timescale problem and easing the difficulties of
large scale structure. As is shown in Figure 1, such a model (with some galaxy evo-
lution included) can roughly fit the number counts (e.g. Lilly, Cowie and Gardner
1991, Fukugita et al. 1990) and yet not violate the spectroscopic constraints.
However, we now have K band number counts which are more sensitive to
cosmological geometry than the blue counts are. This effort has produced the re-
sult that can be seen in Figure 1. Basically, at around K = 19 the counts stop
rising rapidly, they move to a much shallower slope. What is suggested here is
that the fainter magnitudes no longer sample to larger distances; we must have
reached at last the magnitude at which we have broken through to the end of the
galaxy-occupied cosmological volume. At this point the flat or slowly-rising counts
represent the faint-end shape of the galaxy luminosity function averaged through
this volume. As can be seen from Figure 1 the K band counts can be fitted either
by an open unevolving model (which is not very physically plausible) or by a flat
evolving model. Flat models with a cosmological constant overpredict the faint end
counts irrespective of the evolution.
259

B BAND COUNTS K SAND COUNTS


E+~~--r---r---r---r---r---r-~

10000 10000

~
"-
~
1000 "-
....'-'Cl '-'
t..J
1000
Cl
a
(J1
a(/)
"-
a::: 100 "-
.... a:::
t..J
100
a:J III
~ ~
:)
z ~
Z
10 10

0.1
12 26
B MAGNITUDE K MAGNITUDE

ligure 1 A comparison of blue and infrared counts with model predictions. The
blue number counts are taken from the compilation of Metcalfe et al. (1990) in
addition to those of Maddox et al. (1990; diamonds), Tyson (1988; shaded region)
and Lilly et al. (1990; solid squares). The infrared counts are from Glazebrook
et al. (1990; triangles), Jenkins and Reid (1990; shaded region) and Cowie et a1.
(1990; squares and diamonds). For the IR data I have shown error limits. Two
classes of model are shown. The dashed lines show models in which galaxies do
not change at all in luminosity or type. The bottom curve has qo = 0.5, the middle
qo = 0.02, and the top a zero-curvature model with n. = 0.1 and a cosmological
constant. The three solid lines show models with the same three geometries as
above but in which the galaxies evolve with z roughly following the prescription
of Yoshii and Takahara (1988). The galaxies are assumed to form at zJ = 5
in all cases. The K band counts fit best to a qo =0.5 cosmological geometry·
with evolution or a qo = 0.02 model with no evolution but the latter model is
quite physically implausible. All the models which fit the K band counts grossly
underpredict the faint-end blue counts.

3. Dwarf Galaxies and the Faint Blue Number Counts

For any model which roughly fits the K band counts we can see from Fig-
ure 1 that there is a huge excess of faint blue galaxies. Another way of expressing
the relative overabundance of faint blue galaxies and underabundance of faint IR
galaxies is to note t.hat the K-band galaxy sample shows a rapid trend to bluer
colors at fainter magnitudes and that there are relatively few faint objects that are
red in (B - K). This is shown in Figure 3 which is a histogram of the (B - K)
260

SPECTROSCOPIC SAMPLE
1.8r-----------------------------------------------------r---____-,

1.4 r compIete~
• BroadIIIrst et aI. 1988 and COIIesI et aI. 1990
• Hawaii spec;t/OacXIpiC
1.2 -
o SED redshIIt (spec:IR)&C:Opic lBO)
1.0 -

10.8 r
• •

0.8 I""
: .. •
o

0.4 -
. . .........:. -... i·· •..: .
• •
.., .. ' ·c· r___...··
..,.
:
. .--:--.,;,-.".-•
. . --- 0··- - ...
- - - 0

e. • :-:.~!r:.
; . ~-~y_...·r:'·:·.· . :. • •
0.2 -

O.O~~----~-------L-- I I
.. . .-. :.::... .- ,. . ...
.... 4" -.:-~;.-'" ••

______
~ LI _• ______
.1 ••
~
e,,\ ....

I
~

• •

I ________L_______
_______L ~

18 19 20 21 22 23 24 25
isopholal B magnitude

figure 2 Spectroscopic samples for B magnitude limited galaxies. The dots show
the B < 21 and B < 22.5 samples of Broadhurst et al. (1988) and Colless et al
(1990). The latter is about 80% complete. The sold squares show the Hawaii
spectroscopic data (e.g. 'Lilly et al. 1990) which is nearly complete to B 24. =
(Open circles are three objects that remain to be observed spectroscopically shown
at redshifts estimated from their colors.) Only one object at B < 24 defied
spectroscopic identification and could be at high z (upward-pointing arrow); nearly
all objects are at z < 0.6. The dashed line shows a predicted mean redshifi for a
model with no galaxy evolution; it provides a remarkably good fit.

color in various magnitude ranges. Objects which are not detected at the 1 (j level
in B are shown as open regions on the histogram and could lie anywhere to the
red. Compared to the observations the models we discussed above predict redder
distributions. The most extreme are the non evolving models but even a qo = 0.5
evolving model predicts too many red galaxies.
The blueness of the faint galaxy population is of course the cause of the
discrepancy between the blue and IR number counts. What must be explained
in any model is why there are so many very blue faint galaxies. However there
is a more profound conclusion that can be drawn from Figure 3, namely that
only those galaxies with the redder colors are being correctly understood in the
conventional smoothly evoloving models; the blue galaxies must correspond to some
other population or to episodic star formation in the fainter end of the normal
galaxy population (e.g. Broadhurst et al. 1988).
261
15

a:: 10
w
CD
:::!: K=21-22
~
z 5 25 OBJECTS
5 UNDETECTED
0
0 2 12 14
15

a:: 10
w
CD
:::!: K=20-21
~
z 5 46 OBJECTS
11 UNDETECTED

0
0 12 14
15

a:: 10
w
CD
:::!: K=18-20
~
z 5 57 OBJECTS
2 UNDETECTED
0
0 10 12 14
8-K COLOR

I'igure I Histograms of the (B-K) distribution of the K sample as a function


of magnitude. Galaxies which are not detected at the one sigma level in Bare
shown as the open area. They could lie anywhere to the red of their position in
the graph.

We can combine the redshift data with the K-band magnitudes to determine
what the population producing the faint B counts actually is. That is, the abso-
lute K magnitudes should give a good estimate of the galaxy mass except for the
most extreme star-bursting cases where it constitutes an upper limit. (They can be
roughly used without a K correction because of the flatness of galaxy SEDs in the
near infrared.) The absolute K magnitudes for the galaxies giving the B < 24
counts are shown in Figure 4. At brighter B magnitudes most of the galaxies
are at or near the K. of -25.3 typical of the most luminous elliptical galaxies
(Ho = 50 km s-1 Mpc- l ). At the faint end, which is also the position at which the
counts begin to rise rapidly above predictions, we begin to see many much smaller
galaxies that are typically about 4 magnitudes fainter in K. This is an unexpected
effect in a magnitude-limited sample where we expect the counts to be dominated
by the near-L. galaxies that can be seen to the limits of largest volumes. Roughly
two thirds of the faintest galaxies appear to be dwarf galaxies of this type. If we try
262

to turn this into a luminositl function we find that the total K luminosity density
is 8.1 x 108(Ho/50}L0/Mpc with roughly half coming from the dwarfs and half
from normal galaxies.

-28 __ ABSOLUTE K MAGNITUDES OF SPECTROSCOPIC B SAMPLE

-27

w
-26

-25
.
a

0
::l
f-
Z -24-
<.:l
~ a
-23
'"
w
f-
::l -22
...J
0
Vl
co
«: -21

-20

-19

-18
20 21 22 23 24-
TOTAL B MAGNITUDE

figure 4 Absolute K magnitudes for the B < 24 sample. The solid symbols
show objects with spectroscopic redshifts and the open symbols those where the
redshifts have been estimated from colors. The dashed line shows the approximate
magnitude of a typical giant galaxy.

Since the amount of K light in the dwarfs is essentially equal to that in


the normal galaxies the dwarfs must contain a comparable amount of baryonic
matter in low mass stars. Given that they are likely to have considerably more gas
than stars, they are almost certain to dominate the baryonic mass. In particular
if they contain approximately four times as much mass in gas as in stars, typical
of local blue dwarf irregular galaxies, then the baryonic density would be roughly
1.4 x 10- 31 g cm- 3
These estimates are quite approximate but can be compared directly with
homogeneous Big Bang nucleosynthesis which predicts a baryon density of 2 - 9 x
10- 31 g cm- 3 (e.g. Yang et al. 1984) and it is apparent that the dwarfs can provide
the required baryon density without any problem. The interesting thing is that
it is entirely plausible, physically, for the blue galaxy population to contain the
bulk of the baryons. The dwarfs could also be more uniformly distributed than,
or even distributed preferentially away from, normal galaxies at the same epoch,
which would result in quite significant biasing of baryonic matter from present-day
luminous material.
263

4. Summary

The data is still fluid and has yet to be confirmed by other groups, and the
model results are still subject to interpretation. However, to my mind the most
natural conclusion at this time is that we live in a qo = 0.5, A = 0 universe, which
allows us to understand the K band counts naturally', and that there is a population
of dwarf galaxies, responsible for the large number of faint blue galaxies observed,
that flourished at z "" 0.4 but has now burned out and either destroyed itself or
in some other way (perhaps by being too faint) and is no longer counted in the
present galaxy population.
If we accept that the blue number counts and spectroscopy imply the ex-
istence of the dwarf population then it contains at least as much baryonic mass
as (and probably considerably more than) the present-day 'normal' galaxies. This
could account for the missing baryon problem which is suggested by primordial
nucleosynthesis models.
Finally since the dwarfs may not be distributed as the obseved normal galax-
ies are it is possible that they may imply substantial biasing of the present lumi-
nous material from the baryonic matter. Actually determining whether this is in
fact the case will require large faint galaxy samples that can be cross correlated
with brighter normal galaxies at the same redshift.
I am very grateful to Toni Songaila for many interesting discussions which
greatly clarified my thinking on these problems. I would also like to thank John
Kormendy and Ken Freeman for extremely illuminating dicussions about dwarf
galaxies and my student Jon Gardner and colleagues Klaus Hodapp, Esther Hu,
Simon Lilly and Richard Wainscoat on whose work this talk is partly based. This
work was supported in part by NASA grant NAGW 959.

References

Broadhurst, T. J., Ellis, R. S. and Shanks, T. 1988, M.N.R.A.S., 235,827.


Colless, M. M., Ellis, R. S., Taylor, K. and Hook, R. N. 1990, M.N.R.A.S., 244,
408.
Cowie, L. L., Gardner, J. P., Wainscoat, R. J., and Hodapp, K. 1991, preprint.
Fukugita, M., Ta.ka.hara, F., Yamashita, K. and Yoshii, Y. 1990, preprint.
Glazebrook, K., Peacock, J. A., Collins, C. A., and Miller, L. 1990, preprint.
Jenkins, C. and Reid, J. 1990, A. J., in press.
Lilly, S. J., Cowie, L. L., and Gardner, J. P. 1990, Ap. J. Suppl., in press.
Maddox, S. J., Sutherland, W. J., Efstathiou, G., Loveday, J., and Peterson., B. A.
1990, M.N.R.A.S., 247, Ip. .
Metcalfe N., Shanks, T., and Fong, R. 1990, preprint.
Yang, J., Turner, M. S., Steigman, G., Schramm, D. N. and Olive, K. A. 1984,
Ap. J., 281,493.
Yoshii, Y. and Ta.kallara., F. 1988, Ap. J., 326, 1.
264

DISCUSSION:

Guiderdoni: Your model apparently does not fit the deep counts in the visible. Do you
think it is due to the model of spectrophotometric evolution that you use or to Tyson's or
Lilly's et al. data? My feeling is that it is due to the crude modelling of UV spectra in
Yoshii and Takahara's code.

Cowie: These models probably shouldn't be taken too seriously at this level of detail.
There are just too many free parameters and any model of luminosity evolution is not
going to be that accurate. I think the data is reasonably consistent and basically alright
and I would emphasize that we have a very gross problem already in the (B - K) colours
that simple standard models aren't addressing. The more complex models which cover the
dwarf problem have even more freedom of course.
Shanks: At the recent Oxford conference I suggested to you that the total number density
of galaxies in Tyson's B counts corresponded to the number density expected in an no =
1 model, if the local galaxy luminosity function was integrated to include dwarf galaxies.
Although this model has similarities to the model you are now proposing here, at that time
you objected on theoretical grounds to the large amount of evolution required for dwarfs if
a fit to the B counts was to be obtained. What has happened to this theoretical objection?

Cowie: My theoretical object was that for an a = -1.1 Schechter function there is not
enough mass at the faint end to allow enough nucleosynthesis to account for the integrated
blue light. This objection holds only because there are not very many local dwarves. What
now seems to be the case is that there were many more dwarves visible at z tv 0.3 - 0.4 and
these have comparable amounts of mass to the normal galaxies. This instantly solves my
problem but at the expense of having a major change in the galaxy population between z
= 0 and z = 0.4.
Peebles: In your interpretation of the deep K-band galaxy counts are you not concerned
about the effect of redshift on surface brightness which is a factor (1 + Z)4 tv 100 at z tv
3?

Cowie: This effect is of course already included in the count models (i.e. it is contained in
the magnitude). A more critical issue is how complete the counts are at the faint end when
the surface brightness drops near the limiting detectable value. We have simulated and
tested this in various ways. As long as the galaxies are similar to low z normal galaxies in
size they usually drop out of the counts over a relatively limited range of about a magnitude
irrespective of size. The counts shown have not been continued beyond the point where the
correction factor for this effect is larger than 2. We have also simulated cases where the
galaxies are much larger at high z (i.e. expanding by factors of 5 at z = 2). These models
show serious inconsistencies with observed size distributions. Finally, the B gnd K counts
contain roughly the same set of objects. Any missing galaxies are missing in B as well.

Renzini: If the progeny of the "excess" dwarf galaxies at z '" 0.4 is represented by local
dwarf spheroidals, would you expect the excess dwarfs to show the tendency of clustering
265

around giant galaxies, as for example do the dwarf spheroidals in the local group?
Cowie: I don't really know. However, if I was speculating it seems possible that bursting
Ims might find it easier to retain some of their material in higher pressure or density
environments such as the neighbourhoods of giant galaxies or in regions such as the Virgo
cluster, which would allow the dwarf spheroidals to be the descendants of some part of this
population. The long term answer is to determine the correlation function for the blue
dwarf population.
Ellis: (1) Your result that the absolute luminosities of K - selected galaxies apparently
become dwarf dominated at faint limits is evident in our respective blue spectroscopic
surveys. We find (Colless et al, in preparation) that the mean redshift of the flat-spectrum
population at R '" 23 is actually ~ than the no evolution prediction. Why then are the
faintest Mk objects not blue in B - K?
(2) Your new conclusion that the faintest K counts, by themselves, apparently rule out open
models depends critically on the faintest count in a single field. It must also depend on
the precision to which you estimate the correction for incompleteness. In view of the stark
contradiction with the more established result based on optical counts surely you must take
seriously the worry that the faint K counts are underestimated?
Cowie: (1) They are very blue in (B - K) in general though not necessarily always. In
the B < 24 sample for the eleven galaxies with Mk brighter than -24 (Ho = 50 km/s/Mpc)
the median (B - K) is 5.3 while for the eight galaxies with Mk less than -24 the median (B
- K) is 3.7. However, two of these faint galaxies are an Elliptical and an Sa respectively so
the dwarves aren't universally blue.
I should also say that the dwarf effect was clear from the Durham spectroscopic samples.
What is now a little clearer are that these objects genuinely are low mass galaxies and that
there is a lot more of them than we are identifying now.
(2) I think you're being somewhat inconsistent here. We know from our spectroscopic
samples that we have a very serious problem with our understanding of the optical counts
which is caused by the blue dwarves, so that any attempt to claim that these established
counts determine an open geometry is highly suspect. The K band counts to K = 22 are
consistent over three fields though it is true that K = 22 - 23 is based on a single field.
Already at K = 21 -22 however, the realistic evolving open models are a factor of 2 over
the observations. However, the real point to keep in mind is that when we throw in the
dwarf population having high volume geometries gets really hard. What we have then is
an additional population plus the normal galaxies and low observed counts. At this point
flat geometries look extremely attractive.
SELECTION EFFECTS IN REDSHIFT SURVEYS

Y. Yoshii a and M. Fukugitab


a National Astronomical Observatory, Mitaka, Tokyo 181, Japan
b Yukawa Institute for Theoretical Physics, Kyoto University, Kyoto 606, Japan

=
ABSTRACT. The recent Durham redshift survey to BJ 22.5 mag shows a lack of galaxies with a
redshift higher than z ~ 0.7 for their survey magnitude, while cosmological models with the Tinsley-
Bruzual type luminosity evolution of galaxies predict a substantial high-redshift tail in the redshift
distribution. We point out that the suppression of high-redshift galaxies is due to the selection
effects; when the effects of using the isophotal magnitude and finite seeing are taken into account,
we find that a significant fraction of high-redshift galaxies should not be observed in agreement with
the observation. This means that we do not need to modify the current models of galaxy evolution.
We also find that the observed redshift distribution favours a low-density universe in agreement with
the conclusion from the number count of faint galaxies.

1 Introduction

The Durham redshift survey of faint galaxies has given an interesting cosmological puzzle
(Broadhurst et al. 1988; Colless et al. 1990). The authors of the Durham group suggested
that luminosity evolution of galaxies is much smaller than is needed to explain the steep
rise of the number count of faint galaxies as a function of magnitude (e.g., Bruzual and
Kron 1980; Yoshii and Takahara 1988; Guiderdoni and Rocca-Volmerange 1990). The
observed redshift distribution shows a sharp cutoff at a high redshift, while cosmological
models with galaxy evolution taken into account in a canonical way predict a conspicuous
tail of high-redshift galaxies. On the other hand, such a tail is absent in models without
galaxy evolution, and they give a better fit to the data. The analysis leading to this
conclusion, however, does not take account of various selection effects which are necessarily
associated with the observation offaint galaxies (see Ellis et al. 1981; Pritchet and Kline
1981; Phillipps et al. 1990). In this talk we point out that the selection effects, especially
those arising from the algorithm of the magnitude measurement, seriously affect the redshift
distribution. These effects counteract the effect of galaxy evolution in the sense that they
truncate the redshift distribution on its high-redshift side.

267
T. Shanks et al. (etis.), Observational Tests o/Cosmological Inflation, 267-271.
© 1991 Kluwer Academic Publishers.
268

2 Selection Effects with the Isophotal Magnitude Scheme

Galaxy survey programmes generally set the threshold of surface brightness above which
galaxy images are detected. Thus, a certain fraction of high-redshift galaxies would be un-
detected when the combined effects of (1+Z)-4 cosmological dimming and the K correction
lower their peak surface brightness below the threshold. The magnitude of a detected galaxy
is measured by integrating the surface brightness around the image out to an isophote set
by a given brightness threshold, and the magnitude thus defined is called the isophotal
magnitude. On the other hand, when one predicts number of galaxies with some cosmolog-
ical models, one often implicitly uses the total magnitude and assumes the 100% detection
efficiency for galaxies. Our emphasis in this talk is laid on the fact that the mismatch of
these magnitude schemes becomes serious at faint magnitudes.
For an illustrative purpose of demonstrating the effect of isophote, we assume that
galaxies have a radially symmetric luminosity profile of the form

g(.8) = exp( -An.81/ n ) with .8 = r/re ,


where re is an effective radius and the coefficient An is fixed by G(1)/G(oo) = 1/2 by
definition. We take n = 4 for ellipticals and n = 1 for spirals. The integrated profile out to
the normalized radius .8 is given by

G(.8) = 211" L{3 g(.8).8d.8 .

The luminosity profile g(.8) of a galaxy is also smeared by seeing effects. Taking a
Gaussian form with the dispersion 0"0, the smeared profile has the form

with 0" = O"odA/re where dA is the angular diameter distance and Io(x) is the modified
Bessel function of the first kind. We denote the integrated profile of 9(.8) by G«(3).
With these preparations, the apparent isophotal magnitude of a galaxy at the redshift z
having absolute magnitude M enclosed within an isophotal radius (3ph on the object plane
is given by

m = M + K + E + 5Iog(dL/10pc) - 2.5log[G«(3ph)/G(OO)] ,

where J( and E are the corrections for redshift and luminosity evolution, respectively, and
dL is the luminosity distance. Correspondingly the surface brightness is given by

S[mag/arcsec2] = M + K + E + 5log[re(1 + z)2/10pc] + 26.5721- 2.5Iog(g(.8ph)/G(oo)]


The radius .8p h is determined by equating S to its threshold value SL which defines the
isophotal magnitude. To simplify our analysis we assume that the central surface brightness
is constant, i.e, 14.8 mag/arcsec 2 for ellipticals [Fish's (1964) law] and 21.65 mag/arcsec 2
for spirals [Freeman's (1970) law]. This is equivalent to assuming a correlation between the
absolute magnitude M and the effective radius Te of galaxies.
269

The local properties of galaxies of various morphological types are the same as those
taken in Yoshii and Ta.ka.hara (1988). To incorporate the effect of galaxy evolution we adopt
the type-dependent evolution model by Arimoto and Yoshii (1986, 1987). We assume that
the shape of the luminosity function and the comoving number density of galaxies remain
unchanged with redshift.

:I Results and Discussion

Figure 1 shows the histogram for the Durham redshift distribution of galaxies for BJ =
21 - 22.5 mag observed on the AAT. The observed galaxies were selected using the surface
brightness threshold of SL = 26.5 mag/arcsec2 on the AAT plate and the seeing is roughly
1 arcsec FWHM (we found that the result is not sensitive to the seeing for the present case).
The solid curve represents the model redshift distribution calculated for the cosmological
mass density parameter no=1 on the left panel and for no=O.1 on the right panel. The
cosmological constant A is assumed to be zero. The model is normalized to give the observed
number of galaxies. Also shown in the figure are the models both with and without galaxy
evolution for SL = 00 as usually used to interpret the observation. The same normalization
is taken for the model with galaxy evolution (dashed curve) so that the part missed by
the adoption of the isophotal magnitude is explicitly seen. On the other hand, the model
without galaxy evolution (dash-dotted curve) is renormalized to give the observed number.

20 20
(01 ao' I a
(bl o' 0.1

15 15
N ;:.:;-
z z
10 10

5 5

z z
Figure 1. The redshift distribution of galaxies for BJ = 21 - 22.5 mag. The Durham data [ColJess
et aI. (1990)] are compared with the model predictions for the universe with (a) 0 0 =1 and (b) 0 0 =0.1.
The solid curve shows a model with canonical galaxy evolution in which the isophotal selection effect
(SL = 26.5 mag/arcsec2 ) is taken into account. The model is normalized to give the observed number of
galaxies. Dashed curve and dash-dotted curve represent the models with and without galaxy evolution,
respectively, for which the total magnitude scheme (SL = 00) is employed. The same normalization is taken
for the model with galaxy evolution (dashed curve). The model without galaxy evolution (dash-dotted
curve) is renormalized to give the observed number of galaxies.
270

The figure shows that a significant fraction of high-redshift galaxies would be undetected
when the isophotal magnitude scheme is employed. The extended high-redshift tail pre-
dicted by models with galaxy evolution for SL = 00 is strongly suppressed by the selection
effects. As a result, the predictions with the selection effects properly taken into account
resemble those of the models without galaxy evolution but assuming SL = 00, which have
been considered to fit the observation. If the effect of galaxy evolution is taken out of our
models, the predicted number of visible high-redshift galaxies falls considerably short of the
observed. This demonstrates that we need the effect of galaxy evolution in order to explain
the observed redshift distribution; the canonical luminosity evolution model is sufficient.
By looking at two models, (a) for no = 1 and (b) for no = 0.1, we see that the low-
density universe model gives a better fit to the observed redshift distribution, especially for
the high-redshift tail. While we may not be able to rule out the high-density universe model
solely from our analysis because of possible uncertainties in the galaxy evolution model, our
conclusion that the low-density universe is favoured is consistent with the conclusion drawn
from the number count of faint galaxies (Fukugita et al. 1990; Guiderdoni and Rocca-
Volmerange 1990; Yoshii and Peterson 1991). We also find that the low-density universe
of no=O.l with a non-zero A, if it is sufficiently large to make the universe flat, predicts a
little too few high-redshift galaxies, although the possibility is not ruled out because of the
uncertainties again in the galaxy evolution model.
Our present consideration shows that it is indispensable to fully take into account the
selection effects in interpreting the observations. We emphasize that the selection effects
discussed here strongly influence the visibility of high-redshift galaxies, and hence they
show up most dramatically in the redshift distribution. These effects, however, are rather
modest on the number count which is integrated over the redshift space.

Acknowledgements. We are grateful to R.S.Ellis, B.A. Peterson and T.Shanks for discussions. This
work was supported in part by the Grants-in-Aid of the Ministry of Education, Science and Culture
of Japan (Nos.02211207 and 02640207).

References

Arimoto, N., & Yoshii, Y., 1986. Astr. Astrophys., 164, 260.
Arimoto, N., & Yoshii, Y., 1987. Astr. Astrophys., 173, 23.
Broadhurst, T.J., Ellis, R.S., & Shanks, T., 1988. Mon. Not. R. astr. Soc, 235, 827.
Bruzual, A.G., & Kron, R.G., 1980. Astrophys. J., 241, 25.
Colless, M., Ellis, R.S., Taylor, K., & Hook, R.N., 1990. Mon. Not. R. astr. Soc, 244, 408.
Ellis, G.F.R., Perry, J.J., & Sievers, A.W., 1981. Astron. J., 89, 1124.
Fish, R.A., 1964. Astrophys. J., 139, 284.
Freeman, K.C., 1970. Astrophys. J., 160, 811.
Fukugita, M., Takahara, F., Yamashita, K., & Yoshii, Y., 1990. Astrophys. J., 361, L1.
Guiderdoni, B., & Rocca-Volmerange, B., 1990. Astr. Astrophys., 227, 362.
Phillipps, S., Davies, J.I., & Disney, M.J., 1990. Mon. Not. R. astr. Soc, 242, 235.
271

Pritchet, C., & Kline, M.I., 1981. Astron. J., 86, 1859.
Yoshii, Y., & Peterson, B.A., 1991. AstrophYIl. J., 372, in press.
Yoshii, Y., & Takahara, F., 1988. Astrophllll. J., 326, 1.

DISCUSSION:

Guiderdoni: You would not be surprised if I strongly disagree with you. It seems to me
that you do not reproduce the counts with your model because you do not estimate the
uv flux of galaxies consistently from the uv flux of stars. So you are left with the extreme
solution A =I 0 which fits the counts with its large volume elements.

Yoshii: As we go fainter the predicted number count reaches a maximum and then de-
clines. The count at this maximum depends on the geometry of the universe, and the
luminosity evolution of galaxies do not change significantly the value of the count at this
maximum. Tyson's counts reach deep enough to determine the geometry of the universe,
almost free from the uncertainty in galaxy evolution models.

Guiderdoni: Another comment is, we modelled surface-brightness and seeing effects in a


manner similar to yours and we found that providing the building blocks are comparable
to nearby galaxies (giants and/or dwarfs), faint galaxies are visible in the current surveys.

Yoshii: I simply disagree with your opinion that they are all visible.

Blanchard: We heard from Cowie that the count in the ]( band leads to almost exactly
opposite conclusions. What are your comments?

Yoshii: In Hawaii's ](-band survey the surface brightness threshold is set to be 24 mag/arcsec 2
and the seeing is 1.7 arcsec FWHM. Number count models which are not corrected for the
selection effects overestimate the observable number of faint galaxies and hence the value
of no, even if magnitudes of detected galaxies are measured over sufficiently large aperture
to give approximately total magnitudes. This is also the case with Cowie's model. Our
analysis on the 3.5-arcsec aperture magnitude scheme indicates that the predicted count
for no = 1 is about dex 0.5 below the observed count at 22]( mag. On the other hand, the
prediction for no = 0.1 agrees very well with the observation.
An Inflationary Alternative to the Big-Bang

F. Hoyle
10~ Admiral's Walk,
West Cliff,
Bournemouth, BH~ 5HF,
U.K.

ABSTRACT. The model proposed here is inflationary but with inflationary units of the scale of
superclusters of galaxies rather than the much larger scale of more standard inflationary theories.
The physical conditions in an inflationary unit are the same as in a standard big-bang model with
°.
a baryon-to-photon ratio of ..... 3.10- 1 Because the microwave background was smoothed after
galaxies were formed, fluctuations in the background on angular scales of arc minutes to degrees
are very small. The creation phase of the inflationary units is at a look back redshift of ..... 30 and
is potentially observable. A universe with repeated inflationary phases can have relevance to basic
physics, affecting cut-off processes as in quantum electrodynamics.

1 Historical Considerations

The standard big-bang model may provide fruitful ground for the observer and for particle
physicists but it offers little or nothing to the cosmologist. For the reason that essentially all
possibilities for the model had already been discovered by the time of Robertson's famous
article (1933). The metric used took the now well-known form

dr 2
ds 2 = dt 2 - S2(t)[_-
1- kr2
+ r2(d8 2 + sin28dcp2)] (1)

c = 1; k = 0, ±1
The behaviour of 8(t) was determined from Einstein's equation, viz

a
R -"21 g a R = -81fGTEin6tein
a
, (2)

where T~n8tein was the usual physical energy-momentum tensor. It may be noted that a
local body with spherical symmetry, homogeneity and isotropy has an interior solution of
the same form. The exterior solution for a local body is different, however, and is in general
much harder to obtain. It has been found in the simple case where the exterior is empty,
but not for an inhomogeneous exterior. This short-coming will be the reason why, at a later
stage of my contribution, I shall descend from theory to scenario.

273
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 273-278.
e 1991 Kluwer Academic Publishers.
274

All this is for universes that are homogeneous and isotropic of course. By departing
from these restrictions it is possible to obtain other models, of which perhaps the best
known is the rotation model of Godel. Otherwise it is necessary to introduce new physics,
if the situation is to be broadened. In considering this situation (many years ago) it seemed
that the most promising manoeuvre would be to exploit a logical weakness in the big-bang
models, namely that they postulate a sudden origin for matter without describing physically
or mathematically how matter-creation is supposed to take place.
It seemed interesting to define a scalar field with matter-creation as its source. I called
the field C, not having the prescience to call it </>, with the wave-equation

OC=nm (3)
determining it, where the number of particles of mass m created per unit proper 4 - volume
at the space-time point X is n(X). The next step was to define the contribution of C(X) to
the universal action. Not having any guidance from particle physics in those days, I used
a feature of the universe that we then thought to have great importance. Although from
relativity theory there are supposed to be no preferred directions, the universe does in fact
have a preferred direction at every point, the direction of the Hubble flow. Perhaps it is
because of having a good semantic description of this remarkable property that the modern
generation does not see it as so remarkable - semantics tends to destroy curiosity. The
situation is actually so important that even very local properties like the winds blowing
dominantly on the Earth from west to east can be traced to it. Anyway, having obtained a
scalar, the gradient of the scalar gave just such a directivity, so that it seemed a good idea
to construct the universal action from Ci = BC/Bx i , viz.

(4)

in familiar relativity notation. Now one varies (4) with respect to the space-time geometry,
writing the results in the form
J T~
Ok r-;: 4
OgikV -gd X, (5)

T~k being the contribution of the C-field to the energy-momentum tensor. The dynamics
is then the same as in the big-bang models, but with a total energy-momentum tensor Tik
given by
ik ik
T = T Einstein + Tc
ik
(6)
Solving for the simple case n = constant it was possible to show that the universe must
tend to the flat case k = 0, with a metric of the de Sitter form

ds 2 = dt 2 _ e2Ht [dr2 + r2( d8 2 + sin 28d<p2») (7)


the Hubble constant H being related to n by

H = ( --mn
47rG )1 3 (8)
9
Despite matter creation, energy and momentum were conserved, necessarily so because
the gravitational equations gave
275

TJk = 0 (9)
Thinking of conservation of energy in a more intuitive way, the expansion of the uni-
verse contributes negatively and the creation of matter positively, the positive and negative
contributions balancing each other. This can be seen in (8) which shows that the larger the
creation n. the faster the expansion H. It seems worth mentioning this early work because
it has a remarkable family resemblance to the modern concept of inflation. The similarities
are:

(i) Both rely on a scalar field.

(ii) In both the pressure contribution is negative.


(iii) Both lead to the de Sitter metric.

(iv) Both lead to a flat universe.

(v) The explanations of isotropy, homogeneity, and the resolutions of the horizon problem
are the same.

To the extent that if one identifies modern inflation with the Cheshire Cat this was the grin
on the face of the cat. Both lead to a closure condition, but with the interesting difference
of detail that the above led to
(10)
with 41rG instead of 81rG. What seems to be implied is that as soon as one couples a scalar
field to gravity the result is essentially inescapable and, moreover, this is the way to escape
from the strait-jacket of the big-bang models.

2 Statement of the Model

Now let us jump ahead to modern inflation by asking how much freedom in our inflation-
ary concepts is permitted by the observational constraints. I like to divide observational
constraints into those that are tactical and those which are strategic. The relation of the
f
ages of globular clusters to H-1 is an example of a tactical constraint. Here I shall be
concerned more with strategic issues. In addition to the expansion of the universe, the two
strategic constraints are those related to the microwave background and to the primordial
synthesis of 4He. For the model to be described below, the microwave constraint appears to
be rather easily overcome. The key to this success is to have the last interchange of photons
and matter occurring after the galaxies are formed. I will refer back to this issue at a later
stage. The 4He constraint has to be met by taking the universal baryon-to-photon ratio
°.
to be in the region of 3.10- 1 Many people have struggled to break this constraint but
without apparent success, so that I think it should form the centrepoint of the suggested
model. Wagoner, Fowler and Hoyle (1966) expressed this requirement in the form

(11)
276

where T9 was the temperature in units of 109K. Putting T9 = 3.10-9 for the present
day situation (11) gave for the present day baryonic density, '" 3.1O-3I g cm- 3 , agreeing
almost precisely with Oort's estimate for the smoothed density of material in galaxies, an
agreement that strengthened the view that 4He (and D, 4He and perhaps 7Li) were produced
in primordial nucleosynthesis. Since (11) gave a density that was only about 3 percent of
the closure density,
3H2
Ptotal ~ 8'11"G ' (12)
the implication was that {lbaryon ~ 0.03, with the consequence that if {ltotal = 1 most of
the material in the universe had to be non-baryonic.
Wagoner et al. stressed the need to understand why the baryon to photon ratio ~ had
to be '" 3.10- 10 • But even after a quarter of a century of effort by cosmologists this point
(so decisive in nucleosynthesis theory) remains cosmologically obscure. Here I propose to
make ~ ~ 3.10- 10 the centrepoint of the model, which is to differ from that described by
Guth in having its inflationary units much smaller and consequently of being potentially
observable. A mass of material containing thermodynamic radiation of temperature T can
hold the radiation without being blown violently apart if the mass exceeds a Ininimum value
M given by
-M '" 1010T.96 / Ptotal
2 (13)
o -

The model comes from the hypothesis that inflationary units have masses not greatly in
excess of (13), Ptotal in (13) being the total density of gravitating material within the unit.
EliIninating T9 between (11) and (13) leads to

Mbaryon ~ M Pbaryon ~ 1020 (Pbaryon)3 (14)


o 0 Ptotal Ptotal

Putting Pbaryon/ Ptotal ~ 1/30 yields

(15)

a good value for the mass of a supercluster. The circumstance that the required value of
~ leads in this easy way to the masses of superclusters gives the reason for considering the
present model.

3 Discussion

Unfortunately, I must now descend from theory to scenario. The many inflationary units
are not disconnected as they can be if taken much larger. They are connected via a scalar
field and with exterior solutions involving partial differential equations, which one Inight
summon the resolve to tackle if one could be assured of the correctness of the model, but
which must be dealt with by scenario at this early stage. Partial differential equations have
the property of being critically dependent in some circumstances on external boundary
conditions. Proceeding from this known mathematical property I would hope for a solution
in which the inflationary property became triggered in already-existing seeds (e.g. collapsed
277

objects) by a critical tuning imposed by the external scalar field, with the consequence that
the inflationary property could occur more of less synchronously in local regions distributed
everywhere throughout the universe. Inflation changes the scalar field. When it has done so
sufficiently to destroy the mathematically imposed tuning condition, the inflationary phase
comes to an end. Thus one has conceptual provision for turning the inflationary phase both
on and off.
It would be an advantage to basic physics if this process could occur repeatedly. Renor-
malisation in physics depends on working with fields whose Fourier resolutions are cut-off
at some very high but finite frequency. The device of eventually letting cut-off frequencies
tend to infinity leads to a situation that is at best mathematically dubious and at worst
physically absurd, as for instance the theoretical mass of the electron -+ -00 . So long as
the universe does not empty in the future, renormalisation procedures can be formulated in
ways that keep cut-off frequencies finite ('" 1040 times particle masses) and a logical morass
is avoided. I find the possibility of cosmology reacting back on to basic physics in this
fundamental way attractive. A simple analogue for the repeated inflationary phases might
be a two-stroke engine which has short-lived explosion phases interspaced with much longer
evacuation phases. In such a model we live today in an evacuation phase. The overall
space-time metric describing the running of the engine is de Sitter.
There will be an early stage in an evacuation in which neighbouring inflationary units
merge. It seems attractive to associate the formation of galaxies with this stage, occurring
at a universal mean density'" 10- 24 g cm- 3 , or perhaps an order of magnitude less than
this. H the present day total density is '" 10- 29 g cm -3, the redshift z at which merging
from the last inflationary phase occurred is given by

1 + z ~ 105 / 3 , i.e. z ~ 30 to 50 (16)

I promised to refer back to the microwave background. Thinking of the thermalising


agent as being metallic needles produced at the time of galaxy formation and distributed
by radiation pressure on a scale that is fairly uniform in space, and taking the present day
mean-free path for quanta as 104 Mpc, the mean-free path at galaxy formation would be
less that this by (1 + Z)-3, i.e. as little as 100 kpc. During the expansion from 1 + z ~ 30
to the present day the microwaves have been thermalized and made uniform on scales from
100 kpc upwards, to 1 Mpc, to 10 Mpc, to 100 Mpc, .... Remembering that the black-
body distribution is independent of the precise distribution of the thermalising agent there
would be little prospect, in the situation I have just described, of detecting departures from
smoothness, at any rate on angular scales of arc minutes up to several degrees. Whether or
not the situation is smooth on larger angular scales would depend on whether or not the
inflationary process itself was smooth on such scales.
The present model has a desirable property that can hardly be denied. It is intimately
connected with astrophysics. The connection might indeed become very close when the
conditions under which inflationary units merge are examined closely. Random motions
will surely be produced, perhaps several times 10,000 km s-l. Such motions will, however,
have become damped by 1 + z in the following expansion, to '" 300 km S-1 at the present
day. One might think that this was the origin of the present day deviations from the
strict Hubble flow. Galaxy formation itself can be thought of as occurring at the interfaces
278

between adjacent inflationary units, leading to a distribution of galaxies in more or less


spherical shells.
IT we think that galaxies could have existed in cycles preceding the present cycle what
would happen to them during an ensuing inflationary phase? IT too close to an inflationary
unit they would be sheared apart. But if situated at a merger boundary where the density
of new material was rv 1O-24g cm3 , only pre-existing galactic material at lower density than
this would be sheared. Particular regions at higher density would survive. Could globular
clusters be such surviving fragments one can wonder? IT so, the great ages of some globular
clusters would receive an easy explanation.
It seems to me that the model survives all the early observational constraints that one
throws at it. It does so without taking refuge in much that is unobservable. Although
beyond present day techniques, even the last inflationary phase is potentially observable.
As my bottom line : When cosmologists have considered that they were looking back to
the origin of the universe what they were really looking back at was the last inflationary
phase in a repeated number of cycles, which on average can be conceived of as a continuing
inflationary universe of the de Sitter form.

References

Robertson, H.P., 1933. Rev. Mod. Phys., 5, 62.


Wagoner, R.V., Fowler, W.A. & Hoyle, F., 1967. Astrophys. J., 148, 3.
DYNAMICAL ESTIMATES OF no FROM GALAXY CLUSTERING

SIMON D.M. WHITE


Institute of Astronomy,
Madingley Road,
Cambridge CB3 OHA.

ABSTRACT. I review dynamical methods for obtaining masses on scales larger than galaxies, and
the assumptions and uncertainties involved in using these to estimate the density of the Universe.

1 Introduction

The only "good" way to measure the global density parameter, no, from dynamics is to
measure the deceleration of the Universe directly. This requires accurate recession veloci-
ties and relative distances for both nearby and distant objects. Einstein's equations then
give the total density of gravitating matter, including any contribution from a cosmological
constant. This program has yet to produce a believable result because of the difficulty in
obtaining reliable relative distances over cosmological scales; however, the situation may
not be hopeless (see R. Ellis's contribution to this volume). What usually passes for a
dynamical determination of no is in fact very different., Dynamical measurements of an
individual object (the "Great Attractor"), of a number of individual objects (galaxy clus-
ters), or of the statistical properties of an ensemble of objects (binary galaxies - the "Cosmic
Virial Theorem") are used to estimate a mass for the objects in question, and the result
is then extrapolated in some way to obtain the mean mass density of the Universe. The
extrapolation requires some form of the assumption that the mass per galaxy, or the mass
per unit light, MIL, is the same in the objects concerned as in the Universe as a whole.
The result is only as good as this assumption, and unfortunately there is little reason to be
confident of its validity. There is strong evidence that the distributions of mass and light
differ significantly on scales larger than individual galaxies.
The logical sequence in such arguments can be represented by the equation:

"no" = C x MIL -;- Pc. (1)


Here C is the luminosity density of the Universe, MIL is the mass-to-light ratio of the
objects concerned, and Pc is the critical density required to close the Universe. I have put
no in quotation marks to emphasise the assumption that the relation between light and mass
in the Universe as a whole is the same as in the objects measured. Additional uncertainty

279
T. Shanks et al. (eds.), Observational Tests oj Cosmological Inflation, 279-291.
© 1991 Kluwer Academic Publishers.
280

comes from errors in the observational measurements of each quantity in the equation. The
uncertainty in the critical density comes purely from that in the extragalactic distance scale,
Pc oc H6. Since dynamically measured mass-to-light ratios scale as MIL oc Ho, and the
luminosity density obtained from a galaxy redshift survey scales as C oc Ho, the derived
estimate of no is mercifully independent of Ho. However, both MIL and C have substantial
errors from other sources.
The luminosity density, C, is usually measured using photometry in the B band. The
luminosities are then quite sensitive both to obscuration and to the population of young
stars in a galaxy. As a result, the appropriate extrapolation of equation (1) is unclear since
it should plausibly reflect any population difference between the objects used to estimate
MI L and the Universe as a whole. This can be a problem, for example, when using
galaxy clusters, since these usually contain a much smaller fraction of star-forming galaxies
than the rest of the Universe. In principle, photometry in a redder band, perhaps the
near-infrared K band, would be preferable, but it is not yet available for any suitable
redshift survey. Available estimates of the luminosity density at B are quite uncertain
both because the results from various surveys of the nearby universe show large scatter (a
recent compilation by Efstathiou, Ellis, and Peterson (1988) concludes that CB = 1.9 ±
0.06 108hL0/Mpc3, where h = Ho/100 km/s/Mpc) and because galaxy counts seem to show
a serious discrepancy between the galaxy density nearby (rnB < 16) and that at somewhat
larger distances (z rv 0.1, rnB < 19; Maddox et al. 1990). These problems are most likely
due to photometric difficulties or to galaxy clustering, although the second problem could
also be caused by remarkably rapid evolution of the galaxy population. If we adopt the
Efstathiou et al. value for C, then the mass-to-light ratio at B needed to close the Universe
is

(MI LB)crit = 1450 h (MI LB)0' (2)


This value could easily be in error by a factor of two.
The observational uncertainties in MIL depend, of course, on the particular objects
considered. Throughout the rest of this paper I will consider measurements of MIL on
a variety of scales working from the small to the large. This progression will lead us
from dense systems which are almost certainly in dynamical equilibrium to large irregular
structures which are only a relatively weak perturbation on the background expansion. Not
surprisingly, equation (1) applies in rather different ways to the two types of system. A
comparison of the derived "no" over this wide range of scales gives some indication of the
extent to which our fundamental assumption holds, and thus of the kinds of "bias" (defined
as any systematic difference in the large-scale distributions of mass and light) which may
be present.

2 Dynamical Estimates of MIL ratios

2.1 SPIRAL GALAXIES AND THE TIMING ARGUMENT

The blue-band luminosity density of the Universe is dominated by spiral galaxies, so it is


natural to estimate spiral MIL ratios for insertion into equation (1). Unfortunately, the only
281

reliable indication of the mass of an individual spiral comes from measurement of its rotation
curve, and HI gas is very rarely observable beyond about twice the optical (Holmberg)
radius of a galaxy (Sancisi and van Albada 1987). Out to this distance rotation curves
are usually said to be flat, suggesting that the total mass is increasing linearly with radius
despite the fact that the total light has already converged. In fact, systematic deviations
from flatness now seem well established; observed rotation curves are predominantly rising
in faint and diffuse galaxies and can fall significantly in luminous, high surface brightness
systems (Salucci and Frenk 1989; Casertano and van Gorkom 1991). Perhaps faint galaxies
are dark matter dominated while the observed stellar population contains most of the mass
in the inner regions of bright concentrated galaxies. These subtleties provide interesting
clues to galaxy formation but have no significant effect on mass estimates. By combining
standard relations between blue luminosity, maximum rotation velocity, and optical radius
with the luminosity function of Efstathiou et al. (1988), I estimate a luminosity-weighted
mean M/ L of 20 h within two Holmberg radii of the centre of a spiral galaxy. Substituting
into equations (1) and (2) gives "no"= 0.014, so that the mass directly associated with the
optical regions of galaxies contributes less than 2% of the closure density.
One might hope to measure the mass associated with a spiral galaxy by analysing the
motions of its satellites. In practice this has proved to be difficult both because of the small
number of satellite systems available around any given spiral, and because of uncertainties
introduced by observational selection effects and by the unknown shapes of satellite orbits.
The difficulties are well illustrated by results for the best observed satellite system of all;
that of the Milky Way. Using a sample of eleven objects with reliable velocities, Little
and Tremaine (1987) estimated a Milky Way mass of 2.3 lO11Me, based on a point mass
potential and isotropically distributed orbital velocities. Adding good velocities for a further
five satellites, Zaritsky et al. (1989) obtained 12.5 1011 Me for the same model; the 90%
upper confidence limit from the old analysis was half the 90% lower confidence limit deduced
from the new data. Much of this change was due to a single satellite, Leo I, which has both
a large distance (230 kpc) and a large Galactocentric radial velocity (177 km/s). However
even excluding Leo I, Zaritsky et al. obtained twice the mass of Little and Tremaine. They
were able to show that any plausible bound orbit for Leo I required a Milky Way mass in
excess of 1012 Me, corresponding to a M / L of 50 or more. IT the Milky Way mass is so
uncertain, it is clearly hopeless to attempt mass measurements for individual distant spirals
where there are rarely more than two or three observable satellites. The best that can be
hoped for is an estimate of mean halo mass from the satellites of a large ensemble of spirals.
Such an ensemble has been observed by D. Zaritsky and is currently being analysed.
The argument which Zaritsky et al. use to obtain a lower limit on the mass of the Milky
Way from observations of Leo I, has long been applied to the M31/Milky Way system
(Kahn and Woltjer 1959). It is usually referred to as the "timing argument". With modern
data, the assumption that the Local Group can be treated as a two-body system leads to
a lower limit of 3 1012 Me on its total mass, corresponding to a lower limit of about 50 on
its M/ L ratio (Zaritsky et al. 1989). This gives "no"= 0.035 h- 1 , a value which is already
somewhat higher than the maximum baryon density allowed by some recent analyses of
cosmological nucleosynthesis (e.g. Olive et al. (1990) give n" < 0.015 h- 2 ). In fact, the
timing mass limit is sufficiently high that it implies that the halos of the two spirals must
extend over a significant fraction of the distance between them. It is thus unclear whether
282

a two-body orbit will be a good approximation to their relative motion. Several groups
have recently studied the formation of Local Group "lookalikes" in N-body simulations of
structure formation (Peebles et al. 1989; Moore and Frenk 1990; Kroeker and Carlberg
1990). If a sphere is centered on each galaxy in a lookalike with radius equal to half the
separation, the total mass in the two spheres is typically 61012 M 0 . Adopting this value for
the Local Group and using equation (1) suggests "no"= 0.07 h- I • Notice, however, that
all the simulations model flat universes, and that the effective mass per galaxy calculated
in the above manner is smaller than the global mean value in the simulations.

2.2 BINARIES AND GROUPS

A traditional way to estimate the mean mass associated with a galaxy has been through the
dynamical analysis of binary galaxies and groups of galaxies. The result is usually framed
in terms of an estimated mean M / L for a system, or more frequently for an ensemble of
systems. The mass determination problem is much more difficult for such objects than for
a binary star or a stellar group. Orbital times are very long compared to an observer's
lifetime; only the line-of-sight velocity component is measurable; the effective interaction
potential of the objects is unknown since most of the mass is unseen; and the relatively
low contrast of the systems with respect to their surroundings results in strong selection
effects when defining their membership. The M / L ratio of a system is estimated through
a relation of the form,

M rV2
T=k GL · (3)
For a binary, the characteristic radius must be taken as the projected separation and the
velocity as the redshift difference. For groups the classical virial theorem would suggest
taking r and V to be the luminosity-weighted mean harmonic separation in projection,
and the luminosity-weighted velocity dispersion. In practice, however, other estimators of
characteristic size and velocity usually have more stable statistical properties and lower
sampling variance. Whatever estimator is chosen, the distribution of the dimensionless
factor, k, depends on projection and selection effects, on the relative distributions of mass
and of galaxies, and on the shapes of galaxy orbits. A characteristic M / L for a system or an
ensemble of systems cannot therefore be obtained without careful modelling of these factors.
Because the distribution of k is typically very broad, the results are quite sensitive to this
modelling; meaningful results can be obtained for observational data only if the sample
selection and analysis are carefully calibrated against a model. This implies an inevitable
model-dependency in the result which can easily amount to a factor of two or three.
These problems are particularly bad for binary galaxies where the expected distribution
of k can easily span two orders of magnitude, is very skew, and is highly model-dependent.
Useful M / L values can therefore be obtained only for relatively large ensembles of pairs.
Most authors have analysed such ensembles by considering a superposition of orbits with
random orientations, phases, eccentricities, and semi-major axes; observational selection
then eliminates certain configurations (e.g. Turner 1976; van Moorsel 1987; Schweizer
1987). The problem is considerably easier to treat if one starts from the Jeans equations for
the ensemble as a whole (White 1981; White et al. 1983), but this approach has not yet been
283

generally adopted. The data from all recent studies of binary galaxies are in overall agree-
ment, but the conclusions drawn from them differ widely, ranging from confirmation of very
massive halos (Turner 1976) to tentative support for extended mass distributions (White et
al. 1983) to truncation of the mass distribution beyond the optical galaxy (Schweizer 1987).
These differences can all be traced to sensitivity to details of modelling, and as a result
binaries add little to the present discussion of the mean mass associated with a galaxy.
The situation is rather better for groups of galaxies. With more objects the scatter in
k is reduced, although it is still too large to estimate individual M / L ratios for systems
with less than about 10 members. Furthermore, while the typical separation of binaries
in published samples is less than 100 kpc, the groups have typical sizes of order 1 Mpc.
Thus their dynamics gives information about mass which is much more loosely associated
with individual galaxies than is the case for binaries. This looseness has some serious
disadvantages, however. Membership in groups is difficult to assign, and to some extent
arbitrary. The low density of groups implies that their dynamical times are long, and
they may therefore be far from equilibrium. Nolthenius and White (1987) found that the
median inferred collapse time for the groups in their catalogue was longer than the age
of the Universe. The same is true in the catalogues of Huchra and Geller (1982) and
Ramella, Geller and Huchra (1989), although these authors obscure the fact by defining a
"group dynamical time" which is only 8% of the collapse time for a homogenous sphere with
the same mass and binding energy. Both these effects make detailed modelling of group
selection and dynamics imperative for any interpretation of the data in terms of M / L ratios.
The best way to carry out such modelling would seem to be through the use of artificial
catalogues constructed in the same way as the real group catalogues from fully consistent
N-body calculations of clustering. This provides a means for evaluating the consequences
of observational selection effects, and of various assumptions about cosmology, about initial
fluctuations, and about the assumed relation between the mass and galaxy distributions.
In a first attempt at this program, Nolthenius and White (1987) found good qualitative
agreement, and moderate quantitative agreement between the properties of groups in the
CfA survey, and those of groups identified in artificial catalogues generated from simula-
tions of a Cold Dark Matter universe. Assuming an open universe in which the galaxies
are distributed in the same way as the mass, they found a best value, MILB = 220 h
corresponding to "no"= 0.13, with a probable systematic uncertainty of order a factor of
two. However, artificial catalogues generated from CDM models in which galaxies are as-
sumed to be a biased tracer of the mass in a flat universe agreed somewhat better with
the observational data than the open models in which mass follows light. We have recently
confirmed this result in a more detailed analysis of simulations with considerably higher
resolution (Moore, Frenk & White, in preparation). In these models only the positions of
galaxies are biased; their velocities are distributed in the same way as those of the dark
matter. As a result, the observational data require a strong overrepresentation of galaxies
in dense regions if there is to be consistency with n = 1. However, it has recently become
clear that dynamical effects can bias galaxy velocities to smaller values than are typical for
the mass in a group (Carlberg and Couchman 1989). Equation (1) is misleading in this
case, not only because the true MIL value of a group is unrepresentative of the universe
as a whole, but also because the mass calculated from galaxy velocities is a substantial un-
derestimate of the true group mass. Until the strength of this effect is properly established
284

in real groups, it is unclear whether their apparent M / L ratios can be trusted. This issue
needs to be resolved before one can make a confident attack on the question of how group
M / L ratios should be extrapolated to the Universe as a whole.

2.3 BlCB CLUSTERS OF GALAXIES

Projected positions and radial velocities can be measured for large numbers of galaxies in
rich clusters. As a result, the purely statistical uncertainties entering into the quantities
r and V in equation (3) are quite small. In addition, cluster masses can be estimated
independently from X-ray data. Unfortunately, substantial systematic uncertainties remain
in both techniques. Furthermore, it is unclear how to use equation (1) to extrapolate to a
global "no" because the galaxies which predominate in the richest and densest clusters are
ellipticals and SO's which have stellar populations with a blue M / L a factor of two or so
greater than that of the mean stellar population of the Universe.
Some of the remaining problems with optical mass determinations are a consequence of
observational selection effects. Rich clusters have traditionally been found as enhancements
of the galaxy surface density on photographic plates. Because they are rare systems, this
leads to a tendency for catalogues to contain objects whose richness is enhanced by the
superposition of a foreground or background group. Sometimes the superposed group is
at a sufficiently different redshift that it can clearly be separated from the main cluster.
However, in many cases it may be part of the same supercluster, perhaps falling into the
larger system, and the redshift separation is then insufficient to isolate it. This is a problem
which is again most easily studied by using N-body simulations to construct artificial cluster
catalogues in the same way as an observer constructs a real catalogue. Frenk et al. (1990)
suggest that many of the Abell clusters with measured velocity dispersions of 1000 km/s
or more are likely to have had their dispersions (and hence their masses) overestimated by
a substantial factor as a result of this effect.
If a cluster is regular and is not seriously affected by projection effects of this kind,
then with 100 or so redshifts it is possible to estimate its M/ L to within 10 or 20% if one
assumes that the radial distribution of mass parallels that of the galaxies. In the Coma
cluster, the system with the largest measured number of redshifts, this assumption leads to
M/ LB = 410 h (The & White 1986,1988). Furthermore, the run of velocity dispersion with
radius is then consistent with that expected for a locally isotropic distribution of velocities
within the cluster. Unfortunately, if the assumption that mass follows light is dropped, a
much wider range of possibilities exists, and the data are consistent both with near-radial
orbits within a weakly concentrated but massive dark matter distribution, and with near-
circular orbits in a strongly concentrated distribution of low total mass (The & White
1986; Merritt 1987). The reason for this indeterminacy is easily seen from the hydrostatic
equilibrium equation,

(4)

The galaxy density, Pg, can be obtained directly from observation, but redshift measure-
ments provide only a combination of the radial and tangential velocity dispersions, U r and
Ut. The remaining functional degree of freedom allows many mass profiles, M(r), to be
285

compatible with the data. For Coma, M(r) is best constrained near r = 1.0 h-1Mpc where
M = 6 1014h- 1Me with a total uncertainty which is probably less than 30% (The & White
1986; Merritt 1987). As Merritt (1987) shows, the shapes of the galaxy orbits in a spherical
equilibrium cluster could, in principle, be determined from the shape of the line-of-sight
velocity distribution. In practice, this is unlikely to work both because of the lack of suffi-
cient redshifts and because the distributions in real clusters are perturbed by infall and by
substructure. Such irregularities are also likely to prevent reliable determinations of cluster
masses out to the large radii where characteristic caustic patterns are expected for spherical
infall from an expanding universe (Regas & Geller 1989).
The situation is simpler, at least in principle, when determining cluster masses from
X-ray data. Accidental or selection induced superpositions are expected to be negligible,
and the relevant equilibrium equation,

dp GM(r)
Pgaa dr = --r-2-' (5)

is simpler than equation (4) because both the density, Pgaa(r), and the temperature, T(r),
of the gas can be obtained directly from observation. In practice, current data give good
measurements of the density out to 0.5 h- 1Mpc, and occasionally further, and they also
determine the luminosity-weighted mean temperature well. However, the surface brightness
is too low to measure at larger radii, and there is rather little information about possible
radial variations in temperature. Again, the most detailed observational information is
available for the Coma cluster where the emission has been traced beyond 0.7 h-1Mpc in
some directions, and there are a number of high SIN spectroscopic observations. The data
suggest rather little radial variation in temperature, and they agree well with the simple
mass follows light model for the optical data discussed above (Hughes 1989). This happy
situation is by no means universal; for example there is a serious discrepancy between the
very large velocity dispersion measured for the Perseus cluster and its relatively modest
X-ray temperature.
Taking the MIL estimated for the inner 1.0 h-1Mpc of Coma from both optical and
X-ray data, and applying it to the Universe as a whole, we get "!lo"= 0.28. Because of
the difference in stellar populations, a value about a factor of two smaller would result if
we assumed instead that the stellar mass to total mass ratio in this region is the same as
the global value. Thus galaxy formation must be very substantially biased towards such
regions if indeed !l = 1. A strong enough effect is seen in CDM models with a substantial
positional bias (Frenk et al. 1990), but it seems unlikely that such low apparent MI L values
could be caused by velocity bias of the kind discussed by Carlberg and Couchman (1989).
The latter would have particular difficulty accounting for the low mass derived from X-ray
data and its agreement with the optical value.
The optical and X-ray data on rich clusters can, however, be used to argue directly
against !l = 1 in the following fashion. For galaxies like those which dominate the inner
regions of Coma, optical measurements imply an internal MILB ratio of 19 h-l(Binney
& Tremaine 1987; Table 4.2). The gas mass within 1.0 h- 1 Mpc is 2.9 1Q13h- 2 .5 Me (The
& White 1988; Hughes 1989). Thus the ratio of baryonic to total mass within this region
is at least 0.046 + 0.048 h-1.5 (some of the unseen mass might also be baryonic). This
can be compared with the limit, !lb < 0.015 h- 2 , claimed by Olive et al. (1990) on the
286

basis of calculations of Big Bang nucleosynthesis. H the ratio of baryons to total mass in
the inner parts of Coma is equated to that in the Universe as a whole, we find nil l >
3.1 h2 + 3.2 hO. 5 • The Universe is open by a large margin. This conclusion can only be
averted if the nucleosynthesis limits are greatly relaxed, or if a mechanism is found to
concentrate the baryons in rich clusters by a large factor relative to the total mass.

2.4 COSMIC VIR1AL THEOREMS

A significant part of the uncertainty in determining an appropriate mean M / L for groups


comes from difficulties in deciding on membership. This problem can be avoided by using
equations which describe the statistical equilibrium of the clustering distribution as a whole.
The shift in emphasis is similar to that involved in moving from a description of binary
galaxy dynamics in terms of a superposition of individual orbits to one based on suitable
moments of a Boltzmann equation (see above). H the clustering is assumed to be time-
independent, the appropriate equilibrium equation is directly analogous to (4):

(6)

In this equation 411"np(r)r 2 dr is the number of galaxy pairs per unit vol~me with separations
between r and r + dr, O'r( r) and O't( r) are the relative velocity dispersions of the ensemble
of such pairs in directions parallel and orthogonal to their separation, (p) is the mean mass
density of the Universe, and (p}P(xlr)dV is the mass in a volume element dV located at
position x with respect to a pair member and averaged over all pairs with separation vector
in a given (arbitrary) direction. The factor of 2 on the rhs reflects the fact that both
members of the pair are, on average, accelerated by the force calculated in the integral.
The fundamental assumption underlying this equation is that the distribution of pairs with
separation, r, evolves on a timescale which is long compared to their orbital time. It should
be borne in mind that this assumption can at best hold only approximately, and that it
may, perhaps, be seriously in error (see the discussion of velocity bias above).
An equation similar to (6) was first derived by Peebles (1976). The quantity, np(r),
is simply related to the observable galaxy two-point correlation function. Furthermore, a
combination of O'r and O't is also observable. Thus if an assumption is made about the
relative size of the two dispersions, for example O'r = 0'" the left hand side of the equation
is known. Peebles pointed out that if we are prepared, in addition, to assume that the
relative distribution of mass about a pair of galaxies has the same statistics as the relative
distribution of other galazies, then the function P can be expressed as a combination of the
observed two- and three-point correlation functions. In this case equation (6) can be solved
for (p). A consistency check on this last assumption is provided by the fact that a single
value of (p) should bring the two sides of the equation into agreement for all relevant r.
Davis and Peebles (1983a) showed that the observed scaling of galaxy correlation functions
and pairwise velocity dispersions does indeed satisfy this test, and that the inferred density
corresponds to "no"= 0.2.
It is difficult to assess the reliability of this argument because of uncertainties in the
validity of the two underlying hypotheses - that clustering is statistically stable and that the
galaxy-galaxy-mass and galaxy-galaxy-galaxy three-point distributions are identical. Both
287

assumptions are clearly violated in numerical simulations which attempt to model galaxy
formation and clustering in a flat CDM universe (Carlberg & Couchman 1989). Equation
(6) can plausibly be applied on scales, r < 1.0 h- 1 Mpc, and Davis and Peebles's result, as
expected, is similar to that found when groups and clusters are treated individually.
A different and equally interesting cosmic virial theorem was proposed by Fall (1975)
on the basis of the Layzer-Irvine equation,

!(K + W) +H(2K + W) = o. (7)

This equation relates the kinetic energy density in peculiar motions, K, the gravitational
potential energy density in deviations from uniformity, W, and the Hubble parameter, H,
and is an expression of conservation of energy in expanding coordinates. It is easy to
show that under quite general conditions the energy densitites should satisfy the bounds
0.51WI < K < IWI. Since K/(p) can be estimated from the rms peculiar velocity of
galaxies relative to the comoving frame, and W / (p)2 can be estimated as an integral over the
two-point galaxy correlation function, these bounds provide a way to constrain (p) under
the assumption that mass and galaxies have similar correlations and peculiar velocities.
Fall adopted an rms peculiar velocity of 300 km/s and obtained "flo"- 0.03. Today it
might seem reasonable to adopt the Milky Way's observed motion of 600 km/s through the
microwave background as a "typical" value, resulting in a four times larger density. This
difficulty in obtaining the appropriate velocity is the main disadvantage of the method. Its
main advantage is that the major contributions to both K and W come from relatively large
scales (in linear theory from scales where the power spectrum IOkl 2 varies approximately
as k- 1 ) and so the method is sensitive to the relative distributions of mass and galaxies
over much larger regions than those addressed by the other methods discussed so far. In
fact, Fall's method is more closely related to the quasilinear methods I discuss in the next
section.

2.5 DYNAMICAL TESTS ON LARGE SCALES

Galaxy clusters are the largest objects in the Universe which are gravitationally relaxed,
and thus the largest objects for which one can hope to determine masses using the virial
theorem. However, structure in the galaxy distribution is clearly seen on much larger
scales, and if the mass distribution parallels that of the galaxies, it must induce significant
streaming motions over large volumes of space. Such motions are usually analysed using
linear theory, or some generalisation which allows extrapolation into the mildly nonlinear
regime. Such techniques work in the limit where the induced velocities cause galaxies to
move a small fraction of the overall scale of the structure in a Hubble time. They have been
applied to estimate the mass associated with the Local Supercluster (Davis and Peebles
1983b), with the "Great Attractor" (Lynden-Bell et al. 1988), and clustered like IRAS
galaxies (Yahil 1988). With a suitable "mass follows light" assumption each application
leads to an estimate of "flo". I do not wish to comment in detail on this field, which is
currently very active. Updates can be found in the contributions of Lynden-Bell, Frenk,
and Dekel to this volume. However, I would like to make a point related to the kind of
"bias" needed to reconcile the results of such studies (which consistently give "flo < 1" for
288

mass following light) with a flat universe.


Consider the linear theory model for "infall" into a spherical density enhancement.
This has often been used to describe the retardation of the cosmic expansion in the Local
Supercluster (e.g. Davis and Peebles 1983b). The peculiar velocity induced at the position
of the Local Group can be written as,

(8)

where (15m ) and (6L) are the mean overdensities in mass and light within a sphere centred
on Virgo and passing through the Local Group, d is the Virgo distance, and b is a biasing
parameter defined as the ratio of the two overdensities. As illustrative values let us take
Vp = 250 km/s, Hod = 1350 km/s, and (6L) = 2. IT we assume mass follows light, then
b = 1 and "no"= 0.12. The Universe is open by a factor of eight. Alternatively, if we
assume no = 1, then b = 3.6 and the M / L of the material within the Local Supercluster is
smaller than that ofthe Universe as a whole by a factor, (1 + (15m ) )/(1 + (6L» = 0.52, i.e.
by less than a factor of two. Applying the same model to the Great Attractor we might
take Yp = 535 km/s, Hod = 4200 km/s, and (6L) = 0.58. In this case we find "no"= 0.5
if mass is assumed to follow light, but an M/ L for the material in the Great Attractor
region which is only 13% below the global mean under the alternative hypothesis that the
Universe is, in fact, flat. Thus, when one is working in the quasilinear regime, a relatively
small modulation of galaxy properties can be translated into a much larger error in "no".
Since many properties of galaxies are known to be modulated by clustering, it is important
to be wary of such systematic effects when using these techniques to estimate no.

3 Conclusions

In this review I have tried to emphasize the fact that measurements of galaxy clustering
tell us very little about no unless we are prepared to assume that mass follows light.
Unfortunately the available evidence gives us very little reason to be confident in this
assumption, since the two distributions are clearly very different on scales smaller than a
few hundred kpc. On the other hand there is no convincing evidence against the assumption
on larger scales. In fact, somewhat different assumptions are required on nonlinear scales
(clusters and smaller), where any bias depends on variations in M/ L, and on larger, quasi-
linear scales, where bias depends on the ratio of the fluctuation amplitudes in mass and light.
This difference is at the root of the surprising sensitivity of "no" to systematic variations
in galaxy properties which is discussed in the last section. Although I am personally rather
sceptical that the density of the Universe will ever be truly measured by dynamical methods
of the kind discussed in this article, it is clear that such studies are providing ever more
precise details of the relative distributions of light and mass. This may be the only available
route to an understanding of the nature of the dark matter and of its role in cosmogony.

Acknowledgements. This research was supported in part by NSF grant AST-8822297 to the Uni-
versity of Arizona
289

References

Binney, J., & Tremaine, S.D. 1987 Galactic l}pamic" Princeton Univ. Press.
Carlberg, R.G., & Couchman, H.M.P., 1989. A,trophy,. J., 340, 47.
Casertano, S. & van Gorkom, J.H., 1991. A,tron. J., 101, 1231.
Davis, M. & Peebles, P.J.E., 1983a. A,trophy,. J., 267,465.
Davis, M. & Peebles, P.J.E., 1983b. Ann.Ret/.A,tr.A,trophy,., 21, 109.
Efstathiou, G., Ellis, R.S.,& Peterson B.A., 1988. Mon. Not. R. tIItr. Soc, 232, 431.
Fall, S.M., 1975. Mon. Not. R. tIItr. Soc, 172, 23P.
Frenk, C.S., White, S.D.M., Efstathiou,G., & Davis, M., 1990. Alltrophy,. J., 351, 10.
Huchra, J.P., & Geller, M.J., 1982. A,trophy,. J., 257, 423.
Hughes, J.P., 1989. A,trophy,. J., 337, 21.
Kahn, F.D. & Woltjer, L., 1959. Alltrophy,. J., 130, 705.
Kroeker, T., & Carlberg, R.G. 1990, CITA preprint.
Little, B. & Tremaine, S.D., 1987. Alltrophy,. J., 320, 493.
Lynden-Bell, D., Faber" S.M., Burstein, D., Davies, R.L., Dressler, A., Terlevich, R.J., & Wegner,
G., 1988. A,trophy,. J., 326, 19.
Maddox, S.J., Sutherland, W.J., Efstathiou, G., Loveday, J.,&; Pete!SOn, B.A., 1990. Mon. Not. R.
tIItr. Soc, 247, 1P.
Merritt, D., 1987. AlltrophYIl. J., 313, 121.
Moore, B., & Frenk, C.S., 1990.- in Dynamic, and Interactionll 0/ Galazie" p. 410, ed. Wielen, R.,
Springer.
van Moorsel, G.A., 1987. A,tr. Astrophy,., 176, 13.
Nolthenius, R., &; White, S.D.M., 1987. Mon. Not. R. tIItr. Soc, 235, 505.
Olive, K.A., Schramm, D.N., Steigman, G., &; Walker, T.P., 1990. Physics Letters B, 236,454.
Peebles, P.J.E., 1976. A,trophy,.J.Lett., 205, LI09.
Peebles, P.J.E, Melott, A.L., Holmes, M.R., & Jiang, L.R., 1989. AlltrophYIl. J., 345, 108.
Ramella, M., Geller, M.J., &; Huchra, J.P., 1989. A,trophYIl. J., 344, 57.
RegOs, E. &; Geller, M.J., 1989. A,tron. J., 98, 755.
Salucci, P. &; Frenk, C.S., 1989. Mon. Not. R. tIItr. Soc, 237, 247.
Sancisi, R., & van Albada, T.S., 1987. in Dark Matter in the Unit/er,e, p. 67, ed. Kormendy, J. &;
Knapp, G.R., Reidel.
Schweizer, L., 1987. AlltrophYIl.J.Supp., 64, 427.
Turner, E.L., 1976. A,trophy,. J., 208, 304.
The, L.S., &; White, S.D.M., 1986. Alltron. J., 92, 1248.
The, L.S., & White, S.D.M., 1988. A,tron. J., 95, 15.
White, S.D.M., 1981. Mon. Not. R. tIItr. Soc, 195, 1037.
White, S.D.M., Huchra, J., Latham, D., & Davis, M., 1983. Mon. Not. R. tIItr. Soc, 203, 701.
Yahil, A., 1988. in Large-Ilcale Motionl in the Unit/er,e, p. 219, ed. Rubin, V.C., &; Coyne, G.V.,
Princeton.
Zaritsky, D., Olszewski, E.W., Schommer, R.A., Peterson, R.C., &; Aaronson, M., 1989. Alltrophys.
J., 345, 759.
290

DISCUSSION:

Hoffman: How does dynamical friction affect mass determination of groups and clusters
and what is the mechanism behind the velocity bias reported by R. Calberg?
White: Dynamical friction can cause massive galaxies to sink towards the centres of
groups and clusters. This segregation then results in an underestimate of M/L when stan-
dard techniques are used. I doubt that the effect could be larger than a factor of two in
rich clusters, but it might be more important in groups. The velocity bias discovered by
Carlberg, Couchman and Thomas might be related to dynamical friction, but I am not yet
fully convinced by this interpretation. I !m! convinced that the velocity bias is a real and
significant effect, and that it can result in a substantial bias in the M/L ratios for loose
galaxy group.

Lynden-Bell: All methods of determining 11 via the light of luminous objects end up by
multiplying a Mass/Light ratio by a luminosity density. All such methods really neglect
large volumes of the universe in which nothing shines. If these volumes are not empty 11 is
likely to be larger than estimated.
White: I agree. The equation 11 = (M/L) x C/ Pcrit assures that the galaxies which
contribute to C are each, in the main, associated with the same amount of mass as in the
systems for which M/L is determined. In fact, there is no obvious reason why these systems
should be typical of the Universe as a whole. The "voids" could easily contain much more
mass that one would guess from their galaxy population. However, the fact that peculiar
motions nearby agree roughly with those predicted from the density distribution of observed
galaxies is some indication that we are not missing large mass chunks in the voids, although
a substantial near uniform component could be present.
Liddle: I don't much like the introduction of a cosmological constant as dark matter,
but it seems that a lot of these limits would fail to pick up the cosmological constant
contribution to 11. For instance, mass would certainly not be traced by galaxies, and of
course the cosmological constant energy density will neither cluster nor virialise. Do you
think that the upper limits on 11 are significantly weaker in this case?
White: The virial estimates of M/L ratios which I have been discussing could be almost
entirely unaffected by a cosmological constant because they refer to objects with densities
well above the critical density. My estimate, "11", therefore omits any contribution from a
cosmological constant or any other near-uniform contribution to the cosmic energy density.
Only on larger scales, where we observe dynamical effects in the quasilinear regime, does a
cosmological constant begin to affect inferences from the motions of galaxies.

Shanks: You said that Peebles' version of the cosmic virial theorem was no better than
stacking velocity histograms from galaxy groups. But doesn't the Peebles method at least
circumvent the difficult business of actually having to identify clusters?

White: Fair comment. My point was that in practice one has to make a detailed model of
291

some kind in order to deal with the relationship between the dark matter distribution and
the galaxy distribution. This problem is no easier to solve for statistical virial theorems
than for individual groups. Indeed, I find it easier to think of plausible possibilities for
the structure of individual groups (radial versus circular orbits, smooth versus lumpy dark
matter, mass following light or extended in dark group halos ... ) than for the structure of
the corresponding two and three point functions.
ROSAT OBSERVATIONS OF CLUSTERS OF GALAXIES

H. Bohringer, W. Voges, H. Ebeling, R.A.


Schwarz, A.C. Edge, U. G. Briel, & J. P. Henry

Max-Planck-Institut fur Extraterrestrische Physik


D-8046 Garching
F.R. Germany

ABSTRACT. The expectations for observations of clusters of galaxies with ROSAT in particular
with respect to the All Sky Survey are discussed. A look at the first survey results suggests that
about 4000 to 8000 clusters should be detected in the All Sky Survey. The final catalogue of X-ray
selected clusters of galaxies will probably be quite different from the present optical catalogues. Due
to the higher spatial resolution and sensitivity compared to previous instruments a more detailed
investigation of the internal structure of clusters will be possible with ROSAT as is shown for Perseus
and Abell 2256.

1 Introduction

Clusters of galaxies are the largest gravitationally bound entities that have already de-
coupled from the universal cosmic expansion and are in the process of forming virialized
objects. The fact that a large number of observed clusters of galaxies are not well virial-
ized seems to imply that clusters are actually forming at the present epoch, in contrast to
galaxies for example which had the most active period of formation in the past. These two
facts make clusters of galaxies very interesting probes for the large scale structure in the
primordial density fluctuation field from which they evolved. The present knowledge of the
size distribution and the spatial distribution of clusters of galaxies is not as advanced as
the information on the galaxy distribution from the large redshift samples (e.g. presenta-
tions by Huchra, Frenk, and Dekel at this workshop). Therefore cosmological implications
on the primordial fluctuation spectrum have been mainly derived from the galaxy redshift
samples (e.g. Saunders et al. ,1991). Nevertheless clusters of galaxies will become impor-
tant tracers of the structure on even larger scales than explored up to date and will also
provide important information on the primordial spectrum on scales of the order of ten
Mpc through their size distribution and internal morphology.
Even though thousands of clusters of galaxies have been catalogued notably by Abell
(1958) and Abell, Corwin, and Olowin (1989) not so many redshifts have yet been obtained
for these clusters that can be used for a spatial correlation analysis (Bahcall and Soneira
1983; Bahcall, 1988; Sutherland, 1988; Huchra et al. , 1990). More importantly the selection

293
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 293-308.
@ 1991 Kluwer Academic Publishers.
294

criteria for the detection of clusters on photographic plates are poorly understood and it is
even unclear how many of the objects detected as rich clusters on the plates are actually
chance projections of a series of smaller galaxy groups on the plates (Frenk et al. , 1990).
Therefore observations are needed that lead to a large statistical sample of galaxy clusters
with well understood selection rules. This goal may be achievable with X-ray observations.
Clusters of galaxies are - except for quasars - the most powerful X-ray sources in the
sky with luminosities of 1043 to 3 . 1045 erg S-I. The emission originates from hot gas that
fills the gravitational potential well of the cluster close to hydrostatic equilibrium. The X-
ray surface brightness is roughly proportional to the square of the galaxy density and can
be even more peaked in clusters with cooling flows. Clusters of galaxies can therefore be
detected in X-rays as three dimensional, gravitationally bound entities. The observations
are very much biased to the dense core of the clusters and chance projections are hardly a
problem. Thus the ambiguity adherent to the optical observations is practically removed
in X-ray surveys. So far only a few hundred clusters have been observed in X-rays mostly
with no underlying strategy for the construction of a statistically complete sample. The
two presently available statistical samples of X-ray selected clusters from HEAOI/EXOSAT
observations (Lahav et al. , 1989; Edge et al. , 1990) and from the EINSTEIN Medium
Sensitivity Survey (Gioia et al. , 1990a, b) comprise of the order of 50 and 100 clusters,
respectively. Nevertheless these samples have been found to be very interesting in particular
in showing a trend of evolution in the X-ray luminosity function of clusters with redshift.
Thus there is a great potential for cosmologically relevant observations of clusters of
galaxies for future X-ray missions. Especially the first All Sky Survey with an imaging
X-ray telescope which is currently conducted by ROSAT, where thousands of clusters of
galaxies should be detected, will provide the basis to construct large statistically complete
samples of X-ray selected clusters of galaxies. In addition pointed observations of individual
clusters with ROSAT will help to understand the relation of observable quantities like X-ray
luminosity, spatial extent, X-ray temperature and morphological features. It is especially
important to understand how these observables are related to the cosmologically interesting
parameters as mass and the form of the gravitational potential of the cluster. This can be
learned from detailed observations and modelling (e.g. Hughes, 1989).
This paper provides some details about the ROSAT observatory and the observing
mission in section 2; first results form the ROSAT All Sky Survey with respect to clusters
of galaxies are described in section 3; the first results on individual clusters are discussed
in section 4, and section 5 gives a summary and an outlook on cosmological research that
can be conducted with ROSAT observations of clusters of galaxies.

2 Prospects of the ROSAT Mission

The energy radiation of the hot intracluster gas in clusters of galaxies peaks at one to a
few keY. Clusters are also very large and can be observed as extended objects out to large
distances. Therefore an imaging soft X-ray observatory like ROSAT which was launched
on June 1, 1990, is ideally suited for the study of clusters. The main instrument on board
of the ROSAT observatory is the X-ray telescope that consists of a fourfold nested Wolter
type mirror configuration. It covers an energy range from 0.07 to 2.4keV. The second
295

100

.e-
0>
0"
til
til
Os 1.0
II)

0.1 0.5 2.0


energy (keV)

Figure 1: PSPC X-ray spectra for optically thin thermal plasma at temperatures of 3.106
K, (a), 1.107 K, (b), 3 .10 7 K, (c), and 1.108 K, (d).

instrument is the Wide Field EUV Camera with a spectral window from 20 to 300 eV. Due
to the strong absorption of the interstellar medium of the Galaxy in the EUV band the
Wide Field Camera is only of limited application for extragalactic objects such as clusters.
The X-ray telescope carries two types of focal plane detectors. The Position Sensitive
Proportional Counter (PSPC) has a spatial resolution better than 30 arcsec within a radius
of 20 arcmin around the telescope axis. It has a limited energy resolution which is roughly
about 45 % FWRM at 1 ke V and the relative resolution varies roughly inversely proportional
to the square root of the energy. The second type of detector, the high resolution imager
(RRI), has a better spatial resolution of less than 5 arcsec near the telescope axis. This
detector has no energy discrimination, however, and is about half as sensitive.
The X-ray emitting hot intracluster gas is optically thin and has a temperature roughly
equivalent to the virial temperature of the cluster potential well. The expected and actually
observed temperatures range from 2 to 10 keY. Fig. 1 shows the spectra for hot plasma
with temperatures in this range normalised to the same emission measure. One notes that
in the temperature range relevant for clusters of galaxies the spectra are quite similar.
The differences are becoming larger only for the lower temperatures. The ROSAT energy
window is unfortunately too small to observe the exponential cutoff in the Bremsstrahlung
spectrum for the high cluster temperatures. Also the absolute energy flux integrated over
the ROSAT window (defined as 0.1 to 2.4 keY) is not very sensitive for the gas temperature.
Therefore temperatures of the cluster gas can only be determined for spectra of very high
quality. This effect has the advantage on the other hand that the X-ray flux is directly
related to the emission measure and thus gas densities can be determined without having
precise temperature information.
296

103 100

...
l1li
cu 102

'iI=
10
-...
l1li
l1li
u

be
....0 10 1 ~
Ul
...cu ~
1.0 0.1
~
=
= 0.1 0.01
0 0.2 0.4 0.6 0.8 1.0
redshift

Figure 2: Comparison of the depth of optical versus X-ray cluster surveys. The continuous
line gives the number of galaxies that can be observed on an optical plate with a limiting
magnitude of mv = 20 for a Coma type cluster with an Abell number of galaxies of about
100. The dashed lines give the ROSAT PSPC count rates for clusters with a luminosity of
1044 and 1045 erg S-1 in the 2 - 10 keY band.

Compared to optical observations X-ray detections of clusters of galaxies are biased to


richer and more compact clusters which can more easily be detected out to larger distances.
Fig. 2 compares the dependence of the X-ray lux on distance to the number of galaxies
that can be seen on a photographic plate down to mv = 20 mag for a Coma type cluster
(with an Abell number of galaxies equal to 100). The X-ray luminosities for the two X-ray
lux curves are 1044 and 1045 erg s-1 (in the 2 - 10 keY band), the typical luminosity range
for rich clusters. For the typical exposure time of the ROSAT survey with tv 500 sec very
luminous clusters can be seen out to z tv 0.5 while a distance of z = 1 can be reached
with deeper exposures. One also notes the relatively sharp cut off for the optical plates
compared to the long tail in the X-ray flux curves.
After two month of calibration and verification observations half a year of the ROSAT
mission was devoted to an All Sky Survey which was completed at the end of Jan. 1991
except for a few open patches due to spacecraft and instrumentation problems which will
be scanned later in the mission. During this survey the sky was scanned by the ROSAT
telescope in great circles in a plane perpendicular to the solar direction. Following the sun
the whole sky is thus covered in half a year. Due to the overlap of the scan circles at the
ecliptic poles the exposure is much greater at the poles. The instrument has to be shut
down frequently when the orbit penetrates the radiation belts. This effects the southern
sky more severely than the north due to the South Atlantic Anomaly in the Earth magnetic
field. The minimum exposure time in the equatorial regions is about 500 sec while exposure
297

100
>.
~
ell
80
eo..
0 60
= 40
GI
u
J.o
GI
!:l.
20

0 500 1000 1500 2000


exposure time (sec)

Figure 3: Exposure time histogram for the ROSAT All Sky Survey

times of several 10 000 sec are reached at the ecliptic poles. Fig. 3 gives an exposure time
histogram for the entire sky as predicted from preflight simulations.
Using known X-ray luminosity functions from previous surveys one can now estimate the
number of clusters of galaxies that should be detected during the ROSAT All Sky Survey.
Assuming that 15 to 30 photons are sufficient to detect a cluster the depth of the All Sky
Survey is given by a flux limit of 5 .10- 13 to 10- 12 erg cm- 2 s-l for areas in the sky where
interstellar hydrogen column density is less than 4 .10 20 cm- 2 • In an area of '" 30 deg 2 at
the ecliptic poles the flux limit is a good order of magnitude lower.
Fig. 4 shows the expected number of cluster detections in redshift bins for a flux limit
of 5.10-13 erg cm- 2 s-l for an area of 8 ster of the sky (excluding a 40 deg wide strip
around the galactic plane) and for a flux limit of 5.10- 14 cm- 2 S-l in the area of 30 deg 2
around the ecliptic poles. For the calculations an X-ray luminosity function for the 2 - 10
ke V energy band of

n(Lx) = 3.10- 7 exp(L x /8.2) L;1.6 (1)


has been used where Lx is in units of 1044erg S-l and n(Lx) in units of Mpc 3 .
Eq. (1) is a good fit to the combined results for HEAO 1 of Kowalski et al. (1984) and
the EXOSAT/HEAO 1 results of Edge et al. (1990). A temperature - flux relation of
Lx = 24· (T/I0 8K)2.7 was used to convert these fluxes into fluxes for the ROSAT band.
Standard cosmological paramters (Ho = 50 km Mpc 1 s-l, n = 1, and A = 0) were used
for the calculations. The results show that about 4000 to 8000 clusters of galaxies should
be found in the All Sky Survey for a flux limit between 5.10- 13 and 10- 12 erg cm- 2 S-l.
The above calculations do not take any evolutionary effects of clusters of galaxies into
account. If such effects are taken into account for the rich clusters as they were found in the
298

100

10

0.5 1.0 1.5 2.0


redshift

Figure 4: Histogram of the expected number of cluster detections in the All Sky survey as
a function of redshift for a flux limit of 5.10- 13 erg cm- 2 s-l for the entire sky and for a
flux limit of 5.10- 14 erg cm- 2 S-l for an area of 30 deg 2 around the ecliptic poles.

EINSTEIN Medium Sensitivity Survey (EMSS; Gioia et al. , 1990 a,b) and in the HEAO
1jEXOSAT data (Edge et al. , 1990) the histograms have a steep cutoff at around z rv 0.5
and z rv 0.9, respectively. We also did not account for the effect that for nearby clusters
the detect area is larger and thus for a given background flux the detection efficiency will
decrease for the nearby more extended clusters. In summary most of the clusters to be
detected in the survey will have redshifts below 0.3 but a few rich clusters out to z rv 1
should also be found.

3 First Results from the All Sky Survey

The data handling and processing for the All Sky Survey is in the first period of the analysis
organized in great circular strips of a width of 2 deg. on the sky. So far only a small part
of the survey data has been processed by the standard analysis software. For the following
statistical analysis only three 2 deg. strips from an early test survey during the verification
phase ("MINI Survey", MS) and five 2 deg. strips from the main survey, that have been
analysed by a preliminary version of the standard analysis system, will be used. These two
survey parts cover approximately 8 % of the sky.
A total of 3431 sources were detected in these strips which by extrapolations corresponds
to an expected detection of 43 000 sources over the entire sky. It has been found, however,
that systematic effects in the current detection software are affecting the detection efficiency
so that the number of detected sources is subject to change. One can estimate that the flux
299

limit for the detection efficiency achieved so far is somewhere near 10- 12 erg cm- 2 s-l.
Among these sources 78 were immediately identified as catalogued clusters of galaxies:
48 Abell/ ACO clusters, 19 Zwicky clusters, 6 clusters discovered in the EMSS, 2 known low
galactic latitude clusters, and 3 clusters around prominent radio galaxies. These identifi-
cations were found through a coincidence of the X-ray position with the catalogued cluster
position within a distance of 3 arcmin and the absence of another obvious X-ray candidate
from published catalogues. Only some of these identifications have been checked by detailed
comparison of the optical plate image with the X-ray map and we estimate that up to 5
of the above identifications could be wrong. A closer look at the statistics of the distances
between X-ray positions and positions of catalogued clusters shows that clusters are still
detected with an offset up to 6 arcmin from the X-ray positions for Abell/ ACO clusters
and 10 arcmin for Zwicky clusters, respectively. At this distance the chance coincidences
introduce an uncertainty in the individual identification and all those cases have to be in-
spected in detail. Fig. 5 shows a background corrected cumulative histogram of the Abell
and Zwicky clusters found around the X-ray sources as a function of the offset in the 5
strips of the main survey.
Of the Abell/ ACO clusters 14.5 % belong to the distance class 3, 12.5 % distance class
4, and 73 % to distance classes 2: 5; while the richness class distribution is 27 %, 67 % , 6
% for richness class 0,1,2, respectively. As expected the nearby clusters are relatively over
abundant while richness class 0 clusters are under represented. Among the Zwicky clusters
the more compact ones prevail. The distance distribution of the EMSS clusters roughly
corresponds to the redshift histogram of Fig. 4.
For one of the data strips a detailed search for all known Abell and Zwicky clusters was
performed. In this analysis the clusters with larger offsets where individually inspected and
additional sources were searched for by alternative detection techniques and by eye. For
271 (109 Abell/ ACO) clusters in this strip only 36 (14 Abell/ ACO) clusters were found to
be X-ray sources at a significance for their existance of more than 20". Some uncertainty
remains because of the variation of the detection efficiency of the different methods used.

3.1 X-ray properties of the detected clusters

Clusters of galaxies should be found among the hardest X-ray sources in the survey. Only
neutron stars and young supernova remnants which are confined to the Milky Way and
nearby galaxies have harder X-ray spectra. One also expects to detect a large fraction
of the clusters as extended objects. For the identified clusters the number of registered
photons is 8 - 20, 21 - 39, 2: 40 for 36, 16, and 26 of the clusters, respectively. Only for
the objects with more than 40 counts a good characterization of the X-ray properties is
possible which is given in Table 1. Only about 25 % of the known clusters can be detected
by their extent. The fraction of newly detected clusters that can be selected by their extent
is probably even smaller.

3.2 Optical identification of clusters

The identification of ROSAT X-ray Sources which do not coincide with catalogued objects
is in a next step based on the correlation with optical data from scans of the Palomar and
300

aIf') ..J

~ a
,.Qa I- -
~
Q)
u
~ a -
o
III
If')

o 10
Distance [arc min]

Figure 5: Background corrected cumulative histogram of Abell/ ACO and Zwicky clusters
found around the ROSAT X-ray sources as a function of the distance between the catalogued
position and the X-ray source.

Sample hardness extent


hard undefined positive possible negative
2: 40 cts 21 5 18 2 6
< 40 cts 36 16 2 4 46

Table 1: X-ray properties of ROSAT sources identified as clusters of galaxies. The number
of clusters found for each category is given.
301

Sample Abell (> 30-) > 20- > 1.20- > lo-
a 3 8 17 4
b 3 5 12 3

Table 2: Cluster candidates found in a test area of 0.4 % of the sky. The categories are
labled according to the significance of the galaxy overdensity within a radius of 90 arcsec
and the number of clusters found for each category is given. Sample a contains all X-ray
sources with galaxy overdensities while sample b does not contain the cadidates that have
another likely X-ray candidate or sources in very crowded fields.

UK Schmidt Survey plates in collaboration with STScI (Baltimore), ROE (Edinburgh), and
NRL (Washington). So far for one data strip of the MS optical data from the COSMOS
scans of the UK Schmidt survey (e.g. Heydon-Dumbleton et al. , 1989) for a small region
around the X-ray position were used for the further identification of clusters. Excluding
the galactic plane the data cover about 35 % of the strip area. We searched for galaxy
density enhancements locally around the X-ray source in search circles of 50 and 90 arcsec
and calculated the significance of the overdensity above the regional background. The three
Abell clusters in this area all had overdensities at a ;::: 30- significance level for the 90 arcsec
radius. Further fields were found with overdensities between 10- and;::: 20- as given in Table
2. Sample a gives all sources with overdensities while the sources which also have a bright
star or another "contaminating object" have been removed from sample b. There is quite a
large number of possible candidates for cluster X-ray sources in this survey strip where the
highest category seem to be very good candidates and the rest possible cluster candidates.
The further cathegorization of these X-ray sources relies on deeper optical images and
spectroscopic observations of these groups of galaxies which will be conducted for at least
a representative part of the sky.

3.3 Expectations for the All Sky Survey

In the EMSS which is about comparable in depth to the All Sky Survey, 12 % of the sources
detected were clusters of galaxies. Scaled to the above estimated total number of detected
sources in the ROSAT survey one should find about 5000 clusters, which is consistent
with the theoretical predictions above for a more conservative flux limit. From the present
analysis one only expects 1000 to 2000 of these sources to be known clusters, mainly clusters
from the Abell/ ACO catalogue. A larger number of clusters should be newly discovered.
Extrapolation from the optical identification above leads to 1300 good cluster candidates
and about 3000 possible cluster candidates for the whole sky. Distant unknown clusters
not visible on the optical plates constitute another smaller sample of cluster X-ray sources.
In conclusion only a smaller fraction of the known optical clusters will be detected
in X-rays while most of the ROSAT cluster sample should be different from the existing
catalogues. Similar results were obtained for the EMSS where among more than hundred
detected clusters of galaxies only 16 Abell clusters were found.
302

+78 0 50'

DEC (2000)

+78 0 35'

+78 0 20'

RA (2000)

Figure 6: X-ray image of the cluster A 2256 taken with the ROSAT PSPC. The X-ray
contours are in intervals of 2 PSPC counts per 8 X 8 arcsec pixels starting at 2 and the
image has been smoothed by a Gaussian with a width of 48 arcsec.

4 Observations of individual clusters of galaxies

During the calibration phase the Abell cluster 2256 was observed for 17 323 sec. It was
previously considered as a symmetric and well relaxed Coma type cluster even though a
slight ellipticity was observed with EINSTEIN (Fabricant et al. , 1984,1989). The ROSAT
image which is shown in Fig. 6 clearly shows two maxima in the X-ray surface brightness.
Further analysis which is described in detail by Briel et al. (1991) shows that the cluster
can be decomposed into two subcomponents.
In the 270 degree wide sector from position angle 310 to 220 the cluster appears azimu-
tally symmetric around the central cD galaxy which coincides with the eastern maximum.
A surface brightness profile ofthe form used by Jones and Forman (1984) can be well fitted
to this part of the cluster yielding a core radius of 480 ± 70 kpc and a scale height param-
eter (3 = 0.755 ± 0.025. If these parameters are used to calculate the surface brightness of
the symmetric component of the cluster in the remaining sector and if this result is sub-
tracted from the original image one is left with a very compact second cluster component
which is shown in Fig 7. This subcluster is clearly distorted indicating that the two cluster
components are interacting.
If the sample of 87 galaxy redshifts is divided into the same two sectors the mean
galaxy velocity in the undistorted sector is 17817 km S-1 while for the sector containing
the second component the mean velocity is 16977 km S-I. The probability that the two
velocity distributions belong to the same distribution function is less than 5 %. A fit
of two Gaussian distributions to the total velocity dispersion histrogram yields a velocity
303

I I I

+78 0 50
,
~ -

o~
DEC (2000)
©
0

,
+78 0 35 - 0
-

I I -

RA (2000)

Figure 7: X-ray image of the second component in A 2256. This contour image has been ob-
tained by subtracting a surface brightness distribution which was fitted to the undisturbed
part of A 2256 and which is symmetric around the central cD galaxy and the eastern max-
imum. The contour level intervals are 2 PSPC counts per 10 X 10 arcsec pixels starting at
2.
304

difference of about 2000 km s-1, where the smaller component has the lower velocity. For a
total mass of the system of'" 1015 Me this velocity corresponds to the case where the two
subcomponents started at rest and are today approaching each other at a distance of 1 to
2 Mpc. At this distance the observed X-ray halos already overlap which is consistent with
the above implications from the X-ray image that the distortion of the smaller component
is due to interaction of the intracluster medium of the two cluster components.
The brightest cluster of galaxies expected to be observed during the All Sky Survey is
the Perseus cluster with a count rate of '" 25 cts S-1 which is shown in Fig. 8. The mean
exposure time of the Perseus region in the survey was 712 sec. Also shown in the Figure
at the same scale is the image obtained previously with EINSTEIN (Branduardi-Raymont
et al. 1981) which clearly demonstrates the advantage of the "unlimited field of view" in
the survey mode. The X-ray halo was detected over a diameter larger than 2.5 Mpc. The
assymetric extension of the inner contourlines to the east which was already seen in the
EINSTEIN image is also visible here. More surprisingly a trace of the large prominent
chain of galaxies in Perseus is also seen for the first time in the X-ray image on an even
larger scale.
The second maximum of the X-ray surface brightness within the western tail in the
X-ray contours following the chain of galaxies, coincides with the galaxy IC 310. The
spectrum of the photons in this region is hard enough to be consistent with hot cluster or
galactic halo gas. The luminosity of this feature around IC 310 is about 2 - 3 . 1042 erg
S-1 in the ROSAT band. This is consistent with the luminosity of a group of galaxies( e.g.
Kriss et al. , 1983). Therefore this feature can be interpreted as a small group of galaxies
just falling onto the main cluster.
Both ROSAT observations reveal very interesting substructure in well known clusters of
galaxies that was not recognized before. Substructure in clusters has been studied mainly
by optical and spectroscopic observations (e.g Geller, 1984; Dressler and Shectman, 1988;
Zabludoff et al. , 1990). But due to the limited statistics which can be achieved with the
number of galaxies that can be studied spectroscopically, evolutionary effects as shown
here have never been discovered at such detail. Both examples imply that these clusters
are currently growing due to further infall of groups of galaxies or even mergers of clusters.
These very first ROSAT results on clusters of galaxies show the great potential of X-ray
observations for the study of the detailed structure and evolutionary state of galaxy clusters.

5 Prospects of ROSAT cluster observations for cosmological research

The analysis of the first ROSAT observations shows that the instrument is well suited
to conduct several interesting cosmologically relevant research tasks pertaining to both
the study of individual clusters as well as the investigation of the statistical properties of
clusters in the Universe. On a more general level and in the context of the topic of this
workshop the most important applications of these observations are related to the study of
the evolution of the large scale structure in the Universe and the cosmological parameters
that govern the cosmic dynamics. Four major research programs will be conducted that
are directed towards these applications: a careful determination of the present day X-ray
luminosity function of clusters, a correlation of various physical and observable properties
305

-...;..;'-'-'-- .....
. ~.~.; ..

. :.. ~ ' .

. .
.... '."'. '1--4 ".
..
'"."

"10 a;c~in:'

Figure 8: X-ray image of the Perseus cluster observed in the All Sky Survey. The contour
levels are logarithmically spaced. The X-ray map is superposed onto the POSS plate. The
main maximum coincides with NGC 1275 while the small western maximum coincides with
IC 310,
306

for a well studied sample of clusters, a study of cluster evolution towards high redshifts,
and the analysis of the spatial correlation of X-ray selected clusters of galaxies.
For a large statistically complete sample of the order of 1000 clusters of galaxies the
X-ray luminosity function should be determined which gives an approximate account of the
size and mass spectrum of clusters of galaxies at the present epoch. In addition the statistics
of more detailed cluster properties such as the spatial extent, the X-ray temperature, the
morphology and state of evolution should be derived from more detailed observations of
a smaller sample of clusters. These observations will then allow to statistically correlate
X-ray luminosities to cluster masses and other properties.
These data are important for the comparison with theoretical predictions for the evolu-
tion of clusters from primordial density fluctuation fields for different cosmological scenarios
(e.g. Frenk et al. , 1990). In particular the distribution of morphologies, as for example
the fraction of observed clusters that show a merger or other signs of ongoing evolution at
present, may provide the basis to discriminate between a closed or open Universe.
Another important clue to the understanding of large scale structure should come from
the investigation of the evolution of clusters of galaxies. The above estimates show that such
studies should be possible for small statistical samples (20 to 40 objects) out to redshifts of z
fV 0.5 where significant evolutionary effects are expected. A small number of clusters should
even be discovered out to distances of z = 1. The comparison of these data with models
will provide another serious constraint on the cosmological model. The magnitude of the
evolutionary effect for accessible redshifts depends very sensitively on the form of the power
spectrum of the primordial fluctuation field at wavelengths around 10 Mpc (Perrenod, 1980;
Kaiser, 1986; Evrard, 1989; B6hringer, 1991).
Finally the large sample of X-ray selected clusters should be used to study the cosmog-
raphy and the spatial correlation of the clusters. One of the most important first results
of the ROSAT survey is an indication that this cluster sample will be different from the
present optical samples and will probably also have quite different properties. The X-ray
data will in particular permit the construction of a large cluster sample with clear selection
criteria that are homogeneous over the sky except for the galactic plane. This is essential
for a precise spatial correlation analysis.
The final goal is a study of the spatial distribution in three dimensions for which redshifts
for the X-ray selected cluster sample are needed. Several optical follow-up programs are
devoted to this aim. A statistically complete all sky study can probably be done out to a
redshift of z = 0.1. Another program is aiming for a deeper sampling scale out to z = 0.25
for a few 1000 deg 2 in the southern sky and a very deep program is planed for the north
ecliptic pole region.
These results will provide information on the power spectrum of the primordial fluctu-
ation field on scales of up to 1000 Mpc. Considering the very dramatic implications the
recently obtained results of the galaxy correlation had on the validity of different cosmo-
logical models - as it was discussed at this workshop - one can expect further interesting
surprises if the above metioned projects are successful.

Acknowledgements. We like to thank H. T. MacGillivray, R.G. Cruddace, and D. Yentis for mak-
ing data from the COSMOS scans of the UK Schmidt Survey available for the optical identifications.
307

References

Abell, G.O., 1958. Astrophys. J.Suppl., 3, 211.


Abell, G.O., Corwin, H.G., Olowin, R.P., 1989. Astrophys. J.Suppl., 70, 1.
Bahcall, N.A. & Soneira, R.M., 1983. Astrophys. J., 270, 20.
Bahcall, N.A., 1988. Ann. Rev. Astr. Astrophys., 26, 361.
Bohringer, H., 1991. in Traces of Primordial Structure in the Universe, p. (in press), ed. Bohringer,
H. & Treumann, R.A., MPE Report.
Branduardi-Raymont, G., Fabricant, D., Feigelson, E., Gorenstein, P., Grindley, I., Soltan, A., &
Zamorani, G., 1981. Astrophys. J., 248, 55.
Briel, V.G., Henry, J.P., Schwarz, R.A., Bohringer, H., Ebeling, H., Edge, A.C., Hartner, G.D.,
Schindler, S., & Voges, W., 1991. Astr. Astrophys., , (submitted).
Cavaliere, A. & Colafrancesco, S., 1988. in Hot Thin Plasmas in Astrophysics, p. 315, ed.
Pallavicini, R., Kluwer Academic.
Dressler, A. & Shectman, S.A., 1988. Astron. J., 95, 985.
Edge, A.C., Stewart, G.C., Fabian, A.C., & Arnaud, K.A., 1990. Mon. Not. R. astr. Soc, 245,
559. .
Evrard, A.E., 1989. Astrophys. J., 341, L71.
Fabricant, D., Kybicki, G., & Gorenstein, P., 1984. Astrophys. J., 286, 186.
Fabricant, D., Kent, S., & Kurtz, M., 1989. Astrophys. J., 336, 77.
Frenk, C.S., White, S.D.M., Efstathiou, G., & Davis, M., 1990. Astrophys. J., 351, 10.
Geller, M.J., 1984. Comments Ap., 10,47.
Gioia, I.M., Henry, J.P., Maccacaro, T., Morris, S.L., Stocke, J.T., & Wolter, A., 1990a. Astrophys.
J., 356, L35.
Gioia, I.M., Maccacaro, T., Schild, R.E., Wolter, A., Stocke, J .T., Morris, S.L., & Henry, J.P.,
1990b. Astrophys. J.Suppl., 72, 567.
Heydon-Dubleton, N.H., Collins, C.A., & MacGillivray, H.T., 1989. Mon. Not. R. astr. Soc, 238,
379.
Huchra, J.P., Henry, J.P., Postman, M., & Geller, M.J., 1990. Astrophys. J., 365, 66.
Hughes, J.P., 1989. Astrophys. J., 337, 21.
Kaiser, N., 1986. Mon. Not. R. astr. Soc, 222, 323.
Kriss, G.A., Cioffi, D.S., & Canizares, C.R., 1983. Astrophys. J., 272, 439.
Lahav, 0., Edge, A.C., Fabian, A.C., & Putney, A., 1989. Mon. Not. R. astr. Soc, 238, 881.
Perrenod, S.C., 1980. Astrophys. J., 236, 373.
Saunders, W., Rowan-Robinson, M., Lawrence, A., Kaiser, N., Efstathiou, G., Ellis, R.S., & Frenk,
C.S., 1989. Nature, 349,32.
Sutherland, W.J., 1988. Mon. Not. R. astr. Soc, 234, 159.
Zabludoff, A.I., Huchra, J.P., & Geller, M.J., 1990. Astrophys. J.Suppl., 74, 1.
308

DISCUSSION:

Huchra: Which one of your X-ray clusters was the "point" source? Could it be an AGN
in a cluster?
Bohringer: I don't know the name of the cluster right now. We have looked into the UK
Schmidt Survey plates and found no bright star within the error box of the X-ray source
and no obvious known AGN. We have to wait for the more detailed data analysis to check
if a possible extended component of the X-ray source has been missed by the maximum
likelihood detection technique used in the standard data reduction.
Lahav: Given that the identification of ROSAT clusters depends on optical samples, is
the resulting catalogue going to be free of the systematic effects (projections etc.) of the
Abell catalogue?

Bohringer: This is indeed a critical point. Only a fraction of galaxy clusters detected in
the all-sky survey will be well characterised by their X-ray properties. The more interesting,
larger sample relies on optical identification. Such an X-ray selected sample will only be
unaffected by biasing effects in the optical data, when the optical identification for all X-
ray sources is essentially complete. Such a complete survey is already being conducted for
selected regions of the sky and its outcome will provide the basis for the evaluation of the
quality of an X-ray selected/optically identified galaxy cluster sample.
A DEEP ROSAT OBSERVATION AT HIGH GALACTIC LATITUDE

I. Georgantopoulos & T. Shanks


Department of Physics, University of Durham
G. Stewart & K. Pounds
X-ray Astronomy Group, Physicil Department, Leicester University
B. J. Boyle
Institute of Astronomy, Cambridge
R. Griffiths
Space Telescope Science Institute

ABSTRACT. A deep (30,000 sec) ROSAT observation of a high galactic latitude field reveals a
high surface density of sources (> 100 deg- 2 ). Spectroscopic follow up observations show that the
majority of these sources are QSOs with a surface density of 66 ± 14 deg- 2 , at least a factor of two
higher than that found by previous X-ray experiments. The discovered QSOs could contribute as
much as ~ 50% to the 1 keY extragalactic X-ray background.

1 Introduction

The most direct way to understand the origin of the X-ray background is through resolving
as many sources as possible. This was attempted by Giacconi et al. (1979) and Griffiths et
al. (1988) who found a surface density of '" 25 - 30 deg- 2 using Einstein High Sensitivity
Survey fields. Giacconi et al. (1979) identified only 4 sources out of 43 as QSOs while
Griffiths et al. (1988) identified most of their 16 sources in the Pavo field with faint QSOs.
Our deep (30,000 sec exposure) ROSAT survey covers a solid angle some four times
larger and has better sensitivity than the deepest Einstein survey. The fields selected for
the ROSAT observations form part of the Durham QSO survey (Boyle et al. 1988) which
selects QSO candidates on the basis of their ultraviolet excess and has provided to date
the largest, complete (B < 21m) QSO sample with redshifts up to 2.2. The high galactic
latitude of the Durham QSO survey (lbl > 500) minimizes the effects of X-ray absorption
and stellar contamination.

2 The ROSAT Observations

One of the five fields was observed in late July 1990 with the Position Sensitive Proportional
Counter (PSPC) (Briel et al. 1990). The PSPC operates in the 0.1-2 keY band and has an
on-axis angular resolution of 25" (FWHM). 93 sources were detected in the full PSPC 20
diameter field of view. However, we confine the discussion here to the central 20' radius field

309
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 309-313.
C 1991 Kluwer Academic Publishers.
..
310

...... ,. ' -,'


:'~ ,
- . : . . . .;;.., ' : .".- . , " \' • ~~
• , •• - ~1
~
. 4O. ,,'' , .
;.'
. .
_

.:.<'. '.-=-:"
.:.... ..... ...
.. I • -, .. .... "' ....

_..........'.
.'

.. ~ •.•:.::.:..'.:.~~-'
.. • : .. 'J " " ,,-
.. ..
'. ,':''':-:-' . .. , . '~ '.':.~' • • • '
." II),
..
.
.,. -.
;. ~
:~:. :\~ )'~r·:. ';:':-M",.,';~::.~~ .
• I •• • •• - ; ' . : .' ' .
I ' •
........ .
"
-.

I'" . . ,,'
.. ~.....
..
-"~ " .
! . '
,.' " -" .,"- " . " · _. -,:
•- . '•I~ -~~
~.-' ~ -• _.', " :-'"
.' ." . ••••.( ..> '. : •• •• . ..,\-~
I"'~' V.··· ,1 ~··:~·· "t·.~.J..··!·
• • I ' -
•• J .... '.
L. ::
.. . ... " I; ..
r. ·.. . ..,, '. · ···
··-.,
_ . , . • •" ( ' ... ". ' • • _ •••• 0< '.:'
• , :•••• •.
'.' " . ... •.•,'_.' ,..." .. .... '. ...
.. ' '\ .._...
+ • •

. .' '.. ..,... . ;.. ':...:_.,:. . ;.,.:.~...


, .,::~='-:
. "~~. ~~~,
' ."..,\~
, .It
.. . ..
" .- ' :'. ."
::. J:';; '• N
.; ' ' ·r··· ·• . _. ... .. ...,
_ • . , . ,' _r$.' •
. .
, - .-li··
. :~:i:
'*"~.·~5:·:(
,.
...
. . .. . . . . . . .0. " -

-.
...-~

." ...
.••. .. _ . ,... ;,
. :' .-'.-." :i;.,
• '..•••
,_ •' ..' '. '....
.•: • ' i/ .
.......:. '" ,;,; ....; •..' ;..~ :r " ' .. ,-
' . ..'
....: .' , .. '" . ' .'
: , ... ,'
.,...,
. .."r ". ' "'.'~
... ::. ., '.'
.' • .
.... I .
• ...,. '.'
1"

,..~.'
. ' ,'':
.~ ~
..- " • ."'';' • • •

. ". ,.-.1 .-". .... .".: '..... .*-,,-,. . . . :


.' :. "'..... ... •. ,. '. ... . " . '1
.. .'.
.. . ....
•• • ..•..•
y ,• •
: "..

''',Jr~ ., ....~.
'
. ,.
.......
:. , '!t' . ...
·~ .:. _. '" .' ... ,.I,:. . :. .." =•..
~. , '
." • ..• . ...
• . . .r. .•.to.
.
" '
• • •• •
..
". , ' •

... ':.~ '"l{ ~.. ..


.. " I, • • .. . .' . ... ••. ' - '. .... - '.

': . ;.:;;..'.. :...., ', '. ' .,':Ii .0:..... ...:.:..


..

.W., ··· .: ~"" >' (10 ' - ' : . '


'. :': }:
.'
...., 'Ii '1'::' • " "":,, - ". ~~,,,' . ;: ..~ .•:tf..~< ..
. . .... ....
. " .' . .-..
.' . ., . .': ; I..~ r .'
. . .' ..A .~ '. ' r····
::'. ".:' : :.
".~.
..........
. .. . . :; .: : . ... : '. ....:.
- :~
' .' .. .. ... ... ' :.
. .: ' .. . .
.. ' •.. ...-
'~,.;--' ":.:"'. , ...,.. ••.~.:"~.... .
' ' ."

,..i.....
.-. ~ .' :,.~.' ,
.
, j~

,
:"t" : '-
• 0- ' t
..
• • .' ...

Fig. 1 The central 20' radius PSPC image with 39 sources detected

o
o .5 1 1.5 2 2.5
Z
Fig, 2 The X-ray .elected qso red.hin distribution
311

where the detector's sensitivity is greatest. The image of this central area, intyegrated over
the full energy range is shown in Figure 1. 39 sources were detected here, yielding a surface
density of 117 ± 19 deg- 2 compared with 36 ± 9 deg- 2 for the Einstein High Sensitivity
survey (Griffiths et al. 1988). While the brightest source detected was a star with B < 12m,
the majority of bright sources were optically selected QSOs from the Durham survey. A
high optically selected QSO detection rate (12 out of 19) was obtained. Previous Einstein
observations of optically selected QSOs proved that X-ray emission is a common property
of the rare, bright, low redshift QSOs (e.g. Tananbaum et al. 1986). Our high rate of
detection extends this to show that moderate to high redshifts, faint (B > 20m ) QSOs are
also typically X-ray emitters.

3 Spectroscopic Follow Up Observations

Optical spectroscopy was necessary to reveal the identity of the X-ray sources not catalogued
in the Durham QSO survey. Optical counterparts were selected from the objects found on
blue UK Schmidt telescope plates within a 15" radius error box, derived from the offsets
IV

of the optically selected X-ray sources. In most cases only one optical counterpart was
found within the error box. Observations took place between 14 and 16 of November 1990
at the 3.9-m Anglo Australian telescope. The use of AUTO FIB allowed for multi object
fibre spectroscopy and enabled us to obtain simultaneously spectra for all 39 sources in the
central field of view with a 6 hour exposure.
fV

Preliminary results show that in addition to the 12 optically selected QSOs detected in
the X-ray, 10 new QSOs were discovered, most of these with magnitudes fainter than the
optical QSO survey's limit. The surface density of X-ray QSOs is high (66 ± 14deg- 2 ) as a
result of the low flux limit 2 X 10- 14 erg cm- 2 sec-I. The redshift distribution of the 22
fV

X-ray selected QSOs is presented in Figure 2. It is evident that the present ROSAT survey
picks up distant luminous QSOs (in general Lz > 1044 erg sec-I, Ho = 50, qo = 0.5) rather
than nearby, low luminosity AGN.
Only three X-ray sources have been positively identified with stars. Two are very bright
(B< 12m) F stars and the third is probably a sdO+G binary.
It is very interesting to examine if any new class of sources is present among the faintest
in our survey. Among the faint sources, of particular interest are 2 objects classified by the
plate measuring machine COSMOS as galaxies: One has a rather poor spectrum showing no
clear redshift while the other object could be classified as a galaxy with a possible redshift
of 0.18, on the basis of its Call, H and K absorption. Neither of these possible galaxies
seems to present emission lines, indicating that these are not starburst galaxies.
The remaining 12 objects are still unidentified. Five of these are very faint giving
no clear spectroscopic identification while the remaining have no optical counterpart on
the Schmidt plate, i.e. their counterpart is fainter than B IV 22 m. The following table
summarizes the spectroscopic identification results.
312

QSO 22
stars 3
galaxies? 2
unidentified 12

Table: Summary of Spectroscopic Identities

4 Contribution of Discrete Sources to the X-ray Background

The contribution of discrete sources detected can be estimated in a number of ways. The
most straightforward is to sum the flux: from the sources and divide by the total background
counts within the central 20 arcmin. This gives a lower limit to the point source contribution
of'" 20% in the 0.1-2 keY band with the 22 detected QSO contributing around half.
The PSPC energy resolution allows to make a similar estimate at a monochromatic
energy of 1 keY. A lower limit to the QSO contribution at 1 keY is '" 30 ± 5%. An
increasing fraction of the observed PSPC background is expected to be galactic as we move
to lower energies (McCammon and Sanders 1990). Extrapolating the 3-40 keY spectrum
observed by Marshall et al. (1980) down to 1 keY suggests that only half of the ROSAT
flux: at 1 keY is extragalactic, resulting in a QSO contribution of '" 50% to the extragalactic
component of the X-ray background.

Acknowledgements. We wish to thank the director of the Anglo-Australian Observatory Dr.


Russell Canon and the support astronomer D. Hatzimitriou for their help during our AAT observing
runs. We also want to thank the Max Planck Institut fUr Extratterestrische Physik for their support.

References

Griffiths, R. E., Tuohy, I. R. and Brissenden, Ward, M., Murray, S. S., and Burg, R., 1988. in Post
Recombination Universe, p. 91, ed. Kaiser and Lasenby, Kluwer.
Giacconi, R., et al., 1979. Ap.J. Letters, 234, L1.
Boyle, B. J., Shanks, T. and Peterson, B. A., 1988. Mon. Not. R. astr. Soc, 235, 935.
Marshall, F. E., Boldt, E. A., Holt, S. S., Miller, R. B., Mushotzky, R. F., 1980. Ap.J, 235, 4.
McCammon, D. and Sanders, W. T., 1990. Ann. Rev. Astron. Astroph., 28, 657.
Briel et al., 1990. Proc. SPIE, 1344, in press.
313

DISCUSSION:

Ellis: Considering Hans Bohringer's talk, the absence of any clusters in your deep X-ray
image is a bit surprising. Did you purposely exclude them from your spectroscopic target
list? Would you care to comment on the proportion of clusters in deep X-ray images from
ROSAT?
Georgantopoulus: No extended sources were picked up in our deep image. No rich
clusters were identified as counterparts to the X-ray sources.
Cristiani: What can you say on the basis of the ROSAT data about the completeness of
the original optical Survey?
Georgantopoulus: The newly discovered X-ray QSOs have in general magnitudes fainter
than those of the original optical Survey. There is only one definite X-ray selected QSO
(B=19.7) not picked up by the optical survey because of its colour (U-B = -0.12); however,
it has redshift Z = 0.564 at which the MgII emission line shifts to the B band and therefore
incompleteness is the greatest. The incompleteness implied (lout of 20) for ultraviolet
excess surveys is well in agreement with previous estimates.
LARGE SCALE STRUCTURE AND INFLATION

JOHN P. HUCHRA
Harvard-Smithsonian Center for Astrophysics
60 Garden Street
Cambridge, MA 02138
United States

ABSTRACT. Recent observations of the large scale distribution of galaxies and


clusters of galaxies are contrasted with predictions of models for the formation of
such structures (primarily n-body simulations) based on the inflationary paradigm
of n = unity. The combined set of observations are not easily matched by any
single model and many of the observations favor low n universes. While the failure
to match the observations of galaxies and clusters is not conclusive, it strongly
suggests that the simplest, n = 1, A = 0 models are incorrect.
1. INTRODUCTION
The inflationary paradigm developed by Guth (1981) and others offers observers an
unprecedented chance to test one of the most simple and basic modifications of Big
Bang cosmology. Cosmic Inflation is extremely attractive not only because it offers
theorists solutions to some of cosmology's conundrums - the flatness and horizon
problems in particular - and provides them with justification for a "heavier," dark
matter dominated universe in which the formation of galaxies without violating
constraints imposed by the uniformity of the microwave background is much easier,
but also because it makes a strong prediction that the Universe is fiat, n = 1.0000 .....
On top of this, the last decade has seen significant advances in both our ability
to develop computer simulations of the formation and growth of structure in the
Universe under a variety of assumptions about the input physics and initial condi-
tions and our ability to observe structures and flows in the nearby universe through
the combination of redshift surveys and redshift independent distance estimates to
galaxies and galaxy clusters.
In the following sections, I would like to describe a series of observations of
large-scale structure in the Universe, primarily those from our group at CfA which
I am most familiar with, and then contrast these and other data with the results of
several simulations based primarily on the assumption that n = 1. I will concen-
trate primarily on what has been learned through redshift surveys and leave direct
determinations of n from galaxy motions to other speakers at this conference. The

315
T. Shanks et al. (ecis.). Observational Tests oj Cosmological Inflation. 315-326.
© 1991 All Rights Reserved.
316

outline of this presentation is similar to many that have been given before (egg.
White 1986) and to the authors of those previous reviews lowe a great debt.
2. GALAXY REDSHIFT SURVEYS
For those of us in this business, the last decade and a half has been one of
great productivity. The total number of galaxies with known redshifts has gone
from less than 1,000 in 1975 to nearly 40,000 in 1991. Large and comprehensive
surveys have been done for example by Haynes and Giovanelli (1988), Kirshner,
Oemler, Schechter and Shectman (1987), Dressler (1991), Da Costa et aI. (1988)
and Strauss et aI. (1990).

e.1 The CIA Survey


Perhaps the largest of the efforts has been the Center for Astrophysics (CfA)
survey and its extensions (Davis et aI. 1982; Huchra et aI. 1983; de Lapparent
Geller and Huchra 1986; Geller and Huchra 1989). The CfA survey is essentially a
ten year project to map the space distribution of galaxies in the northern hemisphere
using a well defined galaxy sample complete to an apparent brightness (magnitude)
limit. The sample is selected from the Zwicky catalpg and has moderately large
sky coverage (in excess of 3 steradians). The first survey contained 2400 galaxies
and was completed in 1982. The second survey containes over 15,000 galaxies to a
limiting blue magnitude of 15.5, was started in 1985 and should be completed by
1994.
Galaxy redshifts surveys can be done via several different strategies (Geller
1987; Geller and Huchra 1989). The strategy we have adopted is similar to that
used in geographic satellite mapping, that of completely covering strips of the area
to be mapped, in our case 6° wide strips of the sky, until the whole area of interest
is completed. We currently have 5 complete strips in the northern galactic cap, 4 of
which are contiguous, and 5 complete and contiguous strips in the southern galactic
cap. Cone diagrams of these regions (including the 62% complete intervening strip
in the north) are shown in Figure 1. I also want to note that with R. Marzke, M.
Geller and I, to the best of our ability, are extending the survey through the galactic
plane with a combination of optically and IRAS identified galaxies.

e.e The Galaxy Distribution


The simplest and most striking "observations" of the galaxy distribution are the
visual impressions of the maps. In Figure 1, despite the large angles over which the
declination axis is projected (36° and 30°, respectively), very large and apparently
coherent structures are still seen. In (a) the "Great Wall," a moniker delivered
by Aveshai Dekel, extends across the wedge from Sh to 17 h between velocities of
approximately 6,000 and 10,000 km s-l. This structure contains many of the well
known smaller structures like the Coma supercluster near the center of the map
and the Abell 2197/99 and Hercules superclusters near 16 h • The Great Wall is
317

12'

r,
1.0 < m $: 155
8.5 S d < 44

5472 entrif'!i (a)

.'
10000

1.0 < m S 155

6.0 ~ 6 <: 36 O· cz (km/s)

2252 entrJes
(b)

Figure 1. The distribution of galaxies in two nearly completed zones of the CfA
Redshift Survey out to a limit of 12,000 km s-l. (a) The 5,472 galaxies in the
north galactic cap between 8.5° and 44.5°, and (b) The 2252 galaxies in the
south galactic cap between 6° and 36°. The "Great Wall" extends across (a)
varying between 6,000 and 10,000 km s-l. The Perseus-Pisces chain can be seen
in (b) between 23 h and 3 h ; galactic extinction strongly diminishes the contrast
in this map before 22h and after 3 h .

approximately 15Ox50 h- 1 Mpc in size (where h is the Hubble Constant in units


of 100 km s-1 Mpc- 1 ). In (b) is seen a somewhat smaller structure, the Perseus-
Pisces supercluster, that may, in fact, connect to the Wall through the galactic
plane. A structure similar in mass although somewhat smaller in dimensions, the
"Great Attractor," has been seen in the southern hemisphere (Lynden-Bell et al.
1988). It thus appears that very large structures (> 50 Mpc h- 1 ) are common.
To me, perhaps the most striking feature of the survey is the "sharpness," or
extreme thinness of the structures. This can be seen in the thin slices through this
region shown in Figure 2. These three 3° thick slices show a fine webwork of struc-
ture where the thickness of the surfaces on which most of the galaxies are distributed
is less than 10% of the diameters of the low density regions they surround. This
is strongly suggestive of a low density Universe. The contrast between the surfaces
318

and the voids is greater than 20, but the voids are not completely empty of galaxies.
More detailed analyses of the CfA survey, the KOSS surveys (e.g. Kirshner et al.
1987), the IRAS survey (Dey, Strauss and Huchra 1990) and detailed examinations
of voids (Eder et al. 1989) show that voids are only underdense by factors of 5-10
w.r.t. the mean.

,,' II' ,.' ... II'

,,' '0" ,,' '0"


,.' e' ,.' "
.' "
,0000 ,....
(a)
10 < m' 155 LO<m~ 155
2B5'''< 295" c::r (km/s) 2'95'''< 3lS" c. (/rm/s)
&44 "nfrj.. s
,.' ,,'
,,' '0"
,.' .. ~ .
"
.'

355"< J8S' a (kmhl


386 Inlrjf!S'

Figure 2. Three "microtomed" slices through the northern hemisphere, 3° wide


at constant declinations. These slices illustrate the sharpness of the features we
see in redshift space. The velocity dispersion perpendicular to the surfaces is
very small.
Another important qualitative question asked about the galaxy distribution is
its "topology" (Gott et al. 1989). "Bubblelike," "meatball," "spongelike" and other
terms have been used to describe the distribution. Perhaps all can be said to apply
to some portion of the maps. Figure 3 shows cuts in declination through the upper
4 slices in our northern hemisphere survey (Figure 1a). These declination slices
cut through the Coma cluster, the Coma-Virgo void (in front of the Coma Cluster)
and the large void to the east of Coma at ",5,000 km s-I, perhaps best call the
"Near Bootes Void" in contrast to the one found by KOSS. Figure 3b shows that
the Coma-Virgo void (centered near 4500 km s-1) is closed off at high declination.
Figure 3c shows that the Near Bootes void is also closed at high declination, very
suggestive of a bubble-like or closed cell topology. Galaxies generally lie on surfaces,
some of which are topologically closed; galaxy clusters lie at the interstices.
The reader should also note that despite many claims to the contrary, there is
very little evidence that, outside the well known morphology density relation
319

decljnation declination

40· 30·

12.0 is a < 13.0 =t a <

40· 30·

14.0' a < 15.0 h 15.0' a < 160 h

Figure 3. Four cuts in declination in the northern hemisphere; all cover the declina-
tion range 20.5°-44.5°: (a) 12h_13 h , (b) 13 h-14h, (c) 14 h-15\ and (d) 15 h-16 h .
320

(Dressler 1980), there is no good evidence that galaxies of either different lu-
minosity or different morphology trace the structures differently (Thuan, Gott and
Schneider 1989; Huchra et al. 1990). Dwarf galaxies do not fill in the voids.

t.9 Distribution Statistics


In addition to the qualitative aspects described above, a large number of quanti-
tative statistics have been measured for the galaxy distribution (see, for example de
Lapparent, Geller and Huchra 1988,1989, and 1990). The most referenced of these
are the luminosity function and the correlation function. The luminosity function
describes the number density of galaxies as a function of luminosity and presum-
ably mass, and is usually described in terms of the Schechter function (1976), a
combination of a power-law with an exponential cutoff. The correlation function is
usually described as a power-law. The best fit parameters for these functions for
the CfA survey are:
M* = -19.2 ± 0.1; a -1.1 ± 0.1
and
So = 7.5 Mpc
Other statistics like counts-in-cells show that 75-80% of the survey volume is devoid
of bright galaxies and that galaxies are generally found in a network of surfaces.
Measures like the mean pairwise velocity dispersion and measures of the velocity
dispersions of galaxy groups (Ramella, Geller and Huchra 1989) and galaxy clusters
(Zabludoff, Huchra and Geller 1990) indicate that the Universe is "quiet" on small
scales with velocity dispersions of only a few hundred km S-1 outside of rich clusters
and usually less than 1200 km s-1 even in the cores of rich clusters. These are
strongly suggestive of a low density universe even accounting for reasonable bias
factors in galaxy formation (Kaiser 1984). Any successful model will have to account
for these small velocities.
3. GALAXY CLUSTER SURVEYS
We can use galaxy clusters to probe (sparsely) even larger scales. In the last
decade there have been three major surveys of the 3-D distribution of galaxy clus-
ters. The first was the 104 cluster sample of Hoessel, Gunn and Thuan (1980)
that was analysed by Bahcall and Soniera (1983). They found the amazing result
that the amplitude of the cluster-cluster correlation function was very large; the
correlation length was ~ 25 h- 1 Mpc.
Since then, we have completed two major new cluster redshift surveys. The
Nearby Abell Cluster survey consists of all 351 clusters with mlO 2: 16.5 and above
Dec -27.5° (Postman, Huchra and Geller 1991). There are also efforts underway
to complete this survey over the whole sky (Olowin et al. private communication).
Figure 4 shows examples of the distribution of clusters in two wedges between
declination 0°-20° and 20°-40°. The Deep Cluster survey consists of all 145 clusters
321

with R ~ 0, and between lOh < a < 15 h and 58 0 < 0 < 75 0 , (Huchra et al.
1990), an area chosen because of its proximity to the North Ecliptic Pole .

.. I •
.. I.

• 0

... 0

Q)
6)

0 •
<00

...
(b)
(a)

Figure 4. The distribution of nearby Abell clusters out to 35,000 km s-1 between
(a) 0 0 _20 0 declination, and (b) 200 _40 0 declination. Filled circles are richness
class 1 or greater, open circles are richness class o.

9.1 The Cluster Distribution


Despite my early misgivings about the Bahcall and Soniera (1983) result, pri-
marily because of the known problems with the Abell catalog, our nearly indepen-
dent surveys confirm their results. While the cluster angular correlation function
shows a large scatter in amplitude for samples of similar size, the spatial correlation
function appears "robust" and strong on very large scales. In the Deep survey:

for R > 0, Ro 20 Mpc

for R > 1, Ro 21 Mpc


In the Nearby survey:

for R > 0, Ro 21 Mpc

for R > 1, Ro 25 Mpc


Figure 5 shows the cluster-cluster correlation function for the Nearby Cluster survey,
both including and excluding the Corona Borealis supercluster. Although we do not
find evidence for significant cluster-pairwise velocities, there is very strong evidence
for "structure" in the cluster distribution on scales greater than 50 h- 1 Mpc.
322

4. MODELS

Observations of large-scale structure are compared to the predictions of the In-


flationary Paradigm by comparison to simulations based on the two primary models
for the growth of structures. The first is the Gravitational Instability (Top down)
model primarily championed by Peebles and co-workers where the evolution of the
Universe is dominated by gravity produced by Baryons and WIMPS (cold dark
matter). Baryons make up only ",10% of the mass density and cold dark matter
the rest to get n to unity. The second is the Adiabatic or Pancake model (Bottom
Up) promulgated by Zeldovich and co-workers where viscous forces dominate at
early epochs such as those caused by hot dark matter like neutrinos, and gravity
wins in the end. There are other models based on a variety of physical forces such
as the explosion model of Ostriker and Cowie (1981) and Ikeuchi (1981) where the
dominant force in the early universe is the energy input from supermassive stars or
from large negative density fluctuations or cosmic strings.

10

flit
• Cor lor ID
o Cor Bar OU

~
"f
, I
g.....

.1 I

.01
I
I
3 10 30
R ("pc)

Figure 5. The cluster-cluster correlation function for nearby rich clusters (from
Postman, Geller and Huchra 1991).

Major sets of n-body simulations based on the different assumptions outlined


above have been made by many different groups, primarily White, Davis, Efs-
tathiou, and Frenk (1987), Blumenthal, Dekel and Primack (1988), and Melott and
collaborators (e.g. Melott and Scherrer 1988). The n-body models are run until
the model galaxy-galaxy correlation function approximates the observed function.
Biasing is applied (Kaiser 1984) to decide which particles or collections of particles
count as galaxies.
323

4.1 Comparison with the Observations

The zeroth order test of a model is then whether it can produce the observed
shape and amplitude of the galaxy correlation function. Models which do not
pass this first filter are rarely published! Next one compares the qualitative and
quantitative aspects of the model distribution to the observations. These include
general impressions about the "topology," the degree of clumpiness, the strength of
the clustering, the appearance of filamentary structures, the "sizes" of structures
like voids and walls, and the cluster-cluster correlation function. These comparisons
require putting the models into the observed radial velocity space - much easier
than going in the opposite direction.
Finally, the comparison is extended to the velocity field, including the pairwise
galaxy dispersion, cluster internal velocity dispersions (Frenk et al. 1990), and the
measurements of large scale streaming velocities or flows. This last comparison
is usually confined to matching the probability of observing a velocity as large as
our velocity w.r.t. the microwave background or w.r.t. the Great Attractor in the
simulation under consideration - since we really only have one measurement of a
flow velocity at the moment.
The current situation is not promising. The simplest 0 = 1 hot dark matter
models fail because (1) they produce too much random motion on small scales,
and (2) they produce clustering that appears too dense. The simplest 0 = 1
Gravitational Instability models (usually dominated by cold dark matter) are in
trouble because (1) they, too, produce too much random motion on small scales,
(2) they cannot produce largest structures, and (3) they do not generally produce
large coherent flows.
It is possible to adjust the parameters of these models to match a few of the
constraints, but then others are violated. It is difficult to produce large structures in
any of the simple models without strongly adjusting the input fluctuation spectrum;
in most of these models there just isn't enough time to generate structures on such
large scales. It is generally difficult to match the quiescence of the velocity field on
small scales with high 0 models.
5. WHAT DOES THIS MEAN FOR INFLATION?
The overall situation for cosmology is quite unsettled. New simulations by Park
(1990) and Weinberg and Gunn (1990) have confirmed the results of earlier work
and have shown that many of the qualitative features of the galaxy distribution can
be mimicked in cold dark matter simulations, especially those seen on small scales.
However it is still not possible for any of the simulations to produce the largest
structures seen. Low 0 models do a somewhat better job (remember that a low
density Universe is 50% older than one with 0=1 ) on both very small and very
large scales (Blumenthal et al. 1988).
The only significant set of observations that weakly support a high value of 0 are
324

the attempts to match the observed peculiar velocity field with the galaxy density
field (e.g. Strauss and Davis, 1988; Yahil 1988). However, before an absolute
measure of n is derived from such measurements we need a considerably better
understanding both of the biasing of mAS galaxies w.r. t. the optical galaxy density
and the real mass distribution and of the actual velocity field.
Taken at face value, the majority of the observations of large-scale structure
suggest a low n Universe contrary to the predictions of the Inflationary Paradigm.
The case is a long way from closed, however, because no satisfactory model of any
n has been found to fit all the observational constraints. Several large surveys
of the galaxy and cluster distributions and galaxy distances are now underway or
are just beginning. Telescopes like the Princeton-Chicago survey telescope and
the new, wide-field 8-meter class telescopes combined with surveys like 2-mu sky-
survey currently proposed by S. Klienmann and her collaborators will open up new
windows and clear away much of the confusion at low galactic latitudes and caused
by surveys sensitive to only special types of galaxies (e.g. IRAS). Over the next
decade we hope to have considerably deeper, whole sky observations of the galaxy
distribution (to z'" 0.1), the nearby velocity field (to z '" 0.4), and the clustering
properties to redshifts between z = 0.2 and 1.0. Only then might we have· the
answer.

The author would like to thank his many collaborators and colleagues, especially
M. Geller and M. Davis, and many fellow observational cosmologists who's works
have contributed to his world view. Special thanks go to all the people who make
things happen like J. Geary, S. Tokarz, W. Wyatt, L. Feldman and all the telescope
operators of the world. The original research presented here has been supported
by NASA grant NAGW-201 and the Smithsonian Institution and is based partly
on observations obtained with the Multiple Mirror Telescope, a joint facility of the
Smithsonian Institution and the University of Arizona. Typing services provided
by Fang, Inc.

References
Bahcall, N. and Soniera, R. 1983, Ap. J. 270,20.
Blumenthal, G., Dekel, A. and Primack, J. 1988 Ap. J. 326, 539.
da Costa, L.N., Pellegrini, P.S., Sargent, W.L.W., Tonry, J. Davis, M. and Meiksin,
A. 1988, Ap.J., 327,544.
Davis, M., Huchra, J., Latham, D. and Tonry, J. 1982, Ap. J. 253,423.
de Lapparent, V., Geller, M. and Huchra, J. 1986 Ap. J. Lett. 302, L1.
de Lapparent, V., Geller, M. and Huchra, J. 1988 Ap. J. 332,44.
de Lapparent, V., Geller, M. and Huchra, J. 1989 Ap. J. 343,1.
de Lapparent, V., Geller, M. and Huchra, J. 1990, Ap. J. 369,273.
Dey, A., Strauss, M. and Huchra, J. 1990, A. J. 99,463.
Dressler, A. 1980, Ap. J. 236,351.
325

Dressler, A. 1991, Ap. J. Suppl. '15,241.


Eder, J., Schombert, J., Dekel, A., and Oemler, G. 1989, Ap. J. 340,29.
Frenk, C., White, S., Efstathiou, G. and Davis, M. 1990, Ap. J. 351, 10.
Geller, M. 1987, in Large Scale Structures in the Universe, 17th Saas-Fee Lectures,
L. Martinet and M Mayor, eds. (Geneva: Geneva Observatory).
Geller, M. and Huchra, J. 1989, Science, 246,897.
Gott, J. R. lIT, et al. 1989, Ap. J. 340,625.
Guth, A. 1981, Phys. Rev. D23,347.
Haynes, M.P. and Giovanelli, R. 1988, in Large-Scale Motions in the Universe, V.
C. Rubin and G.V. Coyne, S.J., eds. (Princeton University Press: Princeton),
p. 31.
Haynes, M.P. and Giovanelli, R. 1986, Ap.J. Lett., 306, L55.
Huchra, J., Davis, M., Latham, D. and Tonry, J. 1983, Ap. J. Suppl. 52,89.
Huchra, J., Geller, M., de Lapparent, V. and Corwin, H. 1990, Ap. J. Suppl., '12,
433.
Huchra, J., Henry, J. P., Postman, M. and Geller, M. 1990, Ap. J. 365,66.
Ikeuchi, S. 1981, P.A.S.J. 33, 211.
Kaiser, N. 1984, Ap. J. Lett. 284, L9.
Kirshner, R.P., Oemler, A., Schechter, P. and Shectman, S.A. 1987, Ap.J., 314,
493.
Lynden-Bell, D., Faber, S.M., Burstein, D., Davies, R.L., Dressler, A., Terlevich,
R. and Wegner, G. 1988, Ap.J., 326, 50.
Melott, A. and Scherrer, R. 1988, Ap. J. 331, 38.
Ostriker, J. and Cowie, L. 1981, Ap. J. Lett. 243, L127.
Park, C. 1990, M.N.R.A.S. 242, 59P.
Postman, M., Huchra, J. and Geller, M. 1991, Ap. J. in press.
Ramella, M., Geller, M. and Huchra, J. 1989, Ap. J. 344,57.
Schechter, P. 1976, Ap. J. 203,297.
Thuan, T. X., Gott, J. R. ITI, and Schneider, S. 1987, Ap. J. Lett. 315, L93.
Strauss, M., Davis, M., Yahil, A. and Huchra, J. 1990, Ap. J. 361,49.
Strauss, M. and Davis, M., 1988, in Large-Scale Motions in the Universe, V. C.
Rubin and G.V. Coyne, S.J., eds. (Princeton University Press: Princeton), p.
256.
Weinberg, D. and Gunn, J.E., 1990, Ap. J. Lett. 352, L25.
White, S. D. M. 1986, in Innerspace / Outerspace, D. Lindley and E. Kolb, eds.
(University of Chicago Press) p. 228.
White, S. D. M. 1988, Minnesota Lecture Series: Clusters of Galaxies and large-
Scale Structure, J. Dickey, ed. (Provo: Astronomical Society of the Pacific) p.
197.
White, S. D. M., Frenk, C.S., Davis, M. and Efstathiou, G. 1987, Ap.J., 313, 505.
Yahil, A. 1988, in Large-Scale Motions in the Universe, V. C. Rubin and G.V.
Coyne, S.J., eds. (Princeton University Press: Princeton), p. 219.
Zabludoff, A., Huchra, J. and Geller, M. 1990, Ap. J. Suppl. '14,1.
326

DISCUSSION:

Moore: Are you not suspicious of using rich cluster catalogues to constrain models of
galaxy formation when, for example, the 2-point correlation function disagrees with the
predictions of the n = 1 CDM models but a topological analysis of the same data (Gott et
al, 1989) agrees well with the model?

Huchra: No. First of all the topological results come from a small sample with a few
redshifts. Also the above tests check different things. If one set of "observations", eg the
~cc, agrees with the theory, then the theory is presumed incorrect (unless you can show that
the data is wrong!). The topological "shape" of the distribution could be correct, but the
amplitude all wrong.
Melott: I see filaments when I look at your videotape, but you see sheets. Do you plan
any quantitative tests of thes visual impressions?
Huchra: Tough question. Yes, after we get the six contiguous strips in the north and
the six in the south completed. There are papers by M. Vogeley and V. de Lapparent and
Margaret and I in the works.
Shanks: Why do the Weinberg-Gunn CDM simulations look more similar to the obser-
vations than the Frenk-White CDM simulations?
White: This is in fact because they chose their geometry to match that used to present
the "Great Wall" (a 20° wedge) rather than that of the earlier, single slice surveys. I have
made plots from our own simulations with the appropriate geometry and they show heavier
and larger structures than our earlier work (they look like those published by C. Park). The
structure in Weinberg-Gunn's plots tend to exaggerate the thinness and the continuity of
filaments as a result of the approximate method (the "adhesion approximation") they use
to follow the dynamics. Fully consistent dynamical simulations show less coherent structure
than their models.
THE STRUCTURE OF THE UNIVERSE ON LARGE SCALES

A. G. Doroshkevich.
Institute of Applied Mathematics Academy of SciencelJ USSR
Moscow, USSR

ABSTRACT. A possible explanation of the results from a quantitative analysis of the deep galactic
surveys is suggested. The theoretical model taking into account the influence of the potential on
the galaxy distribution is given. This model allows explaining the galaxy distribution on superlarge
scales and the formation of large voids. It is shown that high density of the structure elements
between voids contradicts to standard spectra of the cold and hot dark matter models and to simple
power spectra. This effect can be explained if we assume a special shape of the initial spectrum in
the shortwave region.

1 Introduction

The large scale structure of the Universe ( LSS ) was discovered in the late 1970s but no
adequate methods have yet been developed to analyse and describe it. A very powerful
method of the galaxy correlation function is useful for moderate distances ( smaller 15 +
20Mpc). Observational data for the spatial distribution of galaxy clusters, for the Great
Attractor, Great Wall and Great Voids have a poor statistics and at present do not allow one
to obtain reliable information about LSS. Perhaps, an analysis of the deep surveys of galaxies
and the absorption spectra of quasars may prove more useful for such an investigation. In
this case we would have both good observational statistics and a possibility for quantitative
estimations of the LSS parameters.
Let us draw attention to the evidences of the quasiperiodic structure of the Universe on
the supedarge scale with the period T = 100h- i Mpe, h = H/lOOkm/s/Mpe is Hubble's
constant (Broadhurst et al.,1990) and to the evidences of two characteristic scales for the
LSS element distribution along a random stright line (Buryak et al. ,1991). In the last case
the estimations for the mean distance between separate elements of LSS in the regions with
high density I ~ 7h- I Mpe and for the mean size of voids L ~ 80h- I Mpe were obtained.
It is important that the correlation radius of the galaxy distribution r 9 is near the value I.
It is also interesting that an analysis of the spatial distribution of absorption doublets CIV
in the quasar spectra shows that the mean distance between doublets is L ~ 80h- I Mpe
too and remains approximately constant up to the redshift z ~ 3 .. This suggests that main
elements of LSS were formed at the redshift z > 3. and later evolved very slowly.

327
T. Shanks et al. (etis.), Observational Tests a/Cosmological Inflation, 327-330.
@ 1991 Kluwer Academic Publishers.
328

Below we will discuss a possible theoretical description of the supedarge scale structure
and numerical estimates of the spatial distribution of separate elements of LSS. For this
purpose it is necessary to take into account an influence of the potential f1 (gravitational
or velocity) on the formation of LSS. This approach allows one to realize quantitatively the
old idea about the influence of longwave perturbations on the spatial distribution of LSS
elements which are mainly determined by shorter wave perturbations.

2 Theoretical estimates.

Let us show that the spatial distribution of the LSS elements (galaxies) is governed by the
spatial distribution of the potential. Thus, Great Voids are assosiated with the regions of
the positive potential (around the points of maxima) and the regions with the high density
of galaxies (for instance, Great Attractor) are associated with the regions of the negative
potential (around the points of minima). The quantitative estimate of this influence can be
obtained by culculations of the mean linear density of pancakes as a function of potential.
The method of the calculation was described by Doroshkevich (1989).
According to the general catastrophe theory, the origin of the elements of LSS is deter-
mined by the largest eigenvalue An of the deformation tensor

(1)

were qi are the Lagrangian coordinates of the points. The elements of the LSS ( pancakes,
filaments) arise at the points of the maxima of An ( the earlier the larger a value All) and
extend over the surface
(2)
were Dl is the eigenvector corresponding to the eigenvalue An
As is known, the spatial distributions of the potential f1 and value An are correlated,
and it is possible to calculate the mean linear density of the surfaces (2) along a random
stright line. Evidently, this density is a function of the potential f1 and An. It is convenient
to use a dimensionless estimate Ul ofthe linear density n,( f1, All)

(3)

where T~ = 3./ < k 2 > and <> denotes averaging over the spectrum of perturbations.
Thus, Tc is the charactreristic scale of the spectrum. The value Ul depends also on the
dimensionless coefficient f3c = 0.2T~ < k4 > which influences the profile of the peaks An
The calculation shows that in the region f1 < 0 the function Ul can be written in the
form
(4)
where W is a part ofthe mass with f1 < f1*, All > Ail. For a wide class of spectra ( including
all standard spectra of CDM and HDM models and simple power spectra) f3c < 10. and a
is a weak function in the interval 0.7 ~ a ~ 0.55. But for the special spectra it is possible
to obtain very high values of f3c ( up to 103 710 4 ). In this case there is other approximated
relation a ~ O.114$c and a is independent of f1* and Ail
329

nv of the levels. = .0
To estimate the mean size of voids it is necessary to calculate the mean linear density
along a random line. This value is ( Demianski, Doroshkevich,1991)
(5)
were Lo = 1fL~/le , L~ = 5. < k- 4 >, I~ = (5/3) < k- 2 > and u~ is a dispersion of the
potential. ( if Le and u~ are infinite it is necessary to use a more complicated procedure
). Values Le and Ie ( as a scale Te) are determined by the spectrum of perturbations and
characterize it.

3 Discussion

The comparison of the typical scales obtained by the analysis of the deep surveys with
the above mentioned numerical estimates UI allows us to draw some conclusions about the
process of the structure formation and about the shape of the spectrum.
Firstly, these results show that the massiv pancakes which correspond to the large
value of ~11 > 2.UD, ( u1 is a dispersion of the diagonal components of the tensor du< )
occupy the region with the large negative value of. ( • < -u~ ). Such pancakes were
formed early and contained many galaxies. Probably such pancakes could be associated
with rich superclusters of galaxies. Smaller pancakes corresponding to the smaller values
of ~11 (1.5 < ~l1/UD < 2. ) occupy the regions with -1. < ./u~ < -0.25. Such pancakes
arose later, contained fewer galaxies and could be associated with poor superclusters and
filaments of galaxies.
In the regions with positive. the pancakes arose much later and probably the galaxy
formation in them happened very rarely.
Therefore, the influence of potential leads to the origin of the enormous regions which
practically don't contain the galaxies. It is natural to identify such regions as voids. A
level • ~ -0.25u~ can be accepted as a ( conditional) boundary of voids. Numerical
estimates show that such voids occupy about 70% of the space. The linear density of the
LSS elements in voids is at most 10% of that the regions between voids.
Nevertheless, let us note that the statistical character of the process of the pancakes
formation allows the origin of galaxies in voids with a low probability. Therefore, it is
more correct to treat voids as the regions with a small spase density of galaxies but not as
absolutely empty regions. Probably the observations of some peculiar galaxies in Bootes
confirm this point of view.
Secondly, it is very important to note that the above mentioned numerical estimates of
the UI don't take into account the time evolution of the LSS. This approach is correct for
the various variants of the HDM model in which the evolution of LSS is slow and a value
of Te is close to the correlation radius of the galactic distribution Tg . But if the spectrum
is similar to the spectrum of CDM models the linear density n, quickly decreases and
the correlation radius Tg quickly increases while the main part of the pancake population
is forming. However, one can expect that UI = Tgn, is a weak function of time and an
observational relation UI = Tgnl ~ 1. is approximately correct for a long time.
R this assumption is confirmed by the numerical simulations then it will be necessary to
accept a special shape of the spectrum of the initial perturbations in the shortwave region.
330

For instance, f3c can be large for the spectrum

p(k) = Ak- 1 (1. + k 2R~)-2 kmin < k < k maz , (6)

r~ ~ 3.R~/(1. + xm)ln[xM/(1. + xm)]


f3c ~ 1.8xM /(1.+ m) ln2[xM /(1. + xm)]
were XM = (kmazRc? > 1.,xm = (kminRc)2. Such an initial spectrum (p(k) ()( k- 1 )
differs strongly from the orthodox spectra of CDM and HDM models and can be formed in
complicated models of inflation ( Kofman, Linde, 1987, Mukhanov, Zelnikov,1990 ).
If the relation r9 ~ I is explained, then the observational estimates of the typical size
of voids L ~ 80h- 1 Mpc can be obtained for a wide class of the longwave spectra according
to the relation (5).
I am grateful to O.Buryak for helpful conversations.

References

Broadhurst T.,l., Ellis R.,S., Koo D.,e. &; Szalay A.,S., 1990. Nature, 543, 726.
Buryak O.E., Demianski M. &; Doroshkevich A.,G., 1991. in Observational Tests of Inflation, p. ,
ed. Shanks,T, Kluwer Academic.
Buryak O.E., Demianski M. &; Doroshkevich A.,G., 1991. Astrophys.l., in press, .
Demianski M. &; Doroshkevich A.,G., 1991. Preprint NORDITA., , .
Doroshkevich A.,G., 1989. Sov.Astron., 66, 449.
Kofman L.,A. &; Linde A.,D., 1987. Nuclear Phys., B 282, 555.
Mukhanov V.,F. &; Zelnikov M.,L, 1990. Preprint BROWN-HET-768, , .

DISCUSSION:

Huchra: All the models of biased galaxy formation that produce voids would predict that
the galaxies found in voids should be extraordinary (e.g. Dekel and Silk), but I would like
to point out that when we look in detail at the galaxies found in voids, they are remarkably
like the ones inside normal structure. Any theory, CDM included, must account for that.

Doroshkevich: The models of LSS formation describe the galactic origin only in general
terms. But it is known that galactic parameters depend on the parameters of the LSS - in
the poor filaments we don't see massive galaxies. Therefore the galaxies in voids may be
similar to those in filaments. Perhaps there are elements of LSS without visible galaxies.
It is very probable that such elements occur in voids.It is very useful to obtain statistical
estimations of such galaxies and their correlations with void size.
TESTING THE ZELDOVICH SPECTRUM

WILL SUTHERLAND
Dept. oj Astrophysics,
Keble Road,
OzJord OXl aRH

ABSTRACT. A variety of theoretical models are confronted with the correlation function results
from the APM Galaxy Survey. We can make rather good estimates of the post-recombination power
spectrum out to k '" 1/(100 h-lMpc); its slope is n '" -Ion cluster to supercluster scales, and rolls
over to n ~ 0 at larger scales, but cannot reach n '" 1 until k ;:; 1/(100 h-1Mpc)j thus the microwave
background appears to offer the most direct test.

1 Introduction

Most versions of inflation predict that the power spectrum of density fluctuations in the
very early universe should have the Zeldovich form P(k) oc k. The growth of perturbations
from the early universe to post-recombination is described by the transfer function T( k), i.e.
the post-recombination power spectrum is given by P(k) oc kT(k)2. Thus it is important
to note that we cannot test the Zeldovich spectrum independent of the matter content of
the universe; e.g. in CDM-dominated models, the shape of T(k) scales as 1/(flh2 ), while
observed lengths scale as h- 1 ; thus the power spectra of such models depend only on flh.
In the gravitational clustering scenario, the variances of galaxy counts, bulk motions
and Sachs-Wolfe microwave background anisotropies are related in linear theory by

v 2(r) oc b2 J
fl1.2
P(k)W(kr)dk

(~:):w (0) oc :2 J k- 2 P(k)WT(kORH )dk

(e.g Suto et al. 1988); here b is the usual bias parameter and W(kr) is a window function
depending on the smoothing; e.g. for a top-hat sphere, W(kr) = 9(kr)-6(sinkr-kr cos kr)2.
Thus the different observables are weighted towards different length scales.
In the above, we have normalised P( k) to the variance of galaxy counts; normalisation
to the mass would be more natural, but current observations of bulk motions are too few

331
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 331-335.
e 1991 Kluwer Academic Publishers.
332

to give a useful rms value, and no fluctuations are yet detected in the MWB: hence this
normalisation is more straightforward.

2 Comparison with Observations

We have evaluated the correlation functions and velocities for a variety of model power
spectra from Efstathiou (1990) and Holtzman (1989). We investigated CDM models with
various values of flh and n, hybrid CDM + HDM models, CDM models with high flbarJlOR
and a variety of isocurvature baryon models.
The power spectra are normalised in the usual way with oN/ N = 1 for a sphere of
r = 8 h-1Mpcj we evaluate {(r) using linear theory and then convolve to the angular cor-
relation function w(8) using Limber's equation, with the model galaxy luminosity function
of Maddox et al. (1990a).
Some example power spectra are shown in Figure 1; we show the 'standard' CDM model
with flh = 0.5, and several modified CDM models chosen to provide a reasonable match to
=
the APM w(8). These are a low-density model with flh 0.2, which requires a cosmological
constant (Efstathiou et al. 1990), a hybrid model with flCDM ::::J 0.7, flHDM ::::J 0.3, and
a model with flh = 0.75 and white-noise initial spectrum (which produces a large MWB
quadrupole but is very similar to the above on supercluster scales). We also show an
isocurvature baryon model with the most popular parameters fl = 0.2, n = -1. The
angular correlation functions for these models (at the depth of the Lick survey) are shown
in Figure 2 together with the APM results scaled to the same depth.

3 Discussion

There are a number of significant conclusions to be drawn from these comparisons:


=
a) The standard cold dark matter model with flh 0.5 is excluded, but variants with
flh = 0.2, or a hybrid CDM + HDM model, or a white-noise initial spectrum, can all give
an acceptable match to the APM result. [NB: Carlberg (priv. comm.) has recently claimed
that CDM with flh = =
0.5 and b 0.8 can match the data, due to non-linear clustering
producing large-scale power, and velocity biasing suppressing the otherwise excessive small-
scale velocities. 1
b) Isocurvature baryon models are not favoured by the data; models with primordial
spectrum n '" -1 give excessive large-scale power, while those with n ~ -0.5 give too little
power on intermediate scales. However, in some such models even the APM Survey may
not give a 'fair sample' so this conclusion needs to be verified via numerical simulations.
c) The various models which match the APM results produce very similar predictions of
hN / N (r) as expected. Since the power spectrum seems to roll over to n ~ 0 at the largest
scales, these scales should not contribute greatly to bulk motions so we can also predict
approximate rms values for v(r) up to the factor flo. 6 /b. Some typical values for top-hat
spheres are :
333

r/ h-1Mpc fJN/N v [00.6/b = 0.4]


25 0.40 350
50 0.18 250
100 0.07 180
This enables some interesting tests of Gaussian initial conditions: firstly we can predict
that voids of r ~ 40 h-1Mpc should not exist, and secondly we can perform N-body simula-
tions with a reasonable knowledge of the initial power spectrum, to test for the abundance
of Great Walls etc.
d) The various data sets are consistent; power spectra matching the APM data also
give reasonable agreement with the QDOT redshift survey of lRAS galaxies (Saunders et
al. 1991), the streaming motion data, and the radio galaxy survey of Peacock & Nicholson
(1991). Prelilninary results from the Stromlo/ APM redshift survey (Loveday et al. 1991)
and the COSMOS galaxy survey are also encouraging.
e) The one disappointing conclusion to be drawn from this work is that the Zeldovich
spectrum can only be apparent at wavenumbers k ~ 0.01 h Mpc 1 • This means that it will
be exceptionally difficult to directly detect via galaxy clustering or bulk motions, thus the
microwave background appears to offer the most promising test (Efstathiou, this volume).

Acknowledgements. I wish to thank George Efstathiou and Steve Maddox for many discussions,
and acknowledge the receipt of a SERC Fellowship and a Research Fellowship at Lincoln College,
Oxford.

References

Efstathiou, G., 1990. in Physics of the Early Universe, eds. Peacock, J. et al ..


Efstathiou, G., Sutherland, W.J. & Maddox, S.J., 1990. Nature, 348,705.
Holtzman, J., 1989. Astrophys. J. Supp., 71, 1.
Loveday, J., Peterson, B.A., Efstathiou, G. & Maddox, S.J., 1991. in preparation
Maddox, S.J., Efstathiou, G., Sutherland, W.J. & Loveday, J., 1990. Mon. Not. R. astr. Soc, 242,
43P.
Peacock, J., & Nicholson, R., 1991. Preprint.
Saunders, W., et al. , 1991. Nature, 349, 32.
Suto, Y., Gorski, K., Juskiewicz, R. & Silk, J., 1988. Nature, 332, 328.
334

-COM 0.5
.. -.......... •••••.CDM 0.2

.. / / '-\ __COM 0.75 "-0


./ \
. . -·-.:r:·:.~.·.~~:::::=;=-.~.~::~-. ..... ._.P1S 0.2
:,-:::::::.:......

lot k / h Mpc·'

Figure 1: Power spectra for typical models as described in Section 2. The units of
the ordinate are arbitrary but the relative normalisations are given by unit variance in
8 h- 1 Mpc spheres.

-COM

O·o~ -;.'
0 . •••••COIA 0.2

• •••• COIA 0.75 n-O


° o :~:. "
. :': . . . .. COIA + HOM
..... .:, ~....
..
i
. . :;~3'
•..•• PIB 0.2
•• .:-t"~
.
:§: • 0
~
"0;';" .
l'"
....... "'~.'"
a£ao.,
...-....
N .,~
I
~
:~;
0,,0 •
. ~.~
oo:o:~~.\ :0_0,:
7_~1~~~--~~~~~~~~--~------~~~-·~·~·~~----~~
-0.5 0 0.5 1.5 2

(~ &
Figure 2: Angular correlation functions scaled to the Lick survey depth: the data
points are from the APM Survey, and the lines show the same models as Figure 1.
335

DISCUSSION:

Peacock: I'm not a shareholder in CDM inc., but I do wonder if your results exclude
standard CDM convincingly. All the problems at large 8 in w(8) can be mended by increas-
ing CDM 6p/ p by a factor 2. Problems then appear on small scales, but things are then so
non-linear that it is much harder to be sure there is a problem.

Sutherland: But you then get W (16 h- 1 Mpc) > 1, which is clearly a problem unless
you have anti biasing on such scales.
Shanks: I am also no friend of CDM but I worry about your rejection of that model from
w(8) in a range where its value is of order 0.01. 10% plate-to-plate variations (±0.m1) would
be enough to produce the large-scale power that you see in the APM w(8). Although I
know you have done a very careful job in trying to minimise systematics, I still worry about
the difficulty of keeping systematics in check over such a large survey. Further the results
from the correlation analyses in the Durham Redshift Surveys and in the Las Campanas Z
Survey show smaller power at 1O-20h- 1 Mpc than APM.

Sutherland: I would like to refer you to our 20-page paper on this which is nearly finished.
Briefly, I would emphasize two points.
(1) The 10% errors would have to be correlated across plates which don't even overlap to
get our observed power at '" 10° scales.
(2) We can simply subtract the correlation function of our faint ('" 20m ) galaxies from
that of our bright ('" 18 m ) galaxies; this makes little difference to our result, and therefore
eliminates all errors which are common to bright and faint galaxies.

White: If one assumes that the power spectrum derived from your APM data rolls over
smoothly to the Harrison-Zel'dovich slope on scales beyond 100h- 1 Mpc upper limits on
the quadrupole in the microwave background must translate into lower limits on the bias
parameter, b, at the normalisation scale. What are these limits?
Sutherland: That's part of George Efstathiou's talk!
n ON THE SCALE OF 3 Mm/s

D. LYNDEN-BELL
Institute of Astronomy
Madingley Road
Cambridge, CB3 OHA.

ABSTRACT. The finding and naming of the Great Attractor is briefly reviewed. Although
originally found as a major feature of the optical galaxy distribution, it is shown to be the
most prominent feature in the all-sky distribution of IRAS galaxies.
A map of the surface density of the Supergalactic Plane towards Hydra and Centaurus is
presented.
Optical and IRAS determinations of no. 6 /b are compared. It is evident that the bias
factor b is different for optical and IRAS galaxies. It would then be surprising if the b of
IRAS galaxies is precisely unity.

1. The Extragalactic Velocity Field

Following Slipher's (1917, 1922) demonstration that most galaxies were moving away from
us, Hubble set about determining their distances and in 1929 published Ius result that the
velocity of recession was proportional to distance. Hubble actually had two problems, the
most difficult being that of the distance scale (which he got badly wrong) but the easier was
the finding of the zero point from which the expansion velocity should be measured. Hubble
reduced the velocities to the Heliocentric frame and then assumed that the Sun had its own,
as yet unknown, peculiar motion, v0, relative to the zero point. The cosmic expansion
velocity of a galaxy observed in direction r with a heliocentric radial velocity VH is then

so Hubble's law becomes


(1.1)
This law has four unknowns V0 and H, so one needs at least four galaxies with known
distances and heliocentric radial velocities to determine them. In practice Hubble had 22
galaxies so the equations (1) were inconsistent when considered as exact. Therefore he made
them as true as possible by minimising the sum of the squares of their residuals. Hubble's
result for V0 was in modern notation 305 km/s towards a new galactic longitude of (l,b) =
(71, -14). This is close to the standard IAU de Vaucouleurs value of 300 km/s (90, 0) or
in cartesian coordinates (V cos l cos b, V sin l cos b, V sin b) = (0,300,0). Notice that Hubble
determined the zero point of the Hubble flow which is in principle different from the velocity
of the Local Group since the Local Group might have its own peculiar motion. However,

337
T. Shanks et al. (eds.). Observational Tests of Cosmological Inflation. 337-353.
e 1991 Kluwer Academic Publishers.
338

the velocity of the Sun relative to the Local Group is 299 (96, -5) == (-30,297, -24) with
errors of ±50 km/s so the difference between the IAU standard zero point and the velocity of
the Local Group is zero within the errors. With notable exceptions (e.g. Richter, Tammann
et al. 1987) most astronomers neglect this difference completely. Thus prior to 1973 there
was general agreement that the Sun was moving at about 300 km/s toward (l,b) == (90,0).
Then Rubin, Ford & Rubin found that relative to a shell of ScI galaxies at rv 3500 km/s the
Sun appeared to have quite a different motion. While argument continued over the galaxy
distribution (Fall & Jones 1976) and more seriously, over the magnitude system by which
the galaxies were selected, a still more fundamental result emerged from radio astronomy.
First Conklin (1969, 72) showed that the Right Ascension of the Sun's motion relative to the
cosmic microwave background (CMB) was at 11h rather than the 21h of the standard IAU
motion. However his correction for Galactic emission was as big as his signal. He was right,
but was not believed. Then Corey & Wilkinson, and Smoot et al. showed by balloon and
high flying aircraft that the Sun was moving at 365 km/s with respect to the CMB at 11th
and fJ == _6 0 or in galactic coordinates 365(267,50). Although the speed was similar to the
optical result the direction was utterly different, 130 0 away, not too far from opposite. The
standard (0, 300, 0) zero point was now seen to be moving at 600 km/s towards (269, 28)
relative to the CMB, (see Figure 1).
At first optical astronomers blamed much of this difference on the gravity of the nearest
big cluster of galaxies, Virgo, but as the radio data became firmer the component transverse
to the Virgo direction became established. Also as more Tully-Fisher distance estimates
became available, it became clear that the Virgo infall at the Sun was lower than the early
estimates of 300-500 km/s. Thus most of the Local Group's velocity had to be attributed
elsewhere. Both Shaya & Tully and Tammann & Sandage suggested that much of this motion
was generated by the Hydra-Centaurus supercluster identified by Chincarini & Rood. Lilje,
Yahil & Jones found a tidal distortion of the Virgo inflow that gave quantitative backing to
this. Meanwhile Yahil, Walker & Rowan-Robinson gave the first measurement of the dipole
of the whole-sky distribution of IRAS galaxies and deduced that the density of the universe
must be close to the closure density. As one who believes that n should be greater than
one for theoretical reasons, quite apart from inflation, I believe this paper will be looked
back upon as an important landmark in the subject. However, the method is not without
difficulties to which I will return.
When the Samurai plotted their elliptical galaxy peculiar velocities in the CMB frame,
a strong streaming motion of 521 km/ s towards (l, b) == (307,9) was clearly seen. This
extended at least as far as the Centaurus double cluster named Cen 30 & Cen 45 by Lucey
et al. At the time we believed that the Hydra-Centaurus supercluster consisted of these
clusters and a broad arm across the sky to the strong cluster in Hydra. The fact that we
found the Centaurus double cluster to be moving rapidly away from us after subtraction
of the mean Hubble flow, indicated that the centre of gravity of the any attractor must
lie beyond them (i.e. beyond Hydra-Centaurus ). Some of the difficulties of nomenclature
in this region stem from the fact that many others had included the more distant clusters
in Centaurus in their definition of Hydra-Centaurus whereas we did not realise that this
was done. We thought that the moving of the double cluster implied that, for instance,
Tammann & Sandage were wrong in attributing a significant part of the Local Group's
motion to Hydra-Centaurus. In this we were wrong. Sandage & Tammann had defined
what they meant earlier and they had included in their definition of Hydra-Centaurus a
339

number of clusters stretching out to 5000 km/s and their environment. The bulk flow of the
Samurai ellipticals was not in the same direction as the Local Group's motion towards Hydra
(269,28) but was 42° away from it towards Centaurus , (see Figure 1). At first the Samurai
modelled the gravity field that might cause the motion as that from a density distribution
that fell off as r- 2 from a centre constrained to be in the direction of the streaming motion.
Later that constraint was relaxed and the best fitting model for the gravity field was centred
at (309,18) and the density distribution had a core radius of a third ofits distance. Thus the
core diameter subtends 40° in the sky! The distance to the centre corresponded to 4000±300
km/s of pure Hubble flow (Faber & Burstein 1988). (Note this method of quoting distances
does not change when Hubble's constant is found to be different). Others call this distance
40hlo~ Mpc.
+90·

i-t-IHO·

-90·
Figure 1. Directions of solar motion in Galactic Coordinates.
H Hubble's determination relative to the best fitting Hubble flow of all nearby galaxies.
V de Vaucouleurs standard IAU direction of the Sun's motion relative to galaxies.
T The motion relative to the Local Group
o The Sun's motion relative to the Cosmic Microwave Background (CMB).
+ The Local Group's motion relative to the CMB .
• The direction of the IRAS dipole determined by Rowan-Robinson et al. Theory suggests
this should coincide with +.
P and Q+ show the directions of the optical dipole and the longest positive principal axis
of the optical quadrupole as a function of the velocity limit n X 103 km/s out to which
galaxy light is summed. For P,oo should coincide with +.
00 is the corresponding 'Mean sky' result.

Armed with a direction and an approximate distance to the mass centre, the Samurai
looked at the all-sky distribution of galaxies that had already been plotted by Lahav (see
his map in Lynden-Bell 1986). The relevant direction was close to the divide between
340

Fisure 2a. The 0Jisinal picture of the hemilphere of the .ky c:eatred OIl 307,9 from which
the Great Attractor wu umed. Each point it a ,aluy whole diameter it > 1 arc: minute.
Stamn, to the ript of the Vup dUlter V the Great Attractor extenda put the Centaurua
double dUiter C ud i. 100t behind the dark band of ObecuratiOil in the vertical ,alactic
plue. It re-appean OIl the other aide in Teleacopium-Pavo-IndUl.
341

Figure 2b. Caleb Schan haa decompoeed the mAS galaxy distribution with 601' llux
> 0.6 Fu into spherical harmonica Yt.... When the sky is reconstructed ung only the 120
components with I ~ 10, its primary features sta.nd out strongly. The greatest of these
coincides with the centre of the Great Attra.c:tor . The projection a.nd orientation are those
of Fig. 2a.
342

hemispheres but when Lahav replotted the data with (307, 9) as the centre, the resulting
view of the sky was remarkable (Figure 2a). Not only is the direction of galaxy streaming
very close to the densely populated Supergalactic Plane (which covers the Galactic Plane at
(317,0», but there is a strengthening and broadening of the supergalactic band in northern
Centaurus . Starting just to the right of the Virgo cluster V this 'Great Attractor ' extends
past the Centaurus double cluster C whence it is lost behind the band of obscuration caused
by the galactic plane. However, it re-appears on the other side in Telescopium-Pavo-Indus.
In spite of this large scale enhancement in the galaxy density, the Samurai pointed out
that there was no very prominent central tight cluster dominating this distribution. The
Centaurus double cluster is on one side too close and moving while, as we shall see, Shapley 8
is considerably too far away.
Seeing that the centre ofthe apparent distribution on the sky lay in Northern Centaurus ,
the Samurai sought redshift data towards the Great Attractor's centre. Dressler had already
begun a large scale survey in the area but he was still in the observing phase. Da Costa et al.
had completed their survey in the central region. The distribution showed a dominant peak
at 4350 km/s heliocentric with a spread of (j = 1000 km/s (Figure 3). This corresponds
to 4650 km/s in the CMB frame. We felt this was good confirmation of our distance to
the Great Attractor. Later, Melnick and Moles observed much deeper in the direction of
Shapley 8 which was one of the largest concentrations of galaxies known in the 1930's and is
rivalled only by one or two others even today. This also lies just north of Centaurus but at
a redshift of 16,000 km/s. Recently Raychaudhury has made deep surveys of Schmidt plates
in the whole Hydra-Centaurus region. From the light he concludes that the whole Shapley
Supercluster can not produce more than 15% of the optical dipole so it is unlikely to be the
major contributor to the Local Group's motion as Scaramella et al. had speculated might
be the case. The Shapley concentration, although very massive, is too distant to produce
major streaming here. Since the Great Attractor was named from the all-sky distribution of
known galaxies, the rumour that it is invisible (Hoffman and Zurek 1988, Rowan-Robinson
1990) is untrue.

" 2500 5000 7500 10000 12500 15000 17500 20000

v (KM/S)

Figure 3. The velocity histogram of da Costa et al. in the direction of the Great Attractor.
343

Rowan-Robinson et al. (1990) have also claimed that the Great Attractor is invisible in
the distribution of IRAS galaxies. It is true that the distribution of IRAS galaxies over the
sky is blander than the optical picture. However, Caleb Scharf has recently resolved the
distribution of IRAS galaxies into spherical harmonics. Figure 4 shows his reconstruction
using just the dipole and quadrupole terms while Figure 1b shows the reconstruction using
the 120 coefficients of the Y lm with l :5 10. Almost all the peaks correspond to named
objects in the optical sky and the strongest of all sits on the Great Attractor and not on
the nearby Centaurus cluster C. Figure 1b is in the same projection as Figure 1a. Thus
the IRAS data themselves show that the Great Attractor is the largest concentration of
galaxies in the nearby universe « 10,000 km/s). (The only remaining doubt is the distance
which depends on the survey by da Costa et al.). An interesting consequence of Figure 4 is
that when the quadrupole is added in, the peak moves from near Hydra to sit on the Great
Attractor. Furthermore, another peak Perseus-Pisces lies opposite in the sky. Thus the
Local Group is the subject of a tug-of-war between the Great Attractor in Centaurus and
the Perseus-Pisces supercluster. Although the Great Attractor is winning, the balance of
forces is such that repulsion away from the voids around l = 75 b = ±30 and attraction to
the interesting cluster on the Galactic plane in Puppis (the second strongest peak in Figure
2b) can shift the net motion of the Local Group from Centaurus into Hydra.
New data by Dressler & Faber and by Mathewson, Ford et al. have shown streaming
motions continuing to considerable depths where Malmquist corrections are very important.
Unbiassed methods of reduction are necessary to put the claimed backside infall beyond
contention. It is not at present clear where streaming ceases although it does so somewhere
between 3500 and 5000 km/s. The turnover point in the v versus r diagram may well vary
with the area of sky surveyed. Mathewson considers that streaming continues outwards and
ceases well beyond the major enhancement of galaxy density.
NORTII SOUTH

Figure 4. Scharf's reconstruction of the number density of IRAS sources over the sky using
just the 1 monopole, the 3 dipole and the 5 quadrupole coefficients. The projection is equal
area centred on the N and S galactic poles. The numbers are galactic longitude. The galactic
plane lies around the circles and the Supergalactic plane is horizontal. The Great Attractor
is the strongest feature in the northern sector while Perseus-Pisces is almost diametrically
opposite in the Southern sector.
344

2. The Density in Velocity Space

To get a good idea of the nature of the Supergalaxy we need a 'face on' picture of the super-
galactic plane. Optical surveys are not yet sufficiently well selected to do this accurately but
the combination of Dressler's Supergalactic Plane Survey with the Southern Redshift Sur-
vey does allow us to picture the overdensity in the most relevant sector of the Supergalactic
Plane.
Let n( v, r, D)v2dv dw dD be the number of galaxies with radial velocities v and diameters
D in a solid angle dw about the direction r. Assume a universal diameter function so that
the fraction of galaxies with diameter D is 4>t(D)dD. Then
n(v,r,D) = n(v,r)4>t(D) (2.1)
Approximately the distances will be proportional to the velocities relative to the Local Group
so the apparent diameter fJ will be proportional to D/v. We write
(PI (D) dD =</>{ vfJ)d( vfJ) (2.2)
H a known fraction S(fJ) of the galaxies of apparent diameter fJ are selected for redshift
determination, then the number observed will be
No(v,r,8)dvd8dw = S(8)n(v,r)4>(v8)v3dvdfJdw. (2.3)
Integrating over all fJ and solving for n( v, r) we have

( ") _ JNo(v,r,fJ)dfJ (2.4)


n v,r -J S(8)4>(vfJ)v3dvdfJ'
This is the formula that we use to determine n from the observations. For a given survey
S(fJ) is either known initially or determined directly from the ratios of the numbers in the
redshift list to those in a complete diameter limited catalogue such as ESO.
The diameter function 4>( vfJ) can be found from a redshift survey that is complete to
a given minimum cut-off in apparent diameter fJ m • A pretty way of doing this is my c-
method which gives the cumulative t(> vfJ) in an unbiassed manner without assuming any
parametric form (Lynden-Bell 1971, see also Jackson 1974 & Choloniewski 1987). I shall
now explain this. We need a name for vfJ and by a slight change of definition to absorb the
constant of proportionality we shall call it D.
The numbers of objects observed in such a complete survey will be
(2.5)
where
U(X)={1 x~O
o x<0
Our mathematical problem is given the observed N2 to find the best estimates of 4>(D)
and N1(v). The total number of objects observed with diameters vfJ less than D' will be

N3« D') = LL
D' 00

N2(V,D)dvdD. (2.6)
345

'))'

Figure 5. The data occupies the region above the diagonal cut-off.

Consider the number of objects, C(D'), in the box with D > D' and v < D'18m • In any
vertical strip the profile of N2 will be proportional to 4>(D) wherever we are higher than
the cut-off D = v8m • Hence the number in the horizontal strip at D' divided by C(D') will
reflect this profile. Calling the integrated diameter function <1>(> D)

-d<1> _ dN3 (D')


(2.7)
<1>(> D') - C(D')

Hence
<1>(> D) = exp -
r
D dN3
iDa C(D') (2.8)

where <1>(> D') is the fraction of objects with diameter> D and we have used the boundary
condition that <1> = 1 for D < D m , the least diameter observed. Clearly we can only
investigate the form of <1> for D > D m •
It is quite unnecessary that there should be large numbers of objects in order that this
method produces an answer. As D' comes over a given narrow strip, C decreases by f:lN3 so
that the integral changes by In (C-t:~N.) where C-(D') is the number not counting those
on the edge. Applying this formula point by point, we have

<1>(>D)= II ___
D.<D C-
i_ (2.9)
i Ci +1
where C; are the number of points in the box when the ith point is on its bottom edge (which
is not counted in C-) and the terms in the product are only those with Di < D. This formula
produces a histogram for <1>(> D). The differential distribution 4>(D) = -d<1>ldD can be
obtained by drawing a curve through the histogram and differentiating it.
346

An amusing consequence of the method is that it can be used just as effectively the other
way around to determine the function N1(v). Now the corresponding function C- must
be made a function of v8m rather than D and the jumps in J N1dv occur when there is a
point on the vertical edge of the box rather than the horizontal one. Finally, the integral of
Nl is not a fraction so the normalisation must be chosen to give the observed total number
of points in the region D > v8m • Working the method this other way around is a useful
exercise for those new to it. With the determination of Nl comes an estimate of the number
density of galaxies with diameters> Dm per unit volume of redshift space. This is what is
needed in Section 3. It is also what is needed to calibrate formula (2.4) in terms of the mean
density.
Mike Hudson has computed both the diameter function and the mean density from the
Southern Sky Redshift Survey. He then applies these to determine the density from Dressler's
Supergalactic Plane Survey using formula (2.4). Figure 6 shows his contours of the surface
density within 500 km/s of a supergalactic plane defined by a pole at f = 47 b = 21. This
change of supergalactic plane is because the main run of overdensity in Pavo lies at an angle
to the conventional de Vaucouleurs supergalactic plane, so the southern end of Pavo would
be missed by that plane.

New SGL

Figure 6. Hudson's chart of surface density of optical galaxies within ±500 km/s of
a Supergalactic plane, deduced from Dressler's Supergalactic Plane Survey. Velocity in
the Local Group's frame is plotted radially. To put these velocities in the CMB frame
approximately 500 km/s should be added. The lowest contour is at the mean density of the
SSRS Survey, the highest at six times that density. V marks the approximate position of
the Virgo Cluster.
347

The region close to SGL = 180 has been interpolated from the surface densities on either
side, so there is no certainty that the overdensity falls below three times the mean between
Centaurus and Pavo as the picture indicates. The radial velocities are in the Local Group's
frame. In these directions some 500 km/s should be added to get motions in the cosmic frame
defined by the microwave background. In Centaurus the peak is therefore close to 3000 km/s
with a tongue of density over three times the mean extending behind it to 4200 in the Local
Group's frame, i.e. 4700 in the microwave frame. It must have been this tongue that was
particularly well represented in the survey by da Costa et al. However, it is noticeable that
the contours close to 6000 km/s are almost circles about the the origin. It seems just possible
that Dressler's survey is somewhat too shallow to give a good account of the density out to
8000 while da Costa's went a fraction deeper over a smaller area.
Another interesting facet of this figure is that it lends weight to the Samurai's model
that the Virgo cluster is an outlier of the Supergalaxy whose centre lies in Centaurus, but
probably closer to 3000 km/s in the CMB frame than to the 4000 suggested by the Samurai.
Here the more speculative join to Pavo becomes interesting since the maximum density there
is beyond 4000 even in the Local Group's frame. It is abundantly clear that the Supergalactic
Plane is overdense out to at least 6000 km/s. This is not a great surprise since it was selected
and named for its over density.

3. Dipoles

Each galaxy is a source of both gravity and light. Both falloff as r- 2 so they remain
proportional at all distances. Furthermore, on average, galaxies of twice the mass produce
about twice the light, just as if each was made of two smaller ones. Gott suggested that if
the directed light fluxes (per unit area) from all external galaxies were observed and added
vectionally, the result would be proportional to the net gravitational field at the Local Group.
If F is this total net flux one has

-g=G<M/L>F (3.1)

where M is the mass associated with a galaxy of luminosity Land < M / L > is the average
mass-to-light ratio. The mass must include all dark matter haloes and may include some
more widely distributed material either dark or too faint to be seen. Thus the average
< M / L > may be significantly larger than that usually associated with a single galaxy. If
< PL > is the mean luminosity density produced in the universe and OPL the departure from
that mean, then biassed galaxy formation would lead one to expect that the fluctuations in
luminosity density may be greater than fluctuations in the mass density p. To allow for this
one writes
OPL/ < PL >= bOp/ < P > (3.2)
where b, the bias factors, would be unity if mass followed light exactly as Gott assumed.
The uniform mean density produces no net directed light flux and no net gravitational
acceleration relative to Hubble flow. Thus in the presence of bias (3.1) is modified to

_ = Ci/M)F= Ci <p> F. (3.3)


g b \ L b < PL >
348

If the brightness of the extragalactic sky O"(l, b) is expanded in spherical harmonic compo-
nents in the form
0"( l, b) = M + P . r + r . Q . r + .... (3.4)
where r is the unit vector (coslcosb,sinl,cosb,sinb) then M is the monopole, P the dipole
and the traceless symmetric tensor Q is the quadrupole component of the brightness. The
net directed flux of light here depends only on the dipolar component and taking the z axis
along P it is readily integrated to give F = -!71" P. Thus Gott's relationship can be written
directly in terms of the light dipole.

_ 471"G < P> P


(3.5)
g- 3b < PL > .
Now Peebles has time-integrated the linear equations of motion of growing density pertur-
bations in the expanding universe to give the relationship between current velocity, v and g
as
HnO•6 nO•6 H
(3.6)
v = 471" G < P > g = 3b < PL > P

where H is Hubble's constant and n the ratio of < P > to the critical density Pc =
3H2 1(871" G). We would like to determine n and equation (3.6) is the key.
Now estimates of < PL >, the mean luminosity density of the universe, are dependent on
the Hubble's constant H because volumes scale as H- 3 and luminosities corresponding to
known apparent magnitudes scale as H-2. Thus their ratio < PL > scales as H. Although
both < PL > and H are notoriously hard to measure, the combination < PL > 1H that
appears in (3.6) can be determined since no absolute distance measurement is needed. Thus
no.6 lb can be found from applying (3.6) to the motion of the Local Group through the
CMB and if the objects considered are unbiassed (b = 1), we can obtain n provided we can
measure P. We now turn to the determinations of the dipole P. Different determinations
are listed in Table 1 by date. Each has its own difficulties and compromises so we discuss
the relative merits of the IRAS and the optical determinations.
The great advantage of IRAS is that the measurements were made over almost all the
sky with one instrument and one reduction method. Furthermore, as a result of redshift
surveys the luminosity function is quite well known. A disadvantage is that the luminosity
function is very broad and the mass to 60Jlluminosity ratios vary violently from star-burst
galaxies in which it is very low, to ellipticals in which it is very high. The idea that mass
directly follows 60Jl flux is certainly wrong on an individual basis and something must be
done to cool off the signal from the brightest 30 sources which have so much flux that they
seriously influence the result. Faced with this problem YWR put the sources into bins by
flux. Worked out the number weighted dipole for each flux bin weighted these with the
average flux per source in these bins and added the results. This is a clever thing to do
with noisy data but it does not preserve the significant 'shot noise' gravity of the nearest big
galaxies outside the Local Group which should affect the result. Meiksin &: Davis gave up
using flux altogether and measured a number weighted dipole. Strauss, Davis &: Yahil and
Rowan-Robinson et al. have measured redshifts of IRAS galaxies, made a model of the mass
distribution and summed r- 2 over the sources above some flux limit after suitably weighting
them to allow for the broad IRAS luminosity function. Again this is statistically correct but
does not give the mass weighted nearby 'shot noise' contribution to the gravity field. Most
349

TABLE 1. Optical, IRAS and X-ray Dipole Determinations

Date Authors Catalogue A Method N/100 l b % nib!


Data

1977 Smoot et al. CMB M F 269 28 .1 -


1980 Yahil,Sandage, RSA 0 F 12 217 78 -* .2
Tammann
1982 Davis,Huchra CfA 0 F 12 220 69 -* .2
1983 Shafer ,Fabian XRB X F 277 19 .5 -

1986 Yahil,Walker, IRAS FB 80 248 40 - .85


Rowan-Robinson
1986 Meiksin,Davis IRAS N 90 235 44 12 ~.5

1987 Lahav UGC,ESO, 0 82 169 227 42 49 .3


MCG> 1.3'
1987 Harmon,Lahav, IRAS I F 90 273 31 34 -
Meurs
1987 Villumsen,Strauss IRAS F 80 239 36 - 1.2

1988 Lahav,Rowan- UGC,ESO, 0 82 169 257 34 27 .15


Robinson & MCG> 1.3'
Lynden-Bell IRAS > .6Jy I FB 80 249 38 16 .8
1988 Strauss, Davis, IRAS > 1.9Jy I N/r2 20 231 48 - .5
Yahil
1988 Plionis Lick counts 0 N 810 237 10 5t -
1988 Lynden-Bell, UGC,ESO 0 82 240 261 29 24 -
Lahav MCG> 1.03

1989 Lynden-Bell UGC,ESO}Mock 0 82 240 275 26 41 .14


Lahav,Burstein MCG,ZCAT}Mean 0 82 240 257 37 25 .3
1989 Lahavet al. X-ray clusters X LtT-h- 2 .53 258 76 -

1990 Miyaji,Boldt XAGN X F .7 299 35


1990 Rowan-Robinson IRAS > .6Jy I N/r2 21 237 43 .7
et al. lin 6

1991 Plionis, Abell,ACO 0 Ngfr2 1.5 255 35 t .06


Valdarnini clusters

* Sample too shallow to catch some of the dipole.


t Sample that concentrates on dipoles from very distant objects.
350

elliptical galaxies are not included as they give too little flux and galaxies of twice the mass
do not give a double contribution to the result. Because redshifts have to be measured, the
number of galaxies used to estimate the dipole is far less than in the original catalogue. One
in six for the deeper survey. Although absorption is not a problem at 60J,l, confusion with
galactic cirrus and with stellar sources is a problem at low galactic latitudes. The carefully
colour-selected sample of Harmon, Lahav & Meurs allowed them to go somewhat closer to
the plane but they had to cut out the highest flux sources, most of which were cirrus, but
some of which were galaxies! They got a dipole surprisingly close to the direction of the Local
Group's motion but some bright galaxies were also excluded by their colour criterion. Other
IRAS dipoles have been very consistent in getting dipoles some 35° away from that motion.
There is then some quandary in using the data to deduce nO. 6 lb. H one resolves the dipole
to get its component along the velocity, then the resolution produces a factor (cos 35)-1 in
nO. 6 Ib but if one resolves the velocity along the dipole then the factor is cos 35° . The ratio of
these estimates is cos 2 35 in no.6lb or cos 3 •3 35 = 0.52 in n/b!. Thus significant error occurs
automatically if the dipole and velocity do not align as the theory predicts. Misalignment
also causes a crisis of confidence in the whole method. Two random directions have a chance
of 1 in 11 of marking an angle < 35° purely by chance but a chance of only 1 in 132 of being
better aligned than 10° . Most optical dipoles have been based not on optical flux but on the
squares of the galaxies' angular diameters. These give the correct r- 2 weighting and the use
of major axis diameters makes them insensitive to internal absorption and orientation unlike
apparent magnitudes. Since galaxy masses vary approximately as their diameters squared,
this method weights bigger galaxies appropriately. The major problems lie in homogenising
the measurements from three different galaxy catalogues made by different techniques with
different diameter measures. The MCG catalogue is less deep than the other two and there
is a swathe of no data within 15° of the Galactic Plane.
Altogether about 1/3 of the sky has either poor data or no data. Two procedures have
been used to assess the problems caused by this. 1) Replace such areas with Mean Sky
the average and then measure the dipole. 2) Clone strips of equal area just above and
below the missing galactic strip and move the clones into the missing strip. Then measure
the dipole from this Mock Sky. The difficulties are best illustrated by the results in which
no. 6 /b is 0.49 by the Mean Sky method but 0.31 by the Mock Sky one. The new survey
of the +2.5 > b > -17.5 strip now being completed in Cambridge and Oxford will overlap
with the UGC and ESO catalogues and enable the diameter systems to be compared. A
uniformly calibrated all-sky optical catalogue of galaxies with diameters greater than 1 arc
minute should soon result.
In spite of the difficulties, the optical results have consistently given directions within 15°
of the Local Group's motion. However, for b = 1 they give n ~ 0.2 ~ 2 whereas the IRAS
results give n = 0.7 ± .3. There is direct evidence that the IRAS galaxies show less structure
on large scales and this is no doubt reflected in b for optical galaxies being greater than b for
IRAS galaxies. It could be that all galaxies are biassed but that IRAS galaxies, which are
less common among galaxies in big clusters, are just sufficiently anti-biassed with respect to
all galaxies that they end up unbiassed tracers of the mass. Dekel gives some evidence that
this may be the case but if so, it looks like a lucky fluke rather than anything fundamental.
Methods of measuring the bias for IRAS galaxies are much needed.
351

References

Boldt, E. (1987), Physics Reports 146, 215.


Chincarini, G. and Rood, H.J. (1979), Astrophys. J. 230,648.
Choloniewski, J. (1987), Mon. Not. R. astr. Soc. 226, 273.
Conklin, E.K. (1969), Nature 222, 97l.
Conklin, E.K. (1972), 'External Galaxies and Qsos', in D.S. Evans (ed.) IAU Symposium
44, Reidel, Dordrecht, p. 518.
Corey, B.E. and Wilkinson, D.T. (1976), Bull. Am. Astron. Soc. 8, 35l.
Da Costa, L.N., Nunes, M.A., Pellegrini, P.S. and Willmer, C. (1986), Astr. J. 91,6.
Da Costa, L.N., Willmer, C., Pellegrini, P.S. and Chincarini, G. (1987), Astr. J. 93, 1338.
Dressler, A. (1988), Astrophys. J. 329, 519.
Dressler, A. and Faber, S.M. (1990), Astrophys. J. 354, 13.
Dressler, A., Lynden-Bell, D., Burstein, D., Davies, R.L., Faber, S.M., Terlevich, R.J. and
Wegner, G.A. (1987), Astrophys. J. 313,42.
Faber, S.M. and Burstein, D. (1988), in G.V. Coyne and V.C. Rubin (eds.), Large Scale Mo-
tions in the Universe, Proceedings of the Vatican Study Week, Princeton University
Press, Princeton.
Fall, S.M. and Jones, B.J.T. (1976), Nature 262, 457.
Hoffman, Y. & Zurek, W.H. (1988), Nature 333, 46.
Hubble, E. (1929), Proc. Nat. Acad. Sci. US 15, 168.
Jackson, J.C. (1974), Mon. Not. R. astr. Soc. 166, 28l.
Lahav, O. (1987), Mon. Not. R. astr. Soc. 225, 213.
Lahav, 0., Rowan-Robinson, M. and Lynden-Bell, D. (1988), Mon. Not. R. astr. Soc. 234,
677.
Lahav, 0., Edge, A.C., Fabian, A.C. and Putney, A. (1989), Mon. Not. R. astr. Soc. 238,
88l.
Lauberts, A. (1982), The ESO-Uppsala Survey of the ESO(B) Atlas, European Southern
Observatory.
Lawrence, A., Saunders, W., Crawford, J., Ellis, R., Frenk, C.S., Parry, I., Xiaoyang, X.,
Allington-Smith, J., Efstathiou, G. and Kaiser, W. (1990), Mon. Not. R. astr. Soc.
247, l.
Lilje, P., Yahil, A. and Jones B.J.T. (1986), Astrophys. J. 307, 9l.
Lucey, J.R., Currie, M.J. and Dickens, R.J. (1986), Mon. Not. R. astr. Soc. 222,427.
Lynden-Bell, D. (1971), Mon. Not. R. astr. Soc. 155,95.
Lynden-Bell, D. (1986), Q. Jl. R. astr. Soc. 27, 319.
Lynden-Bell, D. (1987), Q. JI. R. astr. Soc. 28, 186.
Lynden-Bell, D., Faber, S.M., Burstein, D., Davies, R.L., Dressler, A., Terlevich, R.J. and
Wegner, G. (1988), Astrophys. J. 326, 19. [SAMURAI]
Lynden-Bell, D., Lahav, o. and Burstein, D. (1989), Mon. Not. R. astr. Soc. 241, 325.
Mathewson, D., Ford, V. et al. (1990), address to Jerusalem Meeting on Large Scale Structure
of the Universe.
Melnick, J. and Moles, M. (1987), Rev. Mexicana Astron. Astrof. 14,72.
Miyaji, T. and Boldt, E. (1990), Astrophys. J. 353, L3.
Miyaji, T., Jahoda, K. and Boldt, E. (1991), in N. Holt and V. Trimble (eds.) After the
First Three Minutes.
Nilson, P. (1973), Uppsala General Catalogue of Galaxies, Uppsala astr. Obs. ann. 6.
352

Peebles, P.J.E. (1980), The Large Scale Strocture of the Universe, Princeton University
Press, Princeton.
Plionis, M. and Valdarnini, R. (1990), Mon. Not. R. astr. Soc. , in press.
Raychaudhury, S., Fabian, A.C., Edge, A.C., Jones, C. and Forman, W. (1991), Mon. Not.
R. astr. Soc. 248, 101.
Raychaudhury, S. (1989), Nature 342, 251.
Richter, O.-G., Tammann, G.A. and Huchtmeier, W.K. (1987), Astr. Astrophys. 171,33.
Rowan-Robinson, M. (1990), R.A.S. press release.
Rowan-Robinson, M., Lawrence, A., Saunders, W., Crawford, J., Ellis, R., Frenk, C.S.,
Parry, 1., Xiao Yang, X., Allington-Smith, J., Efstathiou, G. and Kaiser, N. (1990),
Mon. Not. R. astr. Soc. 247, 1.
Rubin, V., Ford, W.K. and Rubin, J.S. (1973), Astrophys. J. 183, L111.
Scaramella, R., Baiesi-Pillastrini, G., Chincarini, G. Vettolani, G. and Zamorani, G. (1989),
Nature 338, 562.
Shafer, R.A. and Fabian, A.C. (1983), in G.O. Abell and G. Chincarini (eds.) Early Evolution
of the Universe and Its Present Strocture, IAU Symp. 104, Reidel, Dordrecht.
Slipher, V.M. (1917), Proc. Am. Phil. Soc. 56,403.
Slipher, V.M. (1922), In Eddington A.S. Mathematical Theory of Relativity, Cambridge
University Press 1923, p. 162.
Smoot, G.F., Gorenstein, M.V. and Muller, R.A. (1977), Phys. Rev. Lett. 39, 898.
Strauss, M.A. and Davis, M. (1988), in G.V. Coyne and V.C. Rubin (eds.) Large Scale
Motions in the Universe, Proc. of Vatican Study Week, Princeton University Press,
Princeton.
Tammann, G. A. and Sandage, A.R. (1985), Astrophys. J. 230,648.
Tully, B. and Shaya, E.J. (1984), Astrophys. J. 281,31.
Vorontsov-Velyaminov, B.A. and Arkipova, A.A. (1963-1968), Morphological Catalogue of
Galaxies, Moscow State University, Moscow.
Yahil, A. (1988), in G.V. Coyne and V.C. Rubin (eds.) Large Scale Motions in the Universe,
Proc. of Vatican Study Week, Princeton University Press, Princeton.
Yahil, A., Walker, D. and Rowan-Robinson, M. (1986), Astrophys. J. 301, L1.
353

DISCUSSION:

Kaiser: In the QDOT IRAS redshift survey we do see a major concentration around
the locations of Hydra and Centaurus, so to a first approximation this agrees with the
predictions of the Great Attractor model. Moreover, not only does IRAS predict the local
Group motion, but there is also a correlation between IRAS predictions and measured
motions of other galaxies. When we look at the biggest residuals, one notable discrepancy
is actually in the GA region - we do not see the excess mass which gives the large and
positive peculiar velocities seen. One possibility is that there is something hiding in the
plane.
Lynden-Bell: Putting a significant mass behind the plane is quite possible by me because
some extrapolation of the Supergalaxy there is essential. However, it will increase the dipole
and thereby lower the dipole's nib!.
Sutherland: You said from the sky maps that the optical galaxies are more biased than
IRAS j isn't this partly due to their broader selection function, so if you made a deeper
optical map and then sparse-sampled it, that may look quite similar to IRAS?
Lynden-Bell: The optical maps go to a similar depth but have more objects. I would
have expected lower contrast in the optical due to dilution but I agree the IRAS luminosity
function is broader.
Dekel: (1) In regard to the comment by Kaiser, do you agree that it is inappropriate to
try and do a dynamical analysis based on the distribution of galaxies in a limited conical
volume such as Dressler's redshift survey?

(2) How can we understand the discrepancy in the position of the peak in the galaxy dis-
tribution in the vicinity of the GA as seen in the redshift surveys of Da-Costa vs. Dressler?

Lynden-Bell: (1) I agree but the results are still interesting.

(2) Da Costa, Melnick and Moles I believe went to fainter magnitudes than Dressler.
There may be some sampling problems but I don't yet understand the discrepancy.
TESTS OF INFLATION USING THE QDOT REDSHIFT SURVEY

CARLOS S. FRENK
Department of Physics,
University of Durham,
South Road,
Durham DHt 9LE.

ABSTRACT. The QDOT survey of 2163 redshifts of IRAS galaxies was designed to study clustering
on scales ~10h-l Mpc. It is ideal to test two fundamental predictions of generic models of inflation:
the value of the mean cosmic density and the nature of primordial density fluctuations. The survey
supports the predictions that n = 1 and that the primordial density field is a random Gaussian
process, but is inconsistent with a constant curvature power spectrum on scales :;;50h- 1 Mpc.

1 Introduction

The inflationary paradigm makes two important predictions which are testable: the value
of the mean cosmic density and the nature of the primordial field of density fluctuations.
Although it may be possible to formulate alternatives, the simplest model predicts a cos-
mological density parameter, n = 1, and primordial fluctuations of quantum origin with
random phases and a scale-invariant power spectrum (n = 1 in h~ ()( kn). The former is
established by the rapid expansion which flattens out any relic curvature, while the per-
turbation spectrum reflects the time invariance of the inflationary phase. In principle, the
parameter n can be determined through dynamical tests and the primordial fluctuation
field through studies of galaxy clustering on large scales.
A recently completed redshift survey of galaxies detected by the Infrared Astronomical
Satellite (IRAS) can be used to test inflation. The survey is known as the QDOT survey
after the initials of the participating institutions: QMWC (A. Lawrence and M. Rowan-
Robinson), Durham (J. R. Allington-Smith, R.S. Ellis, C.S. Frenk, I.R. Parry, and X.
Xiaoyang), Oxford (G. Efsthatiou and W. Saunders) and Toronto (N. Kaiser). (Guest
stars include J. Crawford, B. Moore, and D. Weinberg.) This survey is ideal because it
provides almost full-sky coverage by galaxies which are selected uniformly and are virtually
unaffected by inclination effects or galactic extinction. Since the survey was designed for
studies of scales larger that 10 h- 1 Mpc, whereas the characteristic clustering length
fV

of galaxies is 5 h- 1 Mpc, the optimal strategy was to randomly sample the galaxy
fV

distribution (Kaiser 1986). We adopted a sampling rate of one-in-six and a flux liInit at
60{Lm of 0.6 Jy. This gave a sample of 2163 galaxies at galactic latitude Ibl > 10°, covering
about 9 steradians, with a median velocity of 9000 km s-1 and a liIniting velocity of
fV

355
T. Shanks et al. (etis.). Observational Tests oj Cosmological Inflation. 355-364.
@ 1991 Kluwer Academic Publishers.
356

f'V30000 km s-1 . Of these, 1211 galaxies did not have measured redshifts and we obtained
those during 1986-88 using the INT, WHT and AAT. Our survey is approximately 3 times
deeper than and covers about 100 times the volume of the first CfA redshift survey.
The results reported here are part of a series of collaborations involving various subsets
of the authors mentioned above. The determination of n is discussed in detail in papers by
Rowan-Robinson et al. (1990) and Kaiser et al. (1991), while the large-scale density field
is discussed in papers by Efstathiou et al. (1990), Saunders et al. (1991) and Moore et al.
(1991). Some of these results employ the luminosity function of IRAS galaxies determined
by Saunders et al. (1990). A complementary redshift survey of IRAS galaxies is being
carried out by M. Davis, M. Strauss, A. YaWl and collaborators.

2 Morphology of the QnOT survey

An impression of the three-dimensional distribution of the QDOT survey is provided by the


isodensity contours shown in Figure 1. To make this plot we place each galaxy at its redshift
distance, weight it with the inverse of the selection function (computed from the Saunders
et al. (1990) luminosity function), smooth the resulting distribution with a Gaussian of
dispersion A, carve out a sphere of radius R around us, and plot those regions where the
smooth overdensity exceeds v times its 1mS dispersion. In Figure l(a), A = 15 h- 1 Mpc,
R = 100 h- 1 Mpc, and v = 0.7. This plot shows several well-known nearby Northern
structures such as the Local Supercluster (LS), the Abell 1367/Coma complex, and the
"Great Attractor" region (GA). In the South there are several less well-known superclusters,
some of which (e.g. S3, S6) are labelled according to the notation introduced by Saunders
et al. (1991). Figure l(b) shows the largest superclusters in the survey. Many are located
in a shell atf'V100 h- 1 Mpcand, as discussed in Section 4, dominate the number density
fluctuations in the survey. The parameters in Figure l(b) are A = 24 h- 1 Mpc, R = 150
h- 1 Mpc, and v = 1.3.

3 The value of n from the QnOT survey

The QDOT survey provides a map of the nearby universe of galaxies. Let us imagine
smoothing the galaxy distribution on some scale and define bI as the ratio of number
density to mass fluctuations on this scale, bI =(oNIN)/(oplp). Let us assume, further,
that gravity is the dominant interaction on the scales of interest. Then, at each point in
space, we can predict the peculiar acceleration generated by this mass. This prediction
depends on n and on the assumed connection between galaxies and mass. In linear theory
it is easily computed to be (Peebles 1980):

( ) _ nOo6HoJd3 ,ON( ,_ )(i'-r) (1)


v r - 41fbI r N r r (r' _ r)2'

(Note that bI is not the commonly used biasing parameter, b, defined as the ratio of 1mS
fluctuations in the galaxies and in the mass. The exact connection between bI and b is, in
fact, model dependent; see e.g. Frenk et al. 1990.) The integral in equation (1) is assumed
357

Figure 1. Isodensity contours of the smooth galaxy distribution in the QDOT sUnJey for
two different choices of parameters as discussed in the text. The obsenJer is located in the
centre. (Adapted from Moore et al. 1991.)

to be carried out over the whole of space. Thus, all-sky surveys like the QDOT survey,
offer the first opportunity to use equation (1) reliably. (Note that the predicted v(r) is
independent of H o .)
We now smooth the survey on small scales and use equation (1) to predict the radial
component of v(r) at every point. To estimate nO.6lbr we compare this prediction with
the'" 1000 independently measured peculiar velocities compiled and kindly supplied by
D. Burstein. (Details of this procedure, including a prescription to correct for redshift
space distortions, are given in Kaiser et al. 1991). The result is shown in Figure 2. The
correlation between predicted and observed velocities, although noisy, is remarkable: it
suggests that our assumptions concerning a connection between galaxies and mass and the
preeminence of gravity on large scales are not completely misguided. The best fit line to
these data gives nO.6lbr = 0.86 ± 0.16.
In a second application, we obtain a prediction for v( r) at the position of the Milky Way by
setting r = 0 in equation (1) and substituting the integral by n- 1 E4>(r')-1r'/r'2, where n
is the mean density of the sample and t( r) is the selection function of the survey. We com-
pare this prediction with the velocity inferred from the measured dipole in the microwave
background radiation (Lubin and Villela 1986). (See Rowan-Robinson et al. 1990 for de-
tails.) A very encouraging result, originally found by Strauss and Davis (1988), is that the
predicted vector points within only 200 of the dipole, suggesting that the mass traced by
the QDOT survey is responsible for the Milky Way's motion. This identification requires
nO.6lbr = 0.81 ± 0.15, very close to our previous estimate. Note that our two estimates are
358

1000

300 600 900

Figure 2. Predicted and measured peculiar velocities for Burstein's sample of tv 1000
galaxies. The linear sizes of the symbols are inversely proportional to the observational
errors. The lines are the best fit and the ±lu slopes. (Adapted from Kaiser et al. 1991.)

not redundant since they weight different parts of the survey in different ways; Burstein's
galaxies are all within tv 4000 km s-1 , whereas the predicted peculiar acceleration of the
Milky Way converges only at tv 10000 km s-l.
As with all dynamical tests, equation (1) constrains n only up to uncertainties intro-
duced by biasing. (The assumption "galaxies trace the mass" is simply a particular biasing
model.) Even with this uncertainty, our results strongly favour a density close to critical.
Indeed, to have n < 0.2, as suggested by optical studies in which galaxies are assumed to
trace the mass, would require bI < 0.46, implying that IRAS galaxies are much more weakly
clustered than the mass. Although some degree of antibiasing might plausibly be defended,
such a large effect seems ruled out. For example, if optical galaxies traced the mass and
n = 0.2, the autocorrelation function of IllS galaxies would be tv b} = 0.2 times that of
optical galaxies. Yet, as Figure 3 shows IllS galaxies are only slightly less clustered than
optically selected ones.
The straightforward interpretation of our result is that n = 1 and bI = 1.2 ± 0.2. Our
estimate of no.6 /bI is larger than estimates based on optically selected galaxies, partly
because IllS galaxies are somewhat less biased, and partly because our extensive sample
allows us to reliably bypass complications associated with non-linear processes. At the very
least, our results imply that for once it is the supporters of an open universe who must go
through some contorsions to defend their viewpoint.

4 Primordial fluctuations and the QDOT survey

The QDOT survey contains information on the amplitude and phases of primordial den-
sity fluctuations. The former is measured by the autocorrelation function discussed above
and, more reliably, by the counts-in-cells described here. The latter are constrained by
359

0.0
.2 -1

-2
e
-3LL~~~~~-LLL~~~~

2 2.5 3 13.5 4
log s/kms-
Figure 3. A utocorrelation function of galaxies in the Q DO T survey as a function of velocity
separation. The open circles give the estimate for the sample as a whole while the three lines
correspond to samples volume-limited at Sma", = 4000,6000 and 8000 km s-l respectively.
The dashed line is e(s) = (s/500 km s-1)-1.8. (Adapted from Moore et al. 1991.)

topological considerations.
Let us assume that the distribution of galaxies is a stationary point process. Then, the
variance of counts in cells of volume V is simply, (N - nV)2) = nV = N2V2(12 where n is
the mean density and (12 is given by a double volume integral of the autocorrelation function
(Peebles 1980). Assuming each galaxy is at its redshift distance, we have computed this
variance in two ways:
(i) Shell method: The survey is divided into radial shells and the moments of counts in cells
in each shell are computed. For any given cell size, the variance for the sample as a whole is
obtained using a maximum likelihood estimator. This method has two advantages: it does
not require an accurate knowledge of the selection function of the survey and it provides
some information on the radial dependence of the variance. The disadvantage is that it
assumes Gaussian statistics, and so the uncertainities are model dependent. The results
(Efstathiou et al. 1990) are plotted as solid squares in Figure 4.
(ii) Box method: We weight each galaxy by the inverse of the selection function, smooth the
distribution in a manner analogous to that depicted in Figure 1 and compute the moments of
counts in cells. This method has the advantage that it does not assume Gaussian statistics,
but it is sensitive to errors in the selection function. The results (Saunders et al. 1991) are
plotted as solid circles in Figure 4.
For reference, we also show in Figure 4 the variance expected in the standard cold
dark matter model (SCDMj Davis et al. 1985, White et al. 1987, Frenk et al. 1990),
normalised to the data points on the smallest scales we have considered (tV 7 h- 1 Mpc).
This normalisation requires b[ tV 1.2, consistent with the results of Section 3. The QDOT
data have a larger variance on large scales than predicted by the SCDM model. The
latter, in turn, is based on the inflationary prediction of a scale-invariant power spectrum
of Gaussian fluctua:tions, although it does not take on the asymptotic slope, n = 1, until
360

-.5
..l...
it"T
~.. -1 J"t
e
..J o "~
-1.5
1 "
"
-2

1.6
Leg r / h -'Wpc

Figure 4. Variance of counts in cells. Solid squares and solid circles give results for the
QDOT survey using the "shell" method (Efstathiou et al. 1990), and the "box" method
(Saunders et al. 1991) respectively. The vertical error bars are 1(1 errors; the horizontal
error bars reflect uncertainties in scaling between different cell shapes and smoothing func-
tions. The solid line gives the expected variance in a cold dark matter universe, normalized
to the observations at rv 7 h- 1 Mpc. The dashed line gives the slope of the power spectrum
inferred from a topological analysis. (Adapted from Moore et al. 1991.)

scales of order ~100 h- 1 Mpc. Thus, there is no evidence in the QDOT data for a scale-
invariant spectrum, although such a spectrum could, of course, be there on scales larger
than those probed by the survey.
A puzzling feature of our analysis is that most of the "excess" variance (that is, excess
over the SCDM prediction) seems to come from a shell between", 80 - 120 h- 1 Mpc. As
Figure 1( b) shows, the largest superclusters are located there, and are well distributed over
angle on the sky. A deeper survey should reveal how common this type of arrangement is.
The slope of the power spectrum (but not its amplitude) is constrained by the topology
of the galaxy distribution. This provides, in addition, a simple test of deviations from
a Gaussian distribution of fluctuations. The main ideas in this area stem from work by
Bardeen et al. (1986), Gott et al. (1986), and Hamilton et al. (1986) and have been applied
to the QDOT data by Moore et al. (1991). Briefly, one computes the genus of the smooth
galaxy distribution, a quantity related to the difference between the number of holes and
the number of isolated regions. This is calculated from data such as the contours of Figure
1, using the algorithm given by Weinberg (1988). Evidently, the genus can be computed
for different overdensity levels, v, and a plot of the genus per unit volume as a function of
v is called a genus curve. A very useful result is that, for Gaussian fields, the genus curve is
proportional to (1- v 2 )exp( _v 2 /2), and the proportionality constant gives the logarithmic
derivative of the power spectrum on the smoothing scale. In practice, difficulties arise in
real surveys which, like the QDOT survey, are affected by shot noise on large scales. These
difficulties can be overcome using N-body and Monte-Carlo simulations as described by
Moore et al.
361

The genus curve for the QDOT survey, smoothed on scales between 10 - 50 h- 1 Mpc, is
consistent with that expected from a Gaussian distribution. (This is often referred to as the
distribution having a "sponge-like topology.") Although certain non-Gaussian distributions
could have a genus curve of similar shape, this agreement lends support to the inflationary
prediction of quantum primordial fluctuations. Further evidence for gaussianity comes from
the fact that the amplitude of the genus curve implies a power spectrum slope similar to that
derived from the counts-in-cells approach. The topology analysis gives n ~ -1, roughly
constant in the region considered. Thus, the Gaussian nature of the large-scale fluctuations
inferred from the QDOT survey is consistent with the predictions of inflation, but the slope,
in the region considered, is not.

5 Conclusions

The QDOT survey of IRAS galaxies tests two of the fundamental consequences of generic
inflationary models, the value of the cosmic density and the nature of primordial density
fluctuations. The former can be estimated, in combination with the biasing parameter,
b[, from dynamical considerations. The result is n = 1 and bI = 1.2 ± 1.2. The evidence
in favour of a critical density is strong. The survey and the estimation procedures are
reliable. IRAS galaxies appear to be fairer tracers of the mass distribution than optically
selected galaxies or galaxy clusters. Furthermore, the survey is sufficiently extensive that
linear theory can be confidently applied, thus bypassing the uncertain non-linear physics
that shape the small scales.
The QDOT survey is consistent with a Gaussian distribution of fluctuations, on scales
"" 10 - 50 h- 1 Mpc, as expected from inflation, but the slope of the power spectrum on
these scales is n ~ -1 rather than the asymptotic value n = 1 of inflation or the slow
rolloverfrom n ~ -1 at 10 h- 1 Mpcto n ~ 0.5 at 50 h- 1 Mpcofthe inflationary cold dark
matter model. This does not rule out a constant curvature spectrum on larger scales.
The derived amplitude of mass fluctuations is greater than predicted by the standard
cold dark matter model. This result depends sensitively on the "linear biasing assumption"
that galaxy fluctuations are strictly proportional to density fluctuations. This assumption
is poorly justified and could well be wrong. For example, one might imagine effects, such as
the presence of a quasar, which could modulate galaxy properties on large scales and give
rise to superclustering of galaxies unrelated to superclustering of mass. Such effects might
show up as a dependence of the luminosity function of IRAS galaxies on environment.
Ambiguities of this sort will always plague attempts to measure mass fluctuations from
galaxy surveys. Fortunately there are other diagnostics, such as streaming motions or
anisotropies in the microwave background, which are not affected by this limitation. There
is every reason to hope that such tests will produce results in the near future.

Acknowledgements. I would like to thank my QDOT collaborators for allowing me to present


some results prior to publication.
362

References

Bardeen, l., M., Bond, l. R., Kaiser, N. and Szalay, A.S., 1986. A.troph1l'. J., 304, 15.
Davis, M., Efstathiou, G., Frenk,C.S., and White, S.D.M., 1985. A.troph1l'. J., 292, 37l.
Efstathiou, G., Kaiser, N., Saunders, W., Lawrence, A., Rowan-Robinson, M., Ellis, R., and Frenk,
C.S., 1990. Mon. Not. R. /Ultr. Soc, 241, lOp.
Gott, l.R., Melott, A.L. and DIckinson, M., 1986. A.troph1l'. J., 306, 34l.
Hamilton, A.l.S., Gott, l.R. and Weinberg, D., 1986. A.troph1l'. J., 309, l.
Kaiser, N., 1986. Mon. Not. R. /Ultr. Soc, 219, 785.
Kaiser, N., Efstathiou, G., Ellis, R., Frenk, C.S., Lawrence, A., Rowan-Robinson, M., and
Saunders, W. 1991 Mon. Not. R. astr. Soc., in press.
Frenk,C.S., White, S.D.M., Efstathiou, G. and Davis, M., 1990. A.troph1l'. J., 351, 10.
Lubin, P. and Villela, T., 1986. in Ga1az1l diltance. and deviation. form univer.al ezpamion, p.
Reidel, ed. B.F. Madore and R.B. Tully, p. 16l.
Moore, B., Frenk, C.S., Weinberg, D.H., Saunders, W., Lawrence, A., Rowan-Robinson,
M., Kaiser, N., Efstathiou, G., and Ellis, R., 1991, Durham preprint.
Peebles, J.P.E., 1980. The Large-Scale Structure of the Universe, Princeton University
Press.
Rowan-Robinson, M., Lawrence, A., Saunders, W., Crawford, l., Ellis, R., Frenk, C.S., Parry, I.,
Xiaoyang, X., Allington-Smith, l.R., Efstathiou, G., Kaiser, N., 1990. Mon. Not. R. alltr. Soc,
241, l.
Saunders, W., Frenk, C.S., Rowan-Robinson, M., Efstathiou, G., Lawrence, A., Kaiser, N., Ellis,
R., Crawford, l., Xiaoyang, X., and Parry, I., 1991. Nature, 349, 32.
Saunders, W., Rowan-Robinson, M., Lawrence, A., Efstathiou, G., Kaiser, N., Ellis, R., and Frenk,
C.S., 1990. Mon. Not. R. /Ultr. Soc, 242, 318.
Strauss, M. and Davis, M., 1988. in Comets to COlJmolog1l, proc 9rd [RAS conference, p. 00. A.
Lawrence, ed. Springer-Verlag, 56.
Weinberg, D.H., 1988. Pub. /Ultr. Soc. Pacific, 100, 1373.
White, S.D.M., Frenk,C.S., Davis, M., and Efstathiou, G., 1987. AstrophYIl. J., 315, 505.
363

DISCUSSION:

Blanchard: The correlation length of optical galaxies is of the order of 7 - 8h- 1 Mpc in
redshift space, so that a difference of 3 or 4 in the amplitude is possible.
Frenk: The value of the correlation length of optically selected galaxies is, as you imply,
controversial. There are difficulties of interpretation due to the use of redshifts rather than
directly measured distances, the applicability of the "fair sample hypothesis", etc. Some
of these difficulties carry through, of course, to IRAS galaxies. I would, however, argue
that when you use the same kind of estimator on similar samples of optical and IRAS
galaxies, you find little difference in their respective correlation lengths. There is a paper
by Marc Davis and collaborators (1988. Ap.J. letters, 333, L9.) which carried out an
extensive comparison along these lines. Our results are consitent with those and suggest
that it is difficult to have the large difference in the relative biasing between optical and
IRAS galaxies required if n = 0.2 and optical galaxies trace the mass. But I agree that the
current estimates need to be improved.

Dekel: In the determination of n from "topology", how do you decide what length scale
this n corresponds to? Is this necessarily the same as the wavelength (..\ = 21r / k) that you
refer to in the power spectrum?
Frenk: Under the assumption that the field is Gaussian, one can show that the "effective"
spectral index n (i.e the slope of a local power-law spectrum of index n) is determined on
the scale of the softening length. We used a Gaussian window function (without a factor of
2 in the exponent). To relate this to another power spectrum, one must take into account
possible differences in the window functions.

Fong: It is interesting to note that the correlation function at moderate scales reasonbly
follows the usual power law with a smaller ro than for galaxies. IT you also believe the
variance result for your cells", 40h- 1 Mpc, then doesn't this give a rather strange correlation
function, which will have to turn up somewhere between", 5h- 1 Mpc and 40h- 1 Mpc, in
order to give the large variance obtained by the counts-in-cells method.
Frenk: The correlation function and the counts-in-cells must clearly be consistent with
one another since they measure the same property. (The counts are just an integral over
the correlation function ~.) The counts-in-cells estimators we have used, however, are the
most sensitive to what is a rather small signal - '" 20% in the variance at '" 40h- 1 Mpc.
I do not think we can estimate ~ accurately enough at these separations to detect such as
small signal above the noise. In fact, even with our sparsely sampled survey, I do not think
we can say anything very precise about the behaviour of ~ beyond", lOh- 1 Mpc.
Sutherland: Can you really reject low n that strongly? n goes like the dipole to the
1.66, and the APM power spectrum suggests you could get another 25% on the dipole from
beyond", 100h- 1 Mpc and something, as Donald suggests, from within lOh- 1 Mpcj this
would leave n '" 0.3 quite consistent.
364

Frenk: I think I would not appear so confident if the two relatively independent mea-
surements of n from the QDOT survey that I discussed had not given the same result.
The predictions for the convergence of the dipole on the one hand and for the peculiar
velocities of external galaxies on the other, require weighting different parts of the survey
differently. This (and theoretical prejudice, I must admit) make me feel we have a strong
case for n = 1. But, as Blanchard's question also suggests, there are enough loopholes for
the non-believers to get through. But I most stress that, for once, it is the non-believers
who must do the fancy footwork.
TESTING INFLATION WITH PECULIAR VELOCITIES

AVISHAI DEKEL
Racah Institute of Physics
The Hebrew University of Jerusalem
and Meudon Observatory

ABSTRACT. I present three different results which indicate that the velocity
and mass-density fields extracted directly from the observed peculiar velocities
agree with the predictions of Inflation. (a) A comparison with the [RAS galaxy
density yields 0°·6 fbI = 1.35 ± 0.25, where bl is the biasing factor for [RAS galaxies,
and non-linear analysis favors 0 ~ bl ~ 1. (b) The shape of the fluctuation
power spectrum in the wavelength range 10 - 100 h-1Mpc is consistent with an
initially scale-invariant spectrum modified by cold dark matter (CDM). (c) The
one-point probability distribution of the velocity fields by POTENT and by [RAS are
consistent with Gaussian initial fluctuations. I conclude that there is no dynamical
evidence against the standard model of the formation of large-scale structure; the
indicated large-scale excess of power in the galaxy distribution could be attributed
to non-trivial biasing of galaxies relative to the mass distribution.

1. Introduction
The 'standard' Inflation model predicts that the present universe is fiat with
o= 1 (A = 0). Energy-density fluctuations originated from quantum fluctuations
and were stretched to large comoving scales. The assertion is that these fluctuations
were initially a Gaussian random field with a scale-invariant power spectrum. The
fluctuations grew to the present structure via gravitational instability, and the
nature of the dark matter (DM) determines the filtering of the spectrum during
the radiation/plasma era. CDM is currently the favorite DM candidate.
The most direct evidence for the distribution of matter is dynamical, coming
from peculiar velocities. It is safer to assume that the galaxies move as test bodies
in a large-scale gravitational field, independently of how they were selected, then to
assume that they are honest tracers of the mass distribution. Given the observed
line-of-sight peculiar velocities of a sample of ('" 1000) galaxies[l], the POTENT
procedure[21 first smooths them into a radial velocity field, v r ( r'), in a way that
minimizes the effects of sparse sampling and measurement errors. It then imposes
the natural prediction of gravitational instability that the velocity is a potential
v
flow, = -V4>. The velocity potential at each point is calculated by integrating
the radial velocity along radial rays,

(1)

and the two missing components of the velocity are recovered by differentiating this
potential in the transverse directions.
365
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 365-373.
@ 1991 Kluwer Academic Publishers.
366

The underlying mass-density fluctuation field, c( rj, is approximated by[3]

(2)

where the double vertical bars denote the Jacobian determinant, I is the unit
matrix, and 1(0.) = d(lnc)/d(lna) ~ 0.°. 6 with a the universal expansion factor.
This non-linear approximation, which is the exact solution of the continuity
equation under the Zel'dovich assumption that particle displacements grow in a
universal rate [4] , has been tested successfully using cosmological N-body simulations
smoothed with a Gaussian of radius as small as 500 kms- 1 , with an rm3 error less
than 0.1 over the range -0.8 ::; C ::; 4.5. It reduces to the linear approximation,
Co = _(HJ)-l\7. V, when Ici ~ 1.
The uncertainty due to the random distance errors of the data objects is
estimated at each point using the rm3 scatter, Ufj, in Monte Carlo noise simulations.
Further analysis is limited to regions where Ufj is smaller than some value. Malmquist
bias has been estimated to be small compared to the random errors. The bias due
to non-uniform sampling is controlled by limiting the analysis to regions where R4 ,
the distance to the 4th nearest object, is smaller than a certain value.

2. Constraints on 0. by Comparison to IRAS Galaxy Density


c
The POTENT density, p , is 0. dependent, proportional to 1(0.)-1 in the linear
c
regime. The galaxy-density extracted from the [RAS catalog[5], r , is only slightly
affected by 0. and the specific relation between galaxies and mass - only through
the predicted peculiar velocities used to correct the redshifts to real positions. If
we parameterize the relation between galaxy and mass density fluctuations by a
universal 'biasing' factor[6], cr = brc, then we expect in the linear regime a linear
relation of the sort cp = [J(n)/br]cr . Given the uncertainties in the two datasets,
we can ask whether the [RAS data are consistent with being a noisy version of the
POTENT data, or vice versa, and obtain, via linear regression, the best-fit value for
I(n)/br with associated confidence limits[7].
The degeneracy of 0. and br is broken in non-linear regions, where c(v) is no
longer simply proportional to 1(0.)-1. The local volume where the comparison is
possible does contain regions where Icl '" 1 even after 1200 km S-l smoothing.
To include non-linear effects, as in POTENT, we apply the following non-linear
modification to the inverse analysis that predicts velocities from [RAS. A Before
v
integrating the linear relation Co ex: - \7 . forv, we obtain Co from Cr by applying
an empirical relation deduced from N-body simulations[3]: Co = C- 0.2c 2 + 0.OM3.
The deviation about it is less than ±O.l. These compatible quasi-linear treatments
of POTENT and [RAS allow a first attempt at determining 0. and br separately, but
still assuming the simplest biasing model bI = const.
A similar comparison could be made at the level of velocities, but it is
complicated by an unknown quadrupole moment[8]. I focus here on the comparison
of densities, which is local, independent of reference frame, and can be more easily
corrected for non-linear effects.
Figure 1 compares density maps in the Supergalactic plane. Gaussian
smoothing of radius 1200 km s-l has been applied and 0. = br = 1 has been assumed
for this plot. Despite our efforts to minimize the effects of non-uniform sampling,
it does introduce a non-negligible bias into the results of the POTENT smoothing
367

procedure. Instead of comparing the POTENT density (P) to the raw IRAS density
(1), we should compare it to an IRAS density field biased in a similar way: the
IRAS velocities (with 500 kms- 1 smoothing) were sampled non-uniformly as in the
velocity sample and then passed through the POTENT machinery with 1200 kms- l
smoothing [pel)]. To be on the conservative side, we limit the comparison to
the volume where the POTENT noise and emptiness are below 0'6 = 0.2 and
~ = 1500 kms- l . The effective volume for the comparison is '" (5300 kms- I )3.

Figure 1: Density fluctuation fields in the Supergalactic Plane (LG at the origin):
P is mass density by POTENT, 1 is galaxy density by IRAS, and pel) is IRAS
velocities sampled non-uniformly and analyzed by POTENT. Contour spacing is 0.1
(positive solid, negative dashed, and zero thicker). The heavy line is 0'6 = 0.2.

Both maps feature the general "Great Attractor" (GA) phenomenon as a ramp
which peaks beyond the Hydra-Centaurus clusters at about X ~ -4000, Y ~ 1000
and falls off gradually toward Virgo (X ~ -300, Y ~ 1300). The GA peak in the
IRAS map coincides with the original prediction of the Seven Samurai [1] for the
center of the velocity flow (the small cross). The two fields also show an adjacent
"great void" in the region X < 2000, Y < o. There is an apparent discrepancy in
the Perseus region (X > 3000, -2000 < Y < 3000), where the IRAS map shows
a density enhancement while the POTENT map indicates an underdensity, but this
region is both too noisy and undersampled in the current POTENT data. In fact,
POTENT does recover a significant overdensity near Pisces (X ~ 3000, Y ~ -3000);
a preliminary POTENT analysis of new data by J. Willick recovers a significant
density peak on the order of 8 = 1 there.
The POTENT and IRAS densities are plotted against each other at grid points of
spacing 500 kms- l in Figure 2. Our model is a linear relation: 8p = A81 +s. (A shift
s is possible because 81 is normalized by the mean IRAS density within its survey
volume, while the input velocities to POTENT are calculated relative to a mean
Hubble expansion in a different volume.) The parameters A and s are obtained
by linear regression of 8p on 81 , with the weights 0'-;2. The lines drawn are the
best fits: the naive comparison of P and 1 yields A = 0.81, and the bias-corrected
comparison of P and P(I) yields A = 1.15. In ~eneral, there is a strong correlation
between the POTENT and IRAS fields, with f(n)/b 1 of order unity, and with a small
systematic variation of mass-to-light ratio in space. The bias does not change the
growth features - only the details. The exact value of f(n)/b 1 determined this way
should not be regarded as a fully self-consistent estimate yet because these density
368

fields were calculated assuming 0 = br = 1.

.,
.. . ·1

.. ....
0.5 \-----j--.-.---.-,...../'+.'....r7-'---j 0.5 \-----h-----;~~¥-\------l

OI----."!-o1~ 01----""

-0.5 '--'--........<::.....JL.........JL.........JL.........JL.........JL.........JL.........JL.........JL.........JL.........JL.........J---' -0.5 L.........JL.........J'-'---'L.........JL.........JL.........J---I---I---L---L---L---L---L---L----'


o
-0.5
o [I]
0.5 -0.5
° 0 [P(I)]
0.5

Figure 2: POTENT vs. [RAS density fluctuations at grid points (0"6 < 0.2).
To properly correlate the non-linear POTENT and [RAS fields and determine 0
and br in a bias-independent way, given the measurement errors, we have pursued
a likelihood analysis using Monte Carlo simulations, spanning the parameter plane
O-br • I show in Figure 3 preliminary likelihood contours from this ongoing project.
Our robust result, quite insensitive to nonlinear effects, is that 0°·6 fbr ,. . ., 1. As long
as br is not much smaller then unity, I can quote 0°·6 fbr = 1.35±0.25. The nonlinear
effects seem to introduce lower limits on 0 and br , but the results for low br should
be considered uncertain because they involve extreme nonlinear corrections which
are still under investigation. Another limitation of the current nonlinear analysis is
that it assumes that br is independent of position.

2
90%

1.5

Figure 3: Likelihood contours (preliminary). The dashed line is f(O)fb r = 1.36.


369

3. The Fluctuation Power Spectrum


What does the observed velocity field tell us about the power spectrum of the
mass-density fluctuations? Is it consistent with a scale-invariant tail of slope n = 1
on scales larger than 100 h- 1 Mpc and gradual flattening corresponding to 'standard'
CDM on smaller scales?
Figure 4a shows the bulk flow of the POTENT velocity field smoothed by a
Gaussian of 1200 km s-l, inside top-hat spheres of radius R about the Local Group.
For example, the bulk flow in a sphere of radius 4000 km S-1 is V = 388 ± 67 km S-I.
This is fairly consistent with the rm3 prediction for CDM, (V 2 )1/2 = 287 b- 1 kms- 1
(based on the standard normalization of cM/M = b- 1 in spheres of 800 kms- 1 ),
even if b ~ 2.
1000

100

BOO POTENT DATA It.. 1200 km./l <,l>l/1

10

.e SOO

.
"- :;<
0::
'00
v
.1

200

.01

o0L..L..J--1....Lj ...1. 0 ..L..l...-'::.20LOOLLL-J....J300LlO--1....L...1.,...L00-'-0-'::.L.L..SOLOOL...L...J--'S....l00...1.0...1....1...L...J


00..J.. 7000
R (kmj,)

Figure 4: Statistics of the POTENT velocity field, smoothed 1200 kms- 1 .


Error bars mark the standard deviation in Monte Carlo noise simulations. (a)
Bulk velocity V, and the dispersion about it S, in spheres of radius R. (b) The
power spectrum of the mass-density fluctuations derived inside a sphere of radius
6000 kms- 1 . The solid line is the rm3 CDM spectrum (0 = 1,h = 0.5,b = 1),
Gaussian smoothed on small scales as in POTENT.
An analysis of the "Mach number" of this velocity field, defined by the ratio of
bulk velocity to the standard deviation from it, gives M = 1.2 ± 0.2 in a sphere of
radius 4000 kms- 1 . Based on Monte Carlo simulations[9}, this is consistent at the
10" level with the rm3 value predicted for CDM, again indicating a general agreement
of the shape of the power spectrum with the 'standard' model.
Finally, the power spectrum, Pk, has been determined[lO} by FFT of the
POTENT potential field, which has been artificially and gradually damped to zero
at radii 6000 - 8000 kms- 1 to make periodic boundary conditions at the edges of a
cube of side 16,000 km s-l. The resulting power spectrum for the density is shown
in Figure 4b. Its shape agrees nicely with the shape predicted for COM out to
wavelength", 100 h- 1 Mpc, where the logarithmic slope of Pk bends to n ? O. This
is consistent with a Zel'dovich scale-invariant spectrum (n = 1) on larger scales, but
we cannot explore such large scales yet in terms of the dynamical fields.
370

4. Evidence for Gaussian Initial Fluctuations


In the 'standard' model, the fluctuations are assumed to be a random field,
6( i), i.e. a set of random variables, one for each point in space, which is fully
specified by the m-point joint probability distribution functions, P m ( 6)[11]. The
m ~ 2 functions involve the spatial correlation functions, or the power spectrum,
whose time evolution is affected by the nature of the DM. But PI (6) is not explicitly
sensitive to it, so present observations can directly reflect the initial one-point
distribution without speculating about the dark matter.
The natural choice is a Gaussian random field, where Pl (6) DC exp( _6 2 /2( 2 )
(and the joint probabilities are multi-dimensional Gaussians). But non-Gaussian
initial fluctuations have been considered in recent years as a possible explanation
for the indicated structure on very large scales in excess of the predictions of
the "standard" CDM model[12]. In fact, the general inflation picture permits
non-Gaussian fluctuations[13]. They also arise in scenarios where the fluctuations
originate from topological defects such as cosmic strings[14] and textures[15], or from
explosions[16]. The statistical nature of the fluctuations is therefore an important
distinguishing feature between the competing theories.
The distribution functions do not change during linear evolution but the recent
non-linear evolution of structure could introduce strong non-Gaussian features which
might blur any initial non-Gaussianities. The non-linear evolution of an initially
Gaussian density distribution can be followed using the Zel'dovich approximation
or N-body simulations, but its present deviation from Gaussianity depends on the
amplitude of the fluctuations which is ill determined due to the unknown "biasing"
of galaxies.
We find[17]' however, that the probability distribution of velocities (scaled
properly in time) is rather insensitive to non-linear evolution. This can be shown
using the Zel'dovich formalism, which properly approximates a cosmological system
as long as there is no orbit mixing, or when smoothed over a large-enough scale.
A more elaborate argumentation based on kinetic theory shows that this property
of P( v) is valid for the smoothed Zel'dovich velocity field even in the presence of
multi-streams in collapsed pancakes (in preparation). This means that the observed,
smoothed velocity field could almost transparently reflect the statistical properties
of the initial fluctuations.
To test the non-linear effects, I show in Figure 5a the distributions PI (v) and
PI (6) from a cosmological N-body simulation of 'standard' CDM (with 1283 grid cells
and 1283 particles in a cubic periodic box of comoving size 160 h- l Mpc, normalized
such that (6 2 ) = 1 in spheres ofradius 800 kms- l today). The fields are smoothed
with a Gaussian of radius 500 km S-I. The dimensionless variable x is either v or g,
with a zero mean and a standard deviation of unity. The velocity components along
the three axes are regarded as one random variable with 3N entries. The figure
demonstrates that while PI (g) develops a characteristic non-Gaussianity, PI (v) is
hardly affected by non-linear effects.
The distributions of the observed smoothed fields from POTENT and IRAS are
shown in Figures 5b and 5c. In POTENT we use only points with Monte Carlo errors
in the velocities and densities smaller than 400 km S-1 and 0.4 respectively. The
resultant volume corresponds to a sphere of radius 2300 kms- l . In IRAS we use
only points with bootstrap errors smaller than 200 kms- l and 0.2 respectively. The
resultant volume corresponds to a radius 6000 km s-l. The finite volume affects the
distribution mostly via the random .jN effect. Some of the small deviations from
371

a pure Gaussian distribution can be due to this noise, especially in the POTENT
output that is limited to a rather small volume.

-2

p..
.=-4
-6

-10 02 10 -10 02 10 -10 02


X
10
X X

Figure 5: The point probability distribution functions compared with a Gaussian


(dashed) and with the Zel'dovich approximation (dashed-dotted). The random
variable x is v (solid) or 8 (dotted). To stress the behavior at the tails of the
distribution, InPl is plotted against x 2 (the minus signs refer to x, not x 2 ). The
distributions are based on N-body simulation (left), peculiar velocities via POTENT
(middle),. and IRAS galaxy distribution (right).
Given the errors, the observed PI (v) are indistinguishable from the
corresponding N-body distributions and from a Gaussian distribution. The observed
density distributions are also consistent with the N-body simulations of an initially
Gaussian field. Based on these results we can probably rule out non-Gaussian
velocity distributions such as predicted by the Cosmic Strings scenario, where the
velocity field in each rv 40 h- l Mpc region is typically dominated by a single wake.
But it remains to be seen whether other non-Gaussian models, such as Textures
and Explosions, do not produce nearly Gaussian velocity distributions despite the
non-Gaussian nature of their density distribution simply as a result of the central
limit theorem when summing over several weakly-correlated sources of gravity. Note
also that our data, especially in POTENT, is noisy, it is heavily smoothed and it is
limited to a finite volume. Therefore, the distribution functions carry non-negligible
random errors until the sanIples are extended and improved.
5. Conclusion
I presented three results concerning different predictions of the 'standard'
inflation model:
• Cosmological parameters. The I RAS galaxy distribution is correlated with the
mass fluctuations. A robust linear comparison yields bI/flo. 6 ~ 1. A preliminary
comparison with optical data yields bo /flo. 6 ~ 2. These values are consistent with
fl = 1 and typical b values of order unity. An open model, with fl ~ 0.1 say, is
also acceptable, but only if the IRAS galaxies are severely "antibiased": bl :::; 0.25.
Our preliminary attempt to determine fl and bl independently, using non-linear
corrections, seems to favor the standard option: fl ~ bl ~ 1, but this result is still
highly uncertain. These results are insensitive to the value of A.
• Dynamical power spectrum. The bulk flow in a top-hat sphere of radius
4000 km s -1 is V = 388 ± 67 km s-l. Contrary to previous belief, the bulk flow
372

is fairly consistent with the rm" prediction for CDM (287b- l km s-I), even if the
normalization is b ~ 2. An analysis of the Mach number of the smoothed POTENT
velocity field gives values of order unity, which, again, are consistent at the 10"
level with the rm" value predicted for CDM. A direct recovery of the fluctuation
power spectrum, PIc, from the POTENT output reveals a detailed agreement of
shape with the power spectrum of CDM out to wavelength", 100 h- l Mpc, where
the 10garithInic slope have gradually bent to n ~ O. This is consistent with a
Harrison-Zel'dovich scale-invariant spectrum (n = 1) on larger scales, but these
scales cannot be explored dynaInically yet .
• Probability di"tribution. We found that the one-point probability distribution
function of the smoothed peculiar velocity field, PI (v), remains fairly constant
even through non-linear evolution. The Pl(V) of the observed (lD) velocities,
as processed through POTENT or as predicted by the IRAS analysis, is found
to be indistinguishable from a normal distribution, suggesting that the initial
fluctuations were Gaussian, and ruling out extreme non-Gaussian models such as
certain versions of the CosInic Strings scenario where large wakes doIninate the
formation of large-scale structure.
These results lead me to a rather conservative (perhaps too conservative to
my taste) conclusion: the observed velocities and the corresponding mass-density
field are consistent with the 'standard' model: a homogeneous universe with an
inflation phase, n = 1 and A = 0, Gaussian, adiabatic initial fluctuations with a
scale-invariant spectrum, cold dark matter with '" 10% baryons and gravitational
growth of fluctuations to the present structure. In this case, the indications[12] for an
excess of power in the galaxy distribution on scales ~ 40 h- l Mpc must be accounted
for by a non-trivial relation between galaxies and mass, where the biasing factor is
both a function of scale and density. But I would not say that the alternative,
an n '" 0.1 baryonic universe [19] , is ruled out with convincin~ confidence; just
that it requires isocurvature fluctuations and reionization of ~TIT plus extremely
non-trivial anti-biasing on scales of '" 10 h- l Mpc.
Acknowledgements. This talk describes work done in collaboration with Ed
Bertschinger, and in parts with M. Davis, J. Gelb, Y. Hoffman, L. Kofman, T.
Kolatt, A. Nusser, M. Strauss and A. Yahil. The main velocity data is due to D.
Burstein, A. Dressler and S. Faber. This research has been supported by US-Israel
Binational Science Foundation grant 86-00190 and 89-00194.
References
1. Lynden Bell, D., Faber, S.M., Burstein, D., Davies, R.L., Dressler, A., Terlevich,
R. J., and Wegner, G. 1988, Astrophys. J. 326, 19.
2. Bertschinger, E. and Dekel, A. 1989, Astrophys. J. (Lett.) 336, L5; Dekel, A.,
Bertschinger, E. and Faber, S.M. 1990, Astrophys. J. 364, 349; Bertschinger,
E., Dekel, A., Faber, S.M., Dressler, A. and Burstein, D. 1990, Astrophys. J.
364,370.
3. Nusser, A., Dekel, A., Bertschinger, E. and Blumenthal, G.R., 1991, Astrophys.
J. , in press.
4. Zel'dovich, Ya.B. 1970, Astron. Astrophys. 5, 20.
5. Strauss, M.A., Davis, M., Yahil, A. and Huchra, J.P. 1990, Astrophys. J., in
press; Yahil, A., Strauss, M. A., Davis, M. and Huchra, J. 1990, Astrophys. J.
, in press.
6. e.g. Dekel, A. and Rees, M.J. 1987, Nature 326, 455.
7. Dekel, A., Bertschinger, E., Yahil, A., Strauss, M. and Davis, M. 1991, Astrophys.
J. , submitted.
373

DISCUSSION:

White: Can you really say anything about the power spectrum on scales of ...., 10,000
km/s based on a data sample of diameter...., 10,000 km/s? It seems as though one would
need data for a number of such regions in order to estimate an rms power.

Dekel: I agree. Using the potential in a given volume we do recover the dominant waves
with wavelengths larger than the diameter of the volume, but, indeed, this is only one
realization of a random field and it is not a fair estimate of the power-spectrum, which
should be determined as an rms value over an ensemble of such volumes. Still, given only
one universe around us, this is the best estimate we have on such scales.
Hoffman: How sensitive is your probability distribution function of the velocities to the
presumably Gaussian error introduced by the POTENT analysis?
Dekel: We used only velocities in points where the errors, as estimated by our Monte
Carlo noise simulations, are small. Therefore, the distribution beyond 10' and in particular
in the tail, should be mostly signal - not noise. The non-Gaussian distribution of the
densities, where the relative errors are even larger, confirms that.
THE INVISIBLE COSMOLOGICAL CONSTANT

O. LAHAV1 , P.B. LILJE 2 , J.R. PRlMACK3 and M.J. REESl


1 Institute of Astronomy, Madingley Road, Cambridge CB3 OHA, UK.
2 NORDITA, Blegdamsvej 17, DK-2100 Copenhagen, Denmark.
3 Physics Department, University of California, Santa Cruz, CA 95064, USA.

ABSTRACT. We discuss the possibility of measuring the density parameter no and the
cosmological constant Ao == AJ(3HJ) from dynamical tests using linear and non-linear spher-
ical models. While information on both parameters can be obtained by looking at clustering
at different redshifts, in practice the other observables also depend on the cosmology, and
in some cases conspire to give a weak dependence on Ao. We also show that a repulsive A
gives a smaller final radius of a virialized cluster than a vanishing A.

1. Fundamental Limits on The Cosmological Constant

The cosmological constant A has had a checkered past since it was introduced by Einstein,
but is becoming popular in cosmology again in order to help resolve a number of problems:
(1) The age of the Universe; (2) Inflation implies vanishing curvature, i.e. no + AO = 1,
so if no R: 0.2, as suggested by some observations, then AO R: 0.8; (3) Number counts of
galaxies at high redshifts (A can increase the spatial volume); (4) The 'standard' Cold Dark
Matter (CDM) model apparently predicts substantially less large scale structure than obser-
vations indicate. A lower no R: 0.2 gives more large scale structure, but this is incompatible
with observational limits on cosmic background radiation fluctuations for A = 0, but not
incompatible if Ao = 1- no. Observationally, JAJ ~ Peril, which is 120 orders of magnitudes
smaller than the 'natural' value in terms of the Planck mass (e.g. Weinberg 1989). It is
tempting therefore to try to find a physical reason why A = O. Here we discuss what can be
learned about A from the recent redshift and peculiar velocity surveys. More details on the
calculations presented here are given in Lahav, Lilje, Primack & Rees (1991).
Before considering the dynamical tests, we review here the limits on A from basic consid-
erations. The "energy" Friedmann equation for the scale factor a = (1 + z)-l is

(1)

where Pb is the background density (which varies like a- 3 in a matter-dominated universe),


t
and J( = HJ (no - 1) + is the curvature, where Ho is the Hubble constant. In Figure 1
we show the (AO - no) parameter space. We see that the permitted range is fairly small, but
still allows values different from the popular point (no = 1, Ao = 0).

375
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 375-378.
© 1991 Kluwer Academic Publishers.
376

K<O K=O K>()


5 - " -
\,
, "
4
'\
BIG \ , CRUNCH
3

-4

Figure 1. The phase-space of no and ~o = Aj{3HJ) with various fundamental constraints.


The dashed-dotted line indicates an inftationary (i.e. fiat) universe. The solid lines show 4
values for the age of the universe Hoto, and the dashed line is the constraint of Gott et al.
(1989) from a normally lensed quasar at z = 3.27. The boundary (~.) of the shaded "No
Big Bang" region corresponds to a coasting phase in the past.

2. Dynamics in Linear Theory

The solution to the standard linearized equation for the growth of perturbations fJ = fJpj p
as a function of the scale factor a in a A :f. 0 universe was given by Heath (1977). In linear
theory the peculiar velocity vpec is related to the peculiar acceleration g by (Peebles 1980,
§14)
2fg
vpec = 3Hn' (2)

where f =::~! . This relation is basically derived from the linearized continuity equation,
with the assumption that the growth rate of structure is independent of position. Another
variant of this equation is the radial infall velocity around a spherical perturbation of radius
R and with interior average over-density (fJ)R, 1m-
= V(fJ)R.
377

Here we generalize the calculation for any values of no and Ao at any redshift z. We find
the following approximation for I(no, Ao) at z = 0 :

06
I(z = 0) ~ no' + 1
70 Ao ( 1 + 2no
1 )
. (3)

This generalizes the result of Peebles (1984), who only considered the case no + Ao = 1. We
see that at z = 0, I is almost entirely determined by no, with a very weak dependence on
Ao. Figure 2 shows the evolution of I with redshift for different pairs of (no, Ao). We see
that in the range z ~ 0.5 - 2.0 it is possible to discriminate between universes with different
cosmological parameters. The high peak for the K > 0 cases is due to a 'coasting phase'.
These curves are well fitted by I(z) ~ nO.6 , where n is a function of z, no and Ao. The
interpretation of the dynamical tests crucially depends on how the observables are actually
deduced. For example, since I ~ nO. 6 , the product I H depends only weakly on >'0.

,
3

,I
\
2.5 ,I
\
,I
l\ I \,
I \' \
2
I !,
I ,I \
"-

__
(0., ~)
... 1.5 I ,I '.
J........ ---
//"-,r- -----~.~ ___ .__ --- ~-=
(0.1. 1.3)
(1.0,2.0)
u - _. _ _ . _ _ _ _ . _ _ • (1.0, 1.0)
(1.0.0.0)
.~---!---------------------

. . . . . . . . . . . . ..
(0.1. 0.9)
,I _----- - - - - - -.-----
................. (0.1,0.5)
(0.1,0.0)
.5 l~:.~· -::.~.~.~.""

2 4 6 8 10 12 14
l+z

Figure 2. The evolution of the growth factor of the perturbations 1== d(1n c5)jd(ln a) with
red shift for different pairs of (no, Ao).

3. Non-Linear Spherical Models

After showing that dynamical tests for AD in linear theory are only useful at high redshifts,
we explore the possibility of using non-linear theory at z = 0 and at higher redshifts. We
have generalized for A # 0 the procedures of Lilje & Lahav (1991) to estimate the average
density and velocity fields around clusters of galaxies. We considered the "standard" biased
CDM model with no = 1, Ao = 0, h = 0.5 and biasing parameter b = 2, and three other
378

low-density CDM models with no = 0.2, h = 1, and b = 1, for several values of AO (0.0, 0.8
and 1.3). The noh = 0.2 CDM power-spectrum provides a phenomenological fit to the APM
angular correlation function (Efstathiou et al. 1990).
At z = 0 we find that the profiles strongly depend on no, but changing Ao has almost no
effect. This is in accord with N-body simulations (e.g. Martel 1991). We then looked at the
profiles at higher redshift, z = 0.5, as a function of angular separation. Unfortunately the
combination ofthe slope ofthe density profile (close to R-2), the angle-redshift relation, and
the growth factor J conspire to put the zero-velocity surfaces at almost the same angular
distance for all low-density models.
In a Universe with A = 0 a well known result is that the 'final' effective radius of a
collapsing cloud is half its initial radius. In the case when the A-term is included we find by
applying the virial theorem:
RJ (1 - 'T//2)
(4)
Ri ~ (2 - 'T// 2) ,
where 'T/ == A/( 41rGPta) specifies the A-density relative to the turn-around density. The
condition for a shell to turn around is 'T/ < 1. We see that for a repulsive (positive) A the
final radius is smaller compared with a zero-A case.

4. Discussion

We found that present-day observations of density and velocity fields cannot be used to
distinguish between models with different Ao. At z = 0 the independent estimates for f
from the optical and IRAS dipoles and the comparison of the IRAS density field with a
reconstructed matter density (see discussions in these Proceedings) yield J ~ 004 - 1.2
(allowing for biasing in the range b ~ 1 - 2). This corresponds to no ~ 0.2 - lA, almost
for any value of Ao. The present observations of the combination fib still allow different
scenarios which satisfy both inflation and Big Bang nucleosynthesis. By going to higher
redshifts it is possible in principle to decouple the measurement of no from Ao. For example,
if the correlation function of quasars is almost constant in the redshift range z = 0 - 2 (Boyle
1991), then a scenario with J(z) ~ 0, i.e. with a low present-day no and a vanishing Ao,
is more favourable (see Fig 2). Other possible probes of A are clusters at z ~ 0.5 - 1.0,
absorption line systems and gravitational lensing.

References
Boyle, B., 1991. In The Texas/ESO-CERN Sypm., Brighton UK, 1990.
Efstathiou, G., Sutherland, W.J. & Maddox, S.J., 1990. Nature, 348,705.
Gott, J.R., Park, M.G. & Lee, H.M., 1989. Astrophys. J., 338, 1.
Heath, D.J., 1977. Mon. Not. R. astr. Soc., 179, 351.
Labav, 0., Lilje, P.B., Primack, J.R., & Rees, M.J., 1991. Mon. Not. R. astr. Soc.,
submitted.
Lilje, P.B. & Labav, 0., 1991. Astrophys. J., in press.
Martel, H., 1991. Astrophys. J., 366, 353.
Peebles, P.J.E., 1980. The Large Scale Structure of the Universe, Princeton University Press,
Princeton.
Peebles, P.J.E., 1984. Astrophys. J., 284, 439.
Weinberg, S., 1989. Rev. Mod. Phys., 61,1.
SUPPORT FOR INFLATION FROM THE GREAT ATTRACTOR

ALAN HEAVENS
Department of Astronomy
University of Edinburgh
Royal Observatory
Edinburgh EH9 3HI

ABSTRACT. The depth of the velocity potential at the Great Attractor is used to constrain the
density parameter, the spectrum of fluctuations, and the degree of bias ofIRAS galaxies, using only
the following assumptions:
• On large scales, galaxies trace the velocity field (of any dark matter).
• The velocity perturbations are gravitational in origin.
• The perturbations are Gaussian on large scales.
No further assumptions are required. From the absence of fluctuations in the microwave back-
ground radiation, adiabatic CDM models must have 0 > 0.5. From the distribution of IRAS
galaxies, their bias parameter on large scales must satisfy 0/b 5 / 3 > 0.5. For spectra which have
more large-scale power, the lower limits on 0/b 5 / 3 are relaxed, but the ma..'Cimum likelihood values
for power-law spectra exceed 0.6.

1 Introduction

Peculiar velocities of galaxies can be used as a very powerful tool in cosmology; if their
source is gravitational, they are determined by the distribution of mass, which is otherwise
difficult to determine. Comparison of the velocity field with a sample of luminous objects
can therefore be used to constrain the biasing ofthose objects and the density parameter, n.
This sort of comparison has been done in two ways previously: the distribution of luminous
objects has been used to predict the flow locally (the dipole method; e.g. Strauss & Davis
1988); and Poisson's equation has been employed to calculate the mass distribution from
the flow (e.g. Dekel, this volume). Each method has its advantages and disadvantages; the
latter involves considerable work in measuring large numbers of peculiar velocities, but it
does not suffer from potential difficulties in the dipole method arising from incomplete sky
coverage, uncertain luminosity function and sources outside the survey. The disadvantage
of the Poisson method is that it requires a double differentiation of the velocity potential
field, which is done numerically and may be rather noisy. Both these methods, however,
give estimates of n/b5 / 3 which are independent of the primordial fluctuation spectrum.
This report describes an alternative method which uses the properties of the peculiar
velocity potential directly, and can be used to put limits on n in a bias-independent way

379
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 379-383.
© 1991 Kluwer Academic Publishers.
380

by utilising the absence of microwave background fluctuations. It is not independent of


the power spectrum of perturbations, and thus may be used to constrain this also. The
peculiar velocity potential is smoothed with a Gaussian filter of scale 12 h- 1 Mpc, which
reduces the effect of random errors in individual velocity measurements, and probes the
linear regime where the potential is expected to be Gaussian and curl-free.

2 The Peculiar Velocity Potential

The (smoothed) velocity potential derived by Bertschinger et al. (1990) shows a drop in
potential between the Local Group and a minimum at the 'Great Attractor' at a distance
(recession velocity) of around 5000 km s-l. The question is how likely is such a potential
drop over such a distance? This question is addressed in Heavens (1991). The answer
depends on the power spectrum of irregularities (the flow is coherent over a scale of 4000
km s-I), and also on n and the degree of nonlinearity of the density field. If n is low,
or the density field is very uniform on this scale, then a potential drop of the magnitude
seen is exponentially unlikely. As the degree of nonlinearity of the density field grows by
gravitational instability, so too does the distribution of depths oflocal minima in the velocity
potential. Thus, in order to constrain the density parameter, we need to estimate the non-
uniformity of the density field, conveniently via the r.m.s. fractional overdensity, Uo ==
(6p/p)2)1/2. This may be estimated either from the r.m.s. fractional numberoverdensity
of a population of objects (e.g. IRAS galaxies), UN = buo, or as an upper limit from the
absence of microwave background fluctuations. From Saunders et al. (1990) I interpolate
(IN ~ 0.34 for a Gaussian filter of scale 12 h- 1 Mpc.

The final parameter on which depends the statistics of the velocity potential is the
overdensity v, in units of the r.m.s., at the Great Attractor itself. If we assume that the
IRAS galaxies form biased tracers of the mass distribution, then we can obtain v ~ 2 from
the overdensity (Yahil1990), and the r.m.s. above. Full details are presented in Heavens
(1991), but some of the results are shown here. Fig. l(a) shows the likelihood for n and v
for CDM with h = 0.5, where (Jo is as large as possible for marginal consistency with the
UCSB South Pole experiment (Vittorio et al. 1991). Note that, if a position-independent
bias prescription holds at all, with any bias parameter b, then v ~ 2, close to the peak
in the likelihood distribution. Fig. l(b) shows the likelihood for n/b S/ 3 , for CDM with
hb 5 / 3 = 0.5 and 1. Note that values of n and n/b S/ 3 less than about 0.5 are excluded,
and that the maximum likelihood values lie around, or in excess of, unity. Fig. 2 shows
the likelihood plot for a wider range of power spectra, using the IRAS galaxies as tracers.
Gaussian-filtered power-laws 16kl2 <X k n are chosen, for the power spectrum at the current
epoch. A cutoff is applied at the present horizon size. Note that lower values of n are
allowed for n '" -1, but that the maximum likelihood values for fixed n still exceed 0.6.
The value of n required for nucleosynthesis (Olive et al. 1990) if the Universe is baryonic
is strongly excluded, and b ~ 1.
381

Likelihood of 0 and v for CDM (h=0.5) ., Likelihood for CDM


I
8 )(
~-r~~-r~~-r~~~~~~~
~

.,
I
o
)(
M
'0
0.,
01
=~)(
Ql
,>0:(\/
;:s
.,1

CDI/: b- o.a II,- 12.0b-I llpo


-
o

hok Uke!lbood_ 3.1711-0311,·

0.5 1.5 2
v

Figure 1: (a): The likelihood of n and v, using the UCSB South Pole microwave background
experiment to constrain 0"0, for CDM with h = 0.5. Contours at 1, 10, 20 ... 90 and 99%
of peak. (b): Likelihood distribution for CDM, using IRAS galaxies for 0"0 and v, for
hb5 / 3 = 0.5 (full) and 1 (dashed).

~
It)
..a
"'-
c

o L -__ ~ ____- L____ ~ ____L -__ ~ ____-L____ ~ __ ~

-4 -2 o 2 4

Spectral index n

Figure 2: The likelihood distribution for n/b5 / 3 and the density spectral index (present
IOkl 2 ex: k"), for power spectra truncated at the present horizon scale and filtered with a
12 h- 1 Mpc Gaussian filter. b is the bias parameter for IRAS galaxies filtered in the same
way.
382

2.1 Potential Minimum?

If the infall from the far side of the Great Attractor is not confirmed, then we may still
use the potential drop between the Local Group and the Great Attractor to constrain
the density parameter and n. The essential point is that for a given power spectrum the
constraint on n is little changed (Heavens 1991). However, in order to remove the minimum,
one may wish to introduce more large-scale power, and change the spectrum. In this case,
the constraints on high n are less strong. With complete freedom to specify the power
spectrum it is not possible to exclude any value of n.

3 CONCLUSIONS

Assuming that galaxies trace the peculiar velocity field, and that the peculiar velocity
potential arises from Gaussian gravitational perturbations, the depth of the potential at
the Great Attractor is a useful and clean test of fluctuation spectra and the value of the
density parameter. It is sufficiently deep to put a firm constraint !1 <: 0.5 for CDJvI mockls.
If n = 1, then the microwave background limits and the depth of the Great Attractor
constrain the bias parameter for IRAS galaxies on large scales to he in the range 0.7 - 1.4.

References

Bertschinger, E., Dekel, A., Faber, S.M., Dressler, A., & Burstein, D., 1990. Astro1lhys. J., 364,
370.
Dressler, A. & Faber, S.M., 1990. Astrophys. J., 354, 13.
Heavens, A.F., 1991. Mon. Not. R. astr. Soc., submitted.
Olive, K.A., Schramm, D.N., Steigman, G. & Walker, T.P., 1990. Phys. Lett., B236, 454.
Saunders, W.S., Frenk, C.S., Rowan-Robinson, M., Efstathiou, G.P., Lawrence, A., Kaiser, N., Ellis,
R.S., Crawford, J., Xiao-Yang, X. & Parry, 1.,1991. Nature, 349,32.
Strauss, M.A. & Davis, M., 1988. In Large Scale Motions in the Universe, Ed. Rubin, V.C. &
Coyne, G.V. Princeton University Press, Princeton ..
Vittorio, N., Meinhold, Muciaccia, Lubin & Silk, 1991. Preprint.
YahiI, A., 1990. Proceedings of Les Arcs meeting, Rencontres de Moriond.
383

DISCUSSION:

Kaiser: You claimed that the results are insensitive to the backside infall. This seems
puzzling. For c) to have a minimum, v must change sign, so the evidence that c) is a
minimum is exactly that there is backside infall, and such evidence seems slight at present.
Heavens: The principal constraint on n really comes from the large potential drop be-
tween the Local Group and the Great Attractor, and is not very sensitive to whether the
potential turns up at that point. However, this test is dependent on the spectrum offluctu-
ations. IT one believes that an absence of a minimum indicates power on a very large scale,
then one is testing a different spectrum, with different conclusions about n (Fig. 2).
THE ANGULAR LARGE SCALE STRUCTURE

Y. Hoffman
Racah Inst. of Physics,
The Hebrew University,
Jerusalem,lsrael.

ABSTRACT. The angular distribution of the IRAS galaxies has been expanded in spherical har-
monies and its angular power spectrum has been calculated. This distribution is dominated by the
first few multipoles and the amplitude is a decreasing function of I. We find the IRAS data to have
more power on the scale of a few x 10h -1Mpc compared with the standard CDM model, and to be
compatible with an open universe CDM model of noh = 0.2.

1 Introduction

The IRAS survey enables the construction of an almost full sky, flux-limited catalog of
galaxies which can serve as a good data base for the statistical analysis of the large scale
structure of the universe. At a flux limit of Slim = 0.7 J y, the effective depth of the sample is
about 100h- 1 Mpc (where h is Hubble's constant in units of 100km/s/Mpc). On that scale
the departure of the galaxy distribution from uniformity is small and the linear theory is
used to compare the predictions of the cosmogonic models with data. In the present paper
the angular distribution of the lRAS galaxies is analyzed by calculating its angular power
spectrum and it is compared with the predictions of the cold dark matter (CDM) model.
Here, we focused on two variants of the model, both of which obeys the inflation paradigm
of a flat universe with a scale invariant perturbation field. The first is the 'standard' CDM
model with no = 1.0 and h = 0.5 and the other is a model of an no = 0.2 and h = 1
universe dominated by a cosmological constant.
Elsewhere in this Proceedings (C. Scharf and also Scharf et at. 1991) the expansion
of the IRAS catalog in spherical harmonics is presented. The important result presented
there is that the angular distribution of the IRAS galaxies is dominated by the first few
multipoles. In particular, a map constructed out of the I = 1 and 2 recovers the main
features of the observed distribution, namely the Great Attractor/Centaurus complex at
the northern hemisphere and the Perseus-Pices superduster in the south. The angular
distribution of galaxies has been traditionally analyzed by measuring the angular two-point
correlation function, following its pioneering determination by Peebles and Hauser (1974)
from the Lick catalog. Indeed such an analysis has been applied by Lahav, Nemiroff and
Piran (1990) to the lRAS catalog. Here we suggest an orthogonal approach of determining

385
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 385-388.
@ 1991 Kluwer Academic Publishers.
386

the angular power spectrum of the galaxy distribution and use it as a probe of the primordial
power spectrum. This is based on the formalism developed by Peebles (1974) for calculating
the angular correlation function.

2 Theory

Consider now a given angular distribution of galaxies obtained from an incomplete survey
of the whole sky. The incompleteness is represented by a mask which allows galaxies to be
selected only outside the masked region. The spherical harmonics transform of the masked
sky is,
biobs = N
1 L
Yi(i),
obs i~mask
(1)

where Yi( i) is the spherical harmonics evaluated at the angular position of the i-th galaxy
and bi is normalized by the total number of galaxies outside the mask, Nobs. In the the-
oretical domain one considers a given density perturbation field 8(r). To relate the matter
and galaxy distribution we adopt the convenient convention where the biasing parameter is
unity and the power spectrum is normalized by the empirical determination of the variance
of the galaxy counts in spheres of radius R = 8h- 1 Mpc, defined as ol Assuming a full sky
coverage the spherical harmonics transform yields,

(3)

Here, the mean field is subtracted which makes the t = 0 term to vanish and the selection
function is normalized such that, 1 = 41rnJooo ~(r) where n is the mean number density
of galaxies. For an isotropic and stationary Gaussian field the ensemble average ai (th)
vanishes, there is no mode-mode coupling and the angular power spectrum is

(5)

here P( k) is the power spectrum and jl( x) is the spherical Bessel function of order t. In the
absence of an intrinsic clustering the distribution of galaxies is Poisson which is statistically
independent of the 8 field and the total power is given by:

2 2 1
AI = AI (th) + 41rN (6)

Next, the incomplete sky coverage is to be added. For a given mask we define

Ci(th) = ~o:s J d2 nYiW(n)(n) Jr2d~(r)8(r). (7)

where Wen) = 1 for n outside the mask and it vanishes otherwise and nobs = f d2 nW(n).
Note that ci is related to the directly observable bi by:

m
ci = bmI - -41r 1m
nobs- I
(8)
387

where II" is the spherical harmonics transform of W(Q). The ensemble average of cr is
zero, but the mask introduces mode-mode coupling such that

( m c, C*m')
I' n
= (411")2
obs
L..J Iwmm'12A2~
'"
I',m'
I I' ~
I ul,l,um,m" (9)

The transformation tensor is defined by

(10)

The angular power spectrum of the masked data is

nobs
C I2 = (411") 21 1+ 1 '"
L..J
IwImmI' '1 2A",
2 (11)
l',m,m'

Given the Gaussian nature of the perturbation field, C; is X2 distributed with 21 +1 degrees
of freedom. It is convenient to represent the angular power spectrum by 0, = 411"VNC"
where N is the mean angular number density. To normalize the power spectrum of the
various models we adopt here 0'8 = 0.7 as obtained by Lahav et al. (1990) and a Gaussian
filter of a smoothing length of 3h- 1 M pc has been applied for all models.
10 10

,.". .... '


,-------- ~ ...........

.1 ~----~--~~~~~~ .1 ~----~~~~~~~~

1 10 1 10

Fig. 1a Fig. 1b
Fig. 1 The amplitude of the angular power spectrum 0, VS. 1. The rms value is represented
by the solid line, and the two dashed lines are the lower and upper 95% confidence limits.
The solid triangles are the IRAS data. Fig. la is the 'standard' CDM model and Fig. Ib
corresponds to the Q o = 0.2 open CDM model.
388

3 Data vs. Models

To compare the models with the data we correct the observable bi by subtracting the mean
sky. For a flux limit of Slim = 0.7 J y the volume weighted selection function peaks at about
R = 90h- 1 Mpc. The I-th multipole is determined by the structure on a linear scale of
r '" ¥. The mask used here is a Ibl < 20° cut where b is the galactic latitude. This mask
leaves about 7000 IRAS galaxies which corresponds to a shot noise level of '" 10- 5 and at
about 1 = 15 Al is dominated by it. Note that the mask affects only the magnitude of the
shot noise as Eq. 11 is valid for any mask but the convergence of the sum over I' depends
on it. To calculate GI up to 1 = 10 this sum converges at I' = 15.
The predictions of the 'standard' and open CDM models are given in Fig. 1 where
01 is plotted against 1 up to 1 = 10. In each figure the IRAS results are plotted, and
the theoretical models are represented by the rms values and the lower and upper 95%
confidence levels. It is clear that the open CDM model fits the observations best, with all
the IRAS data point except the I = 10 lying within the 95% confidence levels, as opposed to
the standard CDM model where about half the data points lie above the upper confidence
level. The combined effect of most of the GI'S being at or above the upper 95% confidence
level makes the standard CDM to be incompatible with the data.

4 Discussion

The results reported here are in agreement with other recent studies of the large scale galaxy
distribution. The angular correlation function calculated from the APM optical survey is
consistent with the one predicted by the open CDM model and has an excess of power
compared to the 'standard' CDM, in agreement with our findings (Efstathiou et al. 1990).
A count in cells analysis of the 'one in six' red-shift survey of the IRAS galaxies finds an
excess of power on the scale of (20 - 40)h- 1 Mpc compared to the 'standard' CDM, of the
same magnitude as reported here (Saunders et al. , 1991).

Acknowledgements. The work reported here is a joint project with O. Lahav, D. Lynden-Bell and
C. Scharf. The hospitality of the lOA (Cambridge University) is gratefully acknowledged.

References

Efstathiou, G., Sutherland, W.J.,& Maddox, S.J., 1990. Nature, 348, 705.
Lahav, 0., Nemiroff, R., & Piran, T., 1990. Astrophys. J., 350, 119.
Peebles, P.J .E., 1974. Astrophys. J., 185, 413.
Peebles, P.J.E. & Hauser, M.G., 1974. Astrophys. J. Suppl., 196, 1.
Saunders, W. et al. , 1991. Nature, 349, 32.
Scharf, C., Hoffman, Y., Lahav, 0., & Lynden-Bell, D., 1991,preprint.
IS THERE ANY OBSERVATIONAL EVIDENCE FOR NON-GAUSSIAN
PRIMORDIAL DENSITY FLUCTUATIONS?

ADRIAN L. MELOTT
Department of Physics and Astronomy
University of Kansas
Lawrence, Kansas 66045
U.S.A.
ABSTRACT. A brief review of tests of observational data and a report on recent numerical
experiments is used to argue that there is little motivation from observational data for
non-Gaussian primordial density fluctuations. Any such motivation must come at this time
from the wish to consider a specific theoretical model.

1. Inflation and Primordial Density Perturbations

One of the simplest and most elegant theories for the formation of structure in the universe
is that it arose from the growth by gravitational instabilityl from quantum fluctuations
during inflation. 2 ,3 On the other hand, there are a variety of theories which include non-
Gaussian fluctuations. Without enumerating them, I wish to ask whether there is any
existing observational data which demand non-Gaussian primordial density fluctuations.
This question has been studied surprisingly little in a general way; most papers have dealt
only with the success or failure of specific models.

2. Topology Tests

Some time ago we proposed4 that it would be worthwhile to study directly the topology of
the galaxy distributions. The connectivity of superclusters and voids can be examined, after
smoothing to remove discreteness due to the existence of individual galaxies. The connec-
tivity is a topological invariant, and can be calculated as a function of density threshold for
Gaussian distributions. On very large scales, we expect that the universe will remember its
initial conditions and preserve the spongelike topology. Gravitational clustering may have
modified the topology on smaller scales, and numerical simulations have shown us what
to expect. Analysis of galaxy redshift surveys indicates that the distribution is Gaussian
on large scales and has non-Gaussian features attributable to gravitational clustering from
Gaussian initial conditions on smaller scales. This summarizes results to date on three
dimensional topology. 5
It is, perhaps surprisingly, possible to extract topological information from two-dimensional
data. 6 One thing which can be studied in this way is the cosmic microwave background.
Anistropies above dipole can be studied directly when they are reliably detected; this may
give us a window on density fluctuations before gravitational clustering proceeded. 7 An-
other use is the analysis of projected galaxy data, as carried out recently on the Lick Galaxy
Catalog. 8 The conclusions of this study are compatible with those from three-dimensional
studies, as are analysis of "slices" of redshift surveys.9
Analysis of much larger redshift surveys now underway and planned should provide much
stronger constraints on theories of structure formation. Showing that topology is compatible

389
T. Shanks et al. (ells.), Observational Tests ojCosmologicallnflation, 389-393.
© 1991 Kluwer Academic Publishers.
390

with a Gaussian distribution does not (of course) prove that it is Gaussian. However, it is
safe to say that present results provide no motivation for seriously considering alternatives.

3. Filaments, Sheets, and Voids


In recent years redshift surveys have provided spectacular confirmation that the universe
has large nearly empty voids with relatively thin walls of galaxies separating them. 1o This
is obviously a non-Gaussian distribution, and it has been used as an argument against the
simplest and most natural fluctuations that one would expect from inflation.
It is important not to confuse the predictions of initial conditions with their nature after
the nonlinear effects of gravitational clustering. Even very modest clustering can make a
distribution appear not Gaussian, but lognormal. 8 Stronger clustering, which must have
taken place to make galaxies, groups, and clusters can have stronger effects.
Most numerical modeling of gravitational clustering has concentrated on specific kinds of
initial conditions, and on trying to compare them with observation. Almost no attention
has been paid to understanding nonlinear clustering in a systematic way, so that one can
bypass this step understand and anticipate results. In a series of projects with a number of
collaborators, we are trying to remedy this by analyzing the same set of simulations from
many different points of view. l l ,12 Reporting on all these is, of course, beyond the scope of
this report, but I will try to present the main points that bear on inflation and the question
of Gaussian initial conditions.
Although Gaussian fluctuations have random phases for their Fourier components (this is
equivalent to a definition), phase correlations arise during clustering and can be explicitly
measured. 13 Models with relatively more large-scale power develop a stronger phase correla-
tions, more disperion in the size of voids 14 and a stronger three-point correlation function. 15
Since both phase correlations and the three-point function are identically zero for Gaussian
fields, this is simple proof that they can become non-Gaussian through dynamical evolution.
This is not new, but it is often forgotten.
Voids of a preferred size can arise in gravitational clustering, either as a result of a feature
in the initial power spectrum or as a result of nonlinear evolution. 14 We have tested an
algorithm17 which properly picks out a preferred scale and developed an approach which
seems to be insensitive to discreteness errors which can arise as a result of the existence of
point masses in numerical models and/or incompleteness in redshift surveys. Percolation
analysis is remarkably sensitive and shows somewhat different behavior for voids imprinted
in initial conditions versus arising in hierarchial clustering. 18
Taken together, these results indicate that a wide variety of characteristics and result from
nonlinear clustering. A picture may help convey some of the possible variety:
This is an array showing particles from nine different simulations with the same phases
for the Fourier coefficients in the initial conditions but different power spectra. We have
done nine different kinds of initial power spectra and four realizations of each. By going
to two dimensions, we greatly increase our resolution, and can still do many runs and
follow perturbations over large dynamic range. Of course, we lose some generality, but
make possible some experiments which would be nearly impossible in three dimensions.u
Resolution here is 512 2.
In this dot plot we show only 1/64 of the particles. The large dynamic range makes
plot tiny all particles impossible; we have found that color renders much more information
visible. 2 However, we can use this plot to comment on some of our main points.
The vertical columns share common initial power spectra P( k) ex: k" up to some cutoff kc •
The left column has n = 2, the center n = 0, and the right n = -2. The bottom row has
no cutoff down to the Nyquist frequency; the middle row had a cutoff at 32k" where k, is
the fundamental mode; the top row had a cutoff at 4k,. All of these models are shown at
the moment when they are just going nonlinear at 4k,.
391

Many comparisons are possible, but for the purpose of these proceedings I wish only to
call attention to the variety in appearance. A large difference in appearance can result
from simply varying the power spectrum. In particular, the flattened structures reported in
recent years seem only to require substantial large-scale power. The reader will, of course,
be aware of the many other indications of large scale power reported in these proceedings.
392

There is a continuous range of types of structures. Those to the left and bottom are best
described by hierarchial clustering. 1 Those to the top and right are better approximated
by the Zel'dovich pancake theory.18 The transition is continuous. The theories are only
approximations whose validity depends on intial conditions. The bottom center plot rep-
resents the truly intermediate case; it is also our closest analog to the power spectrum we
observe in the universe.
We have studied the relationship of pancake to hierarchial models by doing vertical cross-
correlations. We find that the strongest correlation between the top and bottom rows
happens just after the moment of pancaking, i.e. the onset of nonlinearity at kc • This
means that even in the hierarchial models the physics of the Zel'dovich approximation is
operative. The correlation is stronger as we move from left to right, but it is still significant
even in the left column.
Gaussian density fluctuations give us a rich palette with which to paint possible structures.
We are just beginning to explore the diversity of nonlinear effects. There is no compelling
need to go outside this clearly specified and physically reasonable set of initial conditions.

4. Acknowledgements. I am grateful to the organizers for this workshop, to NASA


for research support under grant NAGW-1288 and NAGW-1793, to NSF for grant AST
8911327, to my collaborators for their talents, and to the National Center for Supercom-
puting Applications, Urbana, Illinois USA for the grant of Cray-2 time which made this
work possible.

References
1. P.J.E. Pebbles, The Large Scale Structure of the Universe (Princeton University Press:
Princeton) (1980).
2. A.H. Guth and S-Y. Pi, Phys. Rev. Lett. 49, 1110 (1982).
3. R.H. Brandenberger, Rev. Mod. Phys. 51, 1 (1985).
4. J.R. Gott, A.L. Melott, and M. Dickinson, Ap.J. 306, 341 (1986).
5. A.L. Melott Physics Reports 193, 1 (1990).
6. A.L. Melott et al., Ap.J. 345, 618 (1989).
7. J.R. Gott et al., Ap.J. 352, 1 (1990).
8. P. Coles and M. Plionis, MNRAS in press (1991).
9. C. Park et al., Ap.J. (submitted) (1991).
10. V. deLapparent, M.J. Geller and J.H. Huchra, Ap.J. Lett. 302, L1 (1986) and references
therein.
11. S.F. Shandarin and A.L. Melott, Ap.J. 364, 396 (1990).
12. J.F. Beacom, K.G. Dominik, A.L. Melott, S.P. Perkins and S.F. Shandarin, Ap.J. in
press (1991).
13. R.J. Scherrer, A.L. Melott and S.F. Shandarin, Ap.J. submitted (1991).
14. G. Kauffmann and A.L. Melott, MNRAS in preparation (1991).
15. J.N. Fry, A.L. Melott and S.F. Shandarin, Ap.J. in preparation (1991).
16. S.F. Shandarin and K.G. Dominik, Ap.J. in preparation (1991).
17. G. Kauffmann and A. Fairall, MNRASin press (1991).
18. S.F. Shandarin and Ya.B. Zel'dovich, Rev. Mod. Phys. 61, 185 (1989).
393

DISCUSSION:

Coles: When you said that one has to smooth on a scale'" 20 Mpc in order to reach 25%
accuracy with the prediction of linear theory, was that the density field you were referring
to? If so, one would expect pI:.edictions of potentials and velocity fields made with gaussian
statistics to be much more accurate.

Melott: It was the density field. I agree with you about the other fields, my point was
that one must separately justify any use with tests. This is not always done.
COBE DMR OBSERVATIONS OF CMB ANISOTROPY

G. F. Smoot
Berkeley Lab fJ Space Science Lab
University of California
Berkeley, CA 94720

ABSTRACT. The COBE Differential Microwave Radiometers (DMR) have produced preliminary
full-sky maps at frequencies 31.5, 53, and 90 GHz. Galactic emission is seen unambiguously at all
three frequencies. The only large-scale anisotropy detected in the cosmic microwave background is
the dipole anisotropy. There is no clear evidence for any other large-angular-scale feature in the
maps. Without correcting for systematic effects, we are able to place limits tl.T /TQ < 3 X 10- 5 for
the rms quadrupole amplitude, and tl.T/To < 4 x 10- 5 for Gaussian fluctuations at the 95% C.L.
with To taken to be 2.735 K. We briefly review the DMR experiment and discuss some implications
of these results for cosmology in general, and inflationary models in particular.

1 INTRODUCTION

In standard Big Bang cosmology decoupling of the radiation and matter occurs at a redshift
of about 103 • Currently we can observe at that redshift roughly 104 causally disconnected
regions. The question before the inflation paradigm was why, on these large angular scales
where the regions were not causally connected, the cosmic microwave background (CMB)
was even comparable much less equal to a part in a thousand or better. A prime motivation
of the inflation paradigm was an explanation of the CMB isotropy Roughly a decade ago
when inflationary cosmology was making its debut, the measured CMB isotropy was at the
level of 10- 3 for a few sample regions of the sky.
How has inflation fared in the intervening time? Inflation has proven robust, which
means difficult to test and easy to adjust. What can measurements of the CMB tell us?

1.1 OBSERVATIONAL TESTS OF INFLATION USING COBE CMB MEASUREMENTS

Even though a high level of CMB isotropy was taken for granted a decade and a half ago,
it was not experimentally demonstrated. Now the isotropy is much better established but
it is still possible for observations to show that inflation fails. Three specific areas for CMB
tests of inflation are:

395
T. Shanks et al. (eds.), Observational Tests a/Cosmological Inflation, 395-412.
© 1991 Kluwer Academic Publishers.
396

1.1.1 Power Spectra Straightforward inflationary models predict a scale free, Zel'dovich,
power spectra and gaussian statistical fluctuations. The maximum power is at long wave-
lengths. One way to set the amplitude scale is to normalize the gravitational potential to
large scale streaming motions or peculiar velocity. Many different authors have estimated
the resulting CMB fluctuation amplitudes and a summary is shown in Table 1. While it

Authors Year ~T ~T/To

Abbot and Wise 1984 lOJLK 3 X 10- 6


K. Gorski 1991 > 5JLK > 2 X 10-6
R. Schaefer 1991 10-20JLK > 3 X 10-6
Kolb and Turner 1990 few X 10- 6
A. Starobinskii 1991 6- 8 X 10- 6

Table 1: Calculated/estimated values for quadrupole anisotropy amplitude from inflation-


ary models normalized to peculiar velocity and large scale motions.

is still possible to have inflation and not produce anisotropies this large, the models lose
some of their efficacy if some other mechanism is producing the peculiar velocities and
the large scale streaming. If the experiment and data analysis go well, then with 2 to 3
years of observations the DMR should achieve a level of sensitivity sufficient to test these
predictions.

1.1.2 Gravity Waves Inflationary models predict significant generation of gravity waves
in the very early uuiverse. After the great expansion of the universe these gravity waves
persist in the wavelength range from about 1 centimeter to significantly larger than the
current horizon. COBE is significantly sensitive to gravity waves with wavelengths larger
than about 1 Mpc. A single local planewave shows up as a quadrupole distortion of the
CMB temperature and gravity waves at the surface of last scattering show as chaotic fluc-
tuations in the CMB temperature. Early estimates of the spectrum (Starobinskii 1979) had
higher energy density at long wavelengths (il grav ~ 10-6 ) than allowed by current limits
on (~T/To)quadrupole. The next generation (e.g. Kolb and Turner 1990) normalized the
spectrum to the long wavelength limit set by anisotropy measurements.

1.1.3 Rotation of the Universe If the vorticity and/or shear in the very early universe
is sufficiently small to allow inflation, then the tremendous expansion during inflation and
the period following reduce them to very low levels, i. e.

vorticity w/H < 10- 7 and


shear (J / H < 10- 7 •

Thus the discovery of significant rotation (vorticity) or shear in the universe would discredit
not only Mach's Principle but also the inflationary models.
397

1.2 OBSERVATIONAL COSMOLOGY USING COBE CMB MEASUREMENTS

COBE CMB measurements are not only powerful in testing inflationary models but also are
a useful probe of other aspects of cosmology. In particular, large angular scale measurements
of the CMB test the geometry of the universe, isotropy of expansion, and the matter and
energy distribution at the time of the early universe. We can search the CMB for fossil
evidence of the progenitors of super clusters , dark matter, cosmic strings, domain walls, and
so forth.

2 COBE DMR EXPERIMENT

The Differential Microwave Radiometers (DMR) experiment aboard NASA's Cosmic Back-
ground Explorer (COBE) is intended to provide precise maps of the full microwave sky at
wavelengths in which competing astrophysical foregrounds (galactic synchrotron, HIl, and
dust emission) are more than 1000 times fainter than the CMB. The resultant maps provide
new information on the large scale structure and evolution of the universe.

2.1 DMR INSTRUMENT

The DMR consists of six differential microwave radiometers, two independent radiometers
at each of three frequencies: 31.5,53, and 90 GHz (wavelengths 9.5, 5.7, and 3.3 mm). Each
radiometer consists of a receiver switched at 100 Hz between two identical corrugated-horn
antennas. The antennas, designed for low sidelobes and compact size, have a main lobe well
described by a Gaussian profile with 7 degree FWHM (Toral et ai. 1989) and are pointed
60 degrees apart, 30 degrees to either side of the spacecraft spin a.xis. A detailed description
of the DMR instrument may be found in Smoot et al. (1990).
Each radiometer measures the difference in microwave power between two regions of
the sky separated by 60 degrees. The combined motions of spacecraft spin (75 s period),
orbit (103 minute period), and orbital precession (1 degree/day) allow each sky position
to be compared to all others through a massively redundant set of all possible difference
measurements spaced 60 degrees apart.

2.2 OBSERVATIONS AND DATA PROCESSING

A Delta rocket launched COBE into a 900 km, near-polar (99 degrees inclination) orbit on
November 18, 1989. Within a week all DMR channels were collecting stable data. The data
from the on-board tape recorder are telemetered daily to a ground station, where a pre-
processor strips the DMR data from the telemetry stream, merges the raw DMR data with
spacecraft attitude and orbit information, and checks the quality of the data. We reject
data of questionable telemetry quality, data without accompanying attitude information,
and spurious data written with the instrument science telemetry disabled. This accounts for
less than 1% of the data. We reject spikes and transients in the data, discarding all points
that lie more than five times the RMS scatter from the daily mean « 0.1% of the data).
There are no celestial sources other than the Moon that would exceed this limit. Finally,
398

the processor flags the data for the presence in the main beam of various celestial sources
(e.g., Jupiter and the Moon) which could contaminate the data. The results presented
below discard all data with the Moon closer than 25 degrees to an antenna.
We subtract a mean baseline, smoothed over a 4-hour period, from the data and apply a
calibration factor derived from the in-flight noise source firings. We convert the calibrated
data to the solar system barycenter rest frame but apply no other corrections for systematic
effects. A sparse matrix algorithm varies the temperatures of 6144 independent map pixels
to provide a least-squares fit to the data. The pixel size is 2.6 degrees, which is smaller
than the 7-degree DMR beam. Consequently, there is some correlation between neighboring
pixels for sources in the sky. Further details of the data processing algorithms may be found
in Torres et al. (1989).
At this time the random errors are dominated by white noise from the receivers. The
noise of a typical pixel pair has a Gaussian distribution with width 30 mK, consistent with
to the expected value 28 mK measured before launch. The COBE orbit provides full, but
not uniform, sky coverage, with the area surrounding the celestial poles observed several
times more frequently than the celestial equator. Rejection of lunar-contaminated data
further limits observations near the ecliptic plane. As a consequence, noise levels are not
uniform across the maps. Typical RMS noise per pixel is on the order tl. T ITo ~ 10-4, thus
the instrumental noise per spherical harmonic coefficient with no systematic effects would
be tl.T/To ~ 3 X 10- 6 (68% C.L.).

2.3 POTENTIAL SYSTEMATIC ERRORS

The preliminary DMR results presented below are limited not by this statistical uncertainty,
but by our current estimates of upper limits on potential systematic effects. These may be
classified into three broad categories. The most obvious source of non-cosmological signals
is the presence in the sky of foreground microwave sources. These include thermal emission
from the COBE spacecraft itself, from the Earth, Moon, and Sun, and from other celestial
objects. Although the DMR instrument is largely shielded from such sources, their residual
or intermittent effect must be considered. Non-thermal radio-frequency interference (RFI)
both from ground stations and from geosynchronous satellites has been searched for but
not yet detected.
A second class of potential systematics is the effect of the changing orbital environment
on the instrument. Various instrument components have slightly different performance with
changes in temperature, voltage, and local magnetic field, each of which can vary with the
COBE orbital position. Longer-term drifts can also potentially affect the data.
Finally, the data reduction process itself may introduce or mask features in the data.
The DMR data are differential; a sparse matrix algorithm is subject to concerns of both
coverage (closure) and solution stability. Other features of the data reduction process,
particularly the calibration and baseline subtraction, are also a source of potential artifacts.
All potential sources of systematic error must be identified and their effects measured or
limited before maps with reliable uncertainties can be produced.
A variety of techniques exist to identify potential systematics and place limits on their
effects. The most general technique is a least-squares fit to the time-dependent effect in the
calibrated data. Over a six month period, the orbital inclination and precession combine
399

to decouple orbit-related effects from large-scale structure on the sky. Another technique,
suitable for celestial foreground sources, is to produce a sky map in an object-centered
coordinate system. The contribution of the source at a given distance from beam center
can be read directly from the maps to the noise limit. As a cross-check, and for instances
where the time dependence of the signal is not known a priori, we derive upper limits
by differencing sky maps produced when the systematic effect is known to be at different
strengths. The degree to which the subtracted maps differ from instrumental noise provides
a limit to the effect in the DMR sky maps.

The extent to which the data reduction process itself may introduce artifacts in the sky
maps is not readily susceptible to analytic solution. Examples include the stability of the
sparse matrix solution in the presence of noise and signals that do not sum to zero over
a closed pixel path (e.g., drifts), cross-talk between nominally independent pixels, and the
propagation of signals too small to be fitted to the daily baselines. We use Monte Carlo
simulations to place a limit on these effects and conclude that the sparse matrix algorithm
is robust and capable of recovering the input sky map with uncertainties consistent with
the Gaussian instrument noise per pixel.

Table 2 summarizes preliminary 95% confidence level upper limits to potential system-
atic effects in the maps for each of the DMR channels. The results in many cases are
limited by sky coverage, signal to noise, and available analysis software. We anticipate in-
creasingly tighter limits to potential systematics as coverage, integration time, and software
improve. The largest limit for the dipole is the uncertainty in the absolute calibration of
the instrument, currently uncertain to 5%. The uncertainty in absolute calibration does
not create artifacts in the maps but affects the calculated amplitude of existing features
(e.g., the dipole anisotropy). We continue to acquire and analyze calibration information
and we anticipate improved calibration in the future. The next largest effect is the modu-
lation of the instrument output in the Earth's magnetic field; the primary effect is on the
dipole term of a spherical harmonics expansion. The magnetic effects will be modelled and
removed in future analysis. The A channel at 53 GHz unambiguously shows a magnetic
susceptibility; we do not include the 53A channel in the dipole results presented below.
The largest potential effects upon the quadrupole and higher-order multi pole coefficients
are magnetic susceptibility, Earth emission, and the possibility of undetected calibration
drifts. The current 95% C.L. upper limits to combined systematic errors in the DMR maps
are D. T ITo < 8 X 10- 5 for the dipole anisotropy and D. T ITo < 3 X 10- 5 for the quadrupole
and higher-order terms.

As the DMR gathers redundant sky coverage and analysis proceeds, we anticipate refined
estimates of, or limits on, these effects. It is important to note, however, that the DMR
is free from some of the systematics of previous large-scale sky surveys. The multiple
differences generated by various chopping frequencies (spin, orbit, and precession periods)
allow separation of instrumental from celestial signals. Two independent full-sky maps,
produced with matched beams at each of three frequencies, provide a powerful tool for
analysis and removal of possible systematic effects.
400

Effect 31A 31B 53A 53B 90A 90B

Signals

Earth 0.070 0.070 0.040 0.040 0.040 0.040


Moon (> 25 degrees) 0.004 0.004 0.004 0.004 0.004 0.004
COBE shield and dewar 0.002 0.002 0.002 0.002 0.002 0.002
Sun 0.001 0.001 0.001 0.001 0.001 0.001

Environmental Susceptibilities

Magnetic (dipole) 0.077 0.100 0.260 0.034 0.143 0.059


Magnetic (quadrupole) 0.014 0.017 0.022 0.024 0.079 0.006
Thermal 0.012 0.010 0.012 0.010 0.009 0.007
Voltage 0.010 0.010 0.010 0.010 0.010 0.010
Cross-talk 0.0001 0.0001 0.0001 0.0001 0.0001 0.0001

Data Reduction and Processing

Absolute CalibrationO 0.165 0.165 0.165 0.165 0.165 0.165


Calibration Drift 0.043 0.079 0.031 0.033 0.049 0.039
Baseline Drift 0.012 0.013 0.005 0.005 0.009 0.006
Antenna Pointing 0.030 0.030 0.030 0.030 0.030 0.030
Solution Stability 0.0001 0.0001 0.0001 0.0001 0.0001 0.0001

Total Systematics, Dipole 0.20 0.23 0.32 0.18 0.23 0.19


Total Systematics, 1~2 0.09 0.12 0.07 0.07 0.11 0.07

°Dipole term only

Table 2: 95% C.L. Upper Limits to Systematic Effects in DMR Maps (mK)
401

3 RESULTS

Figure 1 shows grey-scale maps of the full microwave sky in galactic coordinates. The
independent maps at each frequency enable celestial signals to be distinguished from noise or
spurious features: a celestial source will appear at identical amplitude in both channels. The
three frequencies allow separation of cosmological signals from local (galactic) foregrounds
based on spectral signatures.

3.1 DIPOLE ANISOTROPY

The most noticeable feature in the maps is the dipole anisotropy. This has a natural
interpretation as a Doppler-shifted CMB observed from a frame at motion with respect
to the CMB rest frame. Motion with speed (3 = vic relative to an isotropic blackbody
radiation field yields the observed Doppler-shifted temperature distribution (Peebles and
Wilkinson, 1968).
T(9) = To(l- (32)1/2/(1 - (3cos(9))

~ To(1 + (3cos( 9) + «(32/2)cos(29) + 0«(33)


The first (monopole) term is the unshifted CMB temperature. The second term is a dipole,
varying as the cosine of the angle between the direction of observation and the velocity.
The third term is a quadrupole, varying as the cosine of twice the angle with amplitude
«(3/2) times the dipole amplitude. The temperature distribution for non-Planckian spectra
can be derived by noting that I(v)/v3 is a Lorentz invariant, where I(v) is the spectral
brightness.
The angular distribution on the full sky maps is fully consistent with a dipole anisotropy.
The spectrum of all published dipole parameters, including those from the COBE FIRAS
experiment (Cheng et al. 1990), is consistent with a Doppler-shifted blackbody origin.
Thus we adopt the working hypothesis that the dipole is generated by the galactic peculiar
velocity to the comoving frame and the resulting Doppler shift.
The preliminary DMR-determined dipole amplitude 3.3 ± 0.2 mK implies a solar-system
barycenter velocity of 365 ± 18 km/s in the direction (Ct, 6) = (11.2h±0.2h, -7°±2°), J2000,
or (l,b) = (265°±2°,48°±2°). Unless stated otherwise, all errors are 95% confidence level.
The solar system velocity with respect to the local standard of rest is estimated at 20 km/s
toward (l,b) = (57°,23°) (Kerr and Lynden-Bell 1986), while galactic rotation moves the
the local standard of rest at 220 km/s toward (90°,0°) (Fich, Blitz, and Stark 1989). The
DMR results thus imply a peculiar velocity for the Gala.xy of 547 ± 17 km/s in the direction
(266° ± 2°,29° ± 2°).
Figure 2 shows the DMR maps with this dipole removed from the data. The only
large-scale feature remaining is galactic emission, confined to the plane of the galaxy. This
emission is present at roughly the level expected before flight and is consistent with emission
from electrons (synchrotron and HII) and dust within the galaxy. The ratio of the dipole
anisotropy (the largest cosmological feature in the maps) to the Galactic foreground reaches
a maximum in the frequency range 60-90 GHz. There is no evidence of any other emission
features.
402

Figure 1. COBE DMR full sky maps of the relative temperature of the sky at frequencies
31.5, 53.0, and 90.0 GHz. The maps are in galactic coordinates and have been corrected to
solar system barycenter.
403

Figure 2. COBE DMR full sky maps of the relative temperature of the sky at frequencies
31.5, 53.0, and 90.0 GHz. A dipole corresponding to a solar system barycenter motion of
365 km/s through a 2.735 K blackbody has been subtracted from the maps in Figure 2.
404

3.2 QUADRUPOLE ANISOTROPY

A quadrupole distribution on the sky may be expressed as

The quadrupole is often quoted in terms of the Qnns = (~T/To)rms normalization,

Q~ = 4/15 (3/4Qi + Q~ + Q~ + Q~ + Q~)


We have made a series of dipole and quadrupole fits to the data, excluding data within
several ranges of galactic latitude. We find a limit of Qnns < 0.07 mK (95% C.L.), corre-
sponding to (~T /To)rms < 3 X 10- 5 •
The current limits on the quadrupole are somewhat better than previous anisotropy
studies, but are still limited by uncertanties in estimates of many systematic errors that
may well be removed in further data processing. In an ideal case this might result in a 95%
confidence level upper limit of < 5JlK on the rms quadrupole. One important systematic
error term that may well ultimately limit the detectable quadrupole and other large angular
scale anisotropies is galactic emission. The nns quadrupole amplitude including galactic
emission (DMR 31,53, and 90 GHz data and the Relict 37 GHz data also) decreases with
increasing frequency. However, since the CMB amplitude begins to decrease with frequency,
the DMR 53 and 90 GHz channels bracket the apparent best observing frequency range. The
galactic plane produces a quadrupole amplitude at minimum of 50JLK. To obtain ultimate
sensitivity we need a level 20 to 30 times lower. How much does simply cutting the galactic
plane from the fitting achieve? Tests using the 408 MHz and IRAS 100 micron maps
indicate that reasonable galactic latitude cuts gain roughly a factor of 5 to 10. We will
have to obtain the remaining factor of 2 to 4 from careful modeling.

3.3 HIGHER ORDER SPHERICAL HARMONICS

The CMB temperature distribution may be expanded in spherical harmonics:

T(O,l/J) = LLalmYim(O,</»
I m

= L.: L.:[b,mcos(m</» + b,_msin( ml/J»)k[(21 +(~)(l- )~)!P/2 P,m(cos(O»


I m 411" +m .

where Pt'(cos(O» are the associated Legendre functions, k takes the values V2 (for m -I
'm
0) or 1 (for m=O), and the b are real numbers.

3.4 CMB CORRELATION FUNCTION

The CMB correlation function C( a) is the avera.ge over all pairs of points in the sky sepa-
rated by angle a,
405

where n1 and n2 are two unit vectors such that n1 • n2 = cos(a). C(a) may be determined
from the spherical harmonics expansion.

where ~(cos(a)) is the Legendre polynomial of order 1. Excluding components with 1=0
(mono,pole), 1=1 (dipole) and 1 >20, we find 1C(a) 1< 0.Olm/(2 for the range 15 degrees
< a < 165 degrees, corresponding to (.6.'¥:d
.6.~:2») < 10- 9.
Monochromatic limits are those which present a pattern covering the full sky and are
computed from the spherical harmonic amplitudes in a manner similar to the correlation
function; however, using only a single order I. We find a limit tl.TITo < 4 X 10-5 (95%
C.L.) for monochromatic effects.

3.5 GAUSSIAN FLUCTUATIONS

If there are Gaussian fluctuations in the CMB, the pixel-to-pixel variance observed is the
quadrature sum of the receiver noise and the intrinsic sky fluctuations. The two independent
channels at each frequency provide a means to limit such fluctuations. One can generate a
histogram of all pixel values provided by subtracting the A channel map from the B channel.
Since celestial sources will appear at equal amplitude in both channels, subtraction removes
all true signals and shared systematic errors, therefore the difference map should consist
only of receiver noise and whatever systematic effects are different between the channels.
A similar histogram can be produced by adding the A and B channel maps with dipole
removed, it should be similar to difference histogram with an additional contribution of
(twice) any celestial signals and (twice) whatever systematic effects are the same in both
channels. Using the difference histogram to estimate the receiver noise per pixel (not
constant across the map owing to unequal sky coverage) and the difference in histograms,
we compute a limit tl.T ITo < 4 X 10- 5 (95% C.L.) for Gaussian fluctuations in the CMB.

4 DISCUSSION AND INTERPRETATION

4.1 GEOMETRY

The DMR limits relate directly to the large-scale geometry of the early universe. The data
limit any cosmological feature on angular scales > 7 degrees to tl.T ITo < 10- 4 • From our
vantage point, the Hubble expansion is uniform to this level. If we assume the Copernican
Principle that we do not occupy a special place in the universe and assume that the be-
haviour of light is well described by the Liouville equation, then the combination of isotropy
and homogeneity implies that the large-scale geometry is described by the Robertson-Walker
metric, with perturbations of order tl.T ITo or smaller. Between particle physics experiments
and this large-angular-scale isotropy experiment, we have probed the geometry of the uni-
verse on size scales from 10- 16 to 1028 em. It is well described by a Riemannian manifold
with the simple Robertson-Walker metric, which reduces to the Minkowski for small times
406

and distances. This test continues to support the inflationary models and moves to the
region between post diction and prediction.

4.2 TOPOLOGICAL DEFECTS

Arguments of isotropy and homogeneity constrain perturbations in the metric to order


6T ITo. Cosmic strings provide one mechanism for such local perturbations. They are
nearly one-dimensional topological defects in the metric predicted by many particle physics
gauge theories, and are characterized by large mass per unit length I' (e.g., Vilenkin 1985
and references therein). The tension I' causes transverse oscillations at speeds near that
of light, and the gravitational potential causes lensing with bending angle independent of
impact parameter. Although the lensing causes no anisotropy in the rest frame of the string
(the surface brightness of the CMB is nearly constant), in the rest frame of an observer
there is a step discontinuity of magnitude

where (3 and 'Yare the usual relativistic parameters (Stebbins 1988). The DMR limits
6T ITo < 10- 4 imply a limit on strings with angular scale above 20 degrees such that
GI'Ic 2 < 10- 5 . On large angular scales the metric appears free from topological defects.

4.3 OBSERVATIONAL TESTS OF INFLATION

4.3.1 Rotation of the Universe The limits to quadrupole and higher-order anisotropies
limit global shear and vorticity in the early universe. If the universe were rotating (in
violation of Mach's Principle), the resultant metric would cause null geodesics to spiral. In
a flat universe, the resultant anisotropy is dominated by a quadrupole term (Collins and
Hawking 1973); further limits may be established from the correlation function C(a) and
maps (Barrow, Juskiewicz, and Sonoda 1985). The DMR results limit the global rotation
of the universe to wI Ho < 10- 6 , or less than one ten-thousandth of a turn in the last ten
billion years. This is at a level that begins to probe upper limits allowed by inflationary
models but does not threaten them and it appears that they will survive this test. This
limit is cosmologically significant because it is at a level smaller than the vorticity needed
to make the angular momentum in spiral gala..xies.

4.3.2 Power Spectra The only large-scale CMB anisotropy detected by the DMR is the
dipole, which results from the peculiar velocity of the Earth with respect to the CMB rest
frame. There is no evidence for intrinsic CMB anisotropy. This large-scale isotropy poses
a major problem for cosmology. In standard hot Big Bang models, the last significant
interaction between the CMB photons and matter was Thomson scattering, which ceased
when the universe cooled sufficiently to form neutral hydrogen atoms. This recombination
occurred over a small range (±4%) of redshift at z ;:::: 1100. The particle horizon at a
redshift of 1000 subtends about 3 degrees (;:::: 2(n o lz)1/2) from our epoch. The surface of
last scattering, observed in the microwaves today, contains over 1000 causally disconnected
regions which nevertheless have the same temperature to a part in ten thousand.
407

Inflationary scenarios provide one explanation for the CMB isotropy. In these models,
the universe undergoes a spontaneous phase transition about 10- 32 seconds after the Big

Bang, causing a period of exponential growth in which the scale size increases by 30 to 40
orders of magnitude. The entire observed universe would then originate from a small pre-
inflationary volume in causal contact with itself, eliminating this horizon problem. In the
simplest inflationary models, the pre-inflationary matter and radiation fields are diluted to
zero along with any pre-existing anisotropies. Standard inflation, however, generates scale-
free anisotropies with a Harrison-Zel'dovich spectrum. If these fluctuations are responsible
for the peculiar velocities and the large scale flows, then they should result in small but
detectable CMB anisotropies in the present universe (Gorski 1991, Abbott and Wise 1984
etc. listed in the Table 1). Although current DMR limits are an order of magnitude
above the predicted spectrum, we anticipate achieving sufficient sensitivity over the planned
mission to test the predictions of inflationary models.
A simple check of the sensitivity can be made using the power spectra and correlation
functions obtained from the sum and difference of the 53 A and B channel maps. One
sees a small dipole effect in the correlation function generated from the difference maps.
We understand this to be due to the small magnetic susceptibility of the 53 A channel
and demonstrates two things: (1) there are noticeable systematic effects we have not yet
removed from this data but plan to in future processing and (2) the DMR experiment
is able to detect effects at the level of a several X 10- 11 in the dimensionless correlation
function.) If we then compare the correlation function for the 53A + 53B sum with the
best fitted dipole removed, to that of the difference, then we see a small excess at small
angles. The excess is at about the minimum level predicted (few X 10- 11 ) but cannot be

taken as a detection or confirmation of theories because there are still untreated systematic
effects in the data. This demonstrates that the DMR experiment may be capable of testing
the predicted Zel'dovich spectrum at the level set by large scale motions. If the effect is
not present, then new physics will be required to generate the motions without violating
the CMB isotropy limits. This would erode the power of the inflation paradigm.

4.3.3 Gravity Waves Gravity waves provide another source of perturbations in the metric.
A long-wavelength plane gravity wave propagating through this portion of the universe
would distort the metric and cause a quadrupole anisotropy in the CMB. For a single plane
wave propagating along fJ = 0 the resultant CMB anisotropy is

D.T/T::::: 1/2(A r - Ae)(l- cos(fJ))cos(2¢)

where Ar and Ae are the proper strains at the emitter and receiver, respectively (Burke
1975). A similar effect occurs for gravity waves propagating at the surface oflast scattering
causing a different strain as a function of position at the surface of last scattering (Ae).
The DMR is sensitive primarily to gravitational waves with scale sizes> 7 degrees at the
surface of last scattering, or about 200 Mpc today. Other smaller scale measurements
give information down to about 100 Mpc before scattering on the surface of last scattering
confuses and masks the effects but in general all lead to limits on the order of

d!lGW < 10- ((D.T/To)2)


dInA -
9
10- 5
408

for wavelength range on the order 100 Mpc (Linder 1988). The limits !:!..T/T < 3 X 10-5 for
quadrupole and 10-4 for higher spherical harmonics limit the energy density of single plane
waves or a chaotic superposition to place a statistical limit on the possible energy density
of long-wavelength gravity waves, which, expressed as a ratio of the energy density to the
critical density, is

QGW <4X 1O-3 (AGw/10Mpc)-2h- 2

where h is the Hubble constant in units 100 km/s/Mpc.


Ultra-long wavelength gravity waves or inhomogeneities in the density in the early uni-
verse also lead to temperature anisotropies in the CMB as the CMB photons climb out of
gravitational potential wells varying across the surface of last scattering:

!:!..T/T ~ 1/2!:!"P(Ho L?
P c
(Sachs and Wolfe 1967, Grischuk and Zel'dovich 1978). The results have implications for
structures beyond the present Hubble radiusj in particular, this provides a limit on the
energy density of gravity waves with wavelength larger than the horizon:

QGW < 5 X 1O- 8 h- 2 j A > lH = c/ Ho


Figure 3 is a plot of the limits on gravitational wave energy density versus wavelength
set by the CMB measurements. How does this compare to the level predicted by inflation?
If the expansion of the universe during the hadron era (in standard models, the interval
of time between 10- 43 sec and 10- 6 sec) had been governed by a "stiff" equation of state
p ~ p, then the spectral energy density of high frequency (10 11 Hz) would be enormous.
This is not observed. By contrast in inflationary models the implicit equation of state is
p ~ -p and the graviton spectral energy density grows at low-frequency « 1O- 16 Hz).
Some inflationary models predict a spectral energy density level that is too high, contra-
dictory to the large angular scale CMU anisotropy measurements. (e.g. Starobinskii 1979,
and perhaps Linde 1990)
Inflationary models predict temperature fluctuations from gravitons 6TT ~ mplanck
_H __ ,
where H is the Hubble parameter during inflation (a period characterized by a constant
Hubble parameter). As a result, the possible values of the Hubble parameter at the infla-
tionary stage are significantly restricted by the anisotropy limits.
The relationship between the inflation vacuum Hubble parameter and the energy density
is
H2 = 87rG pv = 87rG M4 = 87r M4
3 3 3 m~lanck
where M is the energy scale of inflation and energy density of the vacuum, Pv M4. The
energy density of gravitons is

(")
Ugraviton = 10-4 mplanck
4
PV = 1O- 4( M )4
mplanck

for wavelengths significantly smaller than the horizon, rising to PV /mPlanck at horizon size.
The limit on the quadrupole CMB anisotropy and smaller angular scales limits Hin/lation <
409

o
GRAVITY WAVE ENERGY DENSITY VS WAVELENGTH

Bubble Collisions
-5 Pulsar Limits
II/////.U

-<
.E
"C -10
"'-
J
"C Max.Inflalion Quantum Fluctuation

-15 Turner & Wilczek

-20 0.9K Thermal Background

o 5 10 15 20 25 30
Log Wavelength (em)

Figure 3. Limits on the energy density of long wavelength gravity waves. The upper curves
are from the DMR e."{periment. There are two curves one based upon the ~ T and strain
to energy density relationships shown here and the corner based upon the formulas from
Linder (1988). The curve marked quantum fluctuation is the predicted spectrum from
inflation from Turner and Wilczek (1990) and at maximum allowed normalization. The
normalization on the prediction can be adjustable depending upon the choice of elementary
particle theory.
410

3 X 1O- 5 mplanck, the energy scale of inflation is far removed from the Planck scale (M
< 1017 GeV), and the energy density of the vacuum is PV = M4 < 1O- 9 mplanck.
At this time it appears that it is possible to adjust parameters to allow the very large
scale gravity wave energy densities to match, there is not unlimited freedom to accomplish
this and it further constrains the possible elementary particle physics models. In the current
theoretical framework inflation must take place between the grand unified energy scale and
the time of baryogenesis. Baryogensis is the period in which the baryon excess relative to
antibaryons is generated. Over the last decade baryogensis has conventionally been thought
to occur shortly below the grand unified energy scale leaving little room for adjustment of
the time of inflation and thus gravitational spectrum, Primack (1991, priv. comm.) has
indicated that a small but growing body of theorists now think that baryogensis may occur
at the quark-lepton phase through the mechanism of sphalerons. This much lower energy
regime would allow more than adequate range to lower the inflation energy scale and gravity
wave background. Turner and Wilczek (1990) point out that inflation also predicts there will
be no 0.9 K thermal background of gravitational waves that might otherwise be expected
in the Big Bang and that extended inflation would produce an extra bump in the spectrum
due to bubble collisions.
During inflation zero point quantum fluctuations produce density fluctuations and gravi-
tons (scalar and tensor fields) at levels comparable to the vacuum energy. Thus at the
horizon wavelength scales they should produce comparable CMB anisotropies. The CMB
anisotropy lhnits are now just pushing down to the level which is needed for observed ve-
locities/gravitational potentials and beginning to limit the energy scale of inflation at a
natural level. If we were optimistic, we could hope to see anisotropies at the 10- 5 level,
giving us the density fluctuations we need and telling us the inflation energy scale.

5 CONCLUSIONS

The DMR instrument is working well and continues to collect data. The current limits
reflect quick-look data before removal of any potential systematic effects. The data show the
expected dipole anisotropy, consistent with a Doppler-shifted thermal spectrum. Galactic
emission is present at levels close to those expected before launch, and is largely confined
to the plane of the Galaxy. There is no evidence for any other large-scale feature in the
maps. We find a limit f:l.T /To < 3 X 10- 5 for the rms quadrupole amplitude, f:l.T /To <
4 X 10- 5 for monochromatic fluctuations, and f:l.T /To < 4 X 10- 5 for Gaussian fluctuations.
There is no feature with f:l.T /To > 10- 4 on angular scales > 7 degrees. The results are
consistent with a universe described by the Robertson-Walker metric and show no evidence
of anisotropic expansion, rotation, or localized defects (strings). Inflation is supported by
these measurements although they do limit the energy scale of inflation and the vacuum
energy density.
The DMR is expected to continue operating until it is shut off. At the end of two years'
operation, we can expect sensitivities on the order of f:l.T /To ~ 2 X 10- 5 per 7 degree field
of view. We have begun the process of identifying and removing systematic effects from
the data. More data and analysis will provide improved calibration and improved limits on
possible systematic effects in the maps.
411

Acknowledgements. The National Aeronautics and Space Administration Goddard Space Flight
Center (NASA GSFC) is responsible for the design, development, and operation of the Cosmic
Background Explorer. GSFC is also responsible for the software development through to the final
processing of the data. We gratefully acknowledge the many workers whose contributions ensured
the success of the COBE mission. The COBE research is a team effort including some 19 original
Science Working Group members and other scientists (see for example, Smoot et al.). This work
supported in part by the Director, Office of Energy Research, Office of High Energy and Nuclear
Physics, Division of High Energy Physics of the u.S. Department of Energy under Contract No.
DE-AC03-76SF00098. We also thank NATO for the workshop and the Local Organizing Committee
for the work they did in providing the workshop, Observational Tests of Inflation at the University
of Durham.

References

Abbott, L.F., and Wise, M.B., 1984. Astrophys. J. Lett., 282, L47.
Barrow, J.D., Juskiewicz, R., and Sonoda, D.H., 1985. Mon. Not. R. astr. Soc, 213, 917.
Bond, J.R. and Efstathiou, G, 1984. Astrophys. J. Lett., 285, L45.
Burke, W.L., 1975. Ap. J., 196,329.
Cheng, E., et al., 1990. BAPS, 35, 971.
Collins, C.B., and Hawking, S.W., 1973. Mon. Not. R. astr. Soc, 162, 307.
Fich, M., Blitz, L., and Stark, A., 1989. Astrophys. J., 342, 272.
Gorski, K., 1991. Astrophys. J. Lett., 370, L5.
Kerr, F.J., and Lynden-Bell, D., 1986. Mon. Not. R. astr. Soc, 221, 1023.
Kolb, E., and Turner, M., 1990. in The Early Universe, p. 290, Addison Wesley.
Linde, A., 1990. in Particle Physics and Inflationary Cosmology, p. 245, Harwood.
Linder, E.V., 1988. Ap.J., 326, 517.
Paczynski, B., and Piran, T., 1990. Astrophys. J., 364, 341.
Peebles, P.J.E., and Wilkinson, D.T., 1968. Phys. Rev., 174, 2168.
Schaefer, R., 1991. in After the 1st 3 minutes, AlP.
Smoot, G.F., et al., 1990. Astrophys. J., 360, 685.
Smoot, G.F., et al., 1991. Astrophys. J. Lett .. , 371, L1.
Starobinskii, A. A., 1979. JETP Lett., 30, 682.
Starobinskii, A. A., 1985. Soviet Astron.Lett., 11, 133.
Starobinskii, A. A., 1991. private communication
Stebbins, A., 1988. Astrophys. J., 327, 584.
Turner, M. S. and Wilczek, F., 1990. Phy.Rev.Lett., 65, 3080.
Toral, M.A., et aI., 1989. IEEE Transactions on Antennas and Propagation, 37, 171.
Torres, S., et aI., 1989. Data Analysis in Astronomy, ed. Di Gesu et aI., Plenun Press, 319.
Vilenkin, A., 1985. Physics Reports, 121, 263.
412

DISCUSSION:

Lasenby: Should we interpret the limit of ¥ < 4 X 10- 5 for Gaussian fluctuations as a
95% confidence limit? Does 'Gaussian' refer to a Gaussian statistical process or a Gaussian
shaped correlation function. How was the angular scale defined for setting the limit? (e.g.
as the coherence scale of a correlation function?)

Smoot: The limit is a 95% confidence level for Gaussian fluctuations in pixel amplitude
added by a Gaussian distribution in the CMB. That is, at the 70 scale the CMB has a
Gaussian distributed fluctuation with a < 4 X 10- 5 at the 95% confidence level.
Starobinsky: In contrast to the shear which is present due to adiabatic perturbations
and the gravitational wave background, the primordial rotation of the Universe should be
exceedingly small. This is the prediction of any theory assuming isotropy at early times, not
only of inflation. For example, it follows from the standard cosmological nudeosynthesis
theory. As for the gravitational wave background, your results may be used to obtain a
model-independent upper limit on the value of the Hubble parameter H(t) at the end of
inflation (strictly speaking, about 70 e-folds before its end) that is of the order of 10- 5tPI.
OBSERVATIONS OF MICROWAVE BACKGROUND ANISOTROPY
AT TENERIFE AND CAMBRIDGE

A.N. LASENBY
Mullard Radio Astronomy Ollservato1'1/,
Cavendish Laoorato1'1/,
Madingley Road,
Camllridge CB9 OHE.

R.D. DAVIES & R.A. WATSON


Nuffield Radio Astronomy Laooratories,
Jodrell Bank,
Macclesfield,
Cheshire SKll 9DL.

R. REBOLO, C. GUTIERREZ & J.E. BECKMAN


lnatituto de Astrofisica de Canaria"
98200 La Laguna,
Tenerife,
Spain.

ABSTRACT. A brief summary of the status of microwave background anisotropy observations at


Tenerife and Cambridge is given. New 15GH. Tenerife results give a preliminary limit of I1T/T
~ 2.5 x 10- 5 on an angular scale of 4°. The features reported previously at 10 GR. appear to
have a galactic spectrum, and can be predicted, though only approximately, from lower frequency
sky surveys. At Cambridge, the newly completed Ryle Telescope is making its first observations of
clusters of galaxies in order to measure the Sunyaev-Zel'dovich effect, and a dedicated interferometer
to detect primordial perturbations is under construction.

1 Introduction

The degree of anisotropy of the cosmic microwave background radiation (CMB) provides
one of the main observational constraints on theories of galaxy formation, and the theory of
inflation. On small angular scales the imprints of the formation of galaxies and clusters of
galaxies are expected, while on degree scales upwards one is probing the long wavelength tail
of the primordial perturbation spectrum, which in the inflationary theory is produced via
the amplification of quantum irregularities. H inflation did not occur then the theory has to
explain how regions of the sky on angular scales ~ (0/(1 +Zre.,))1/2 radians, corresponding
to the horizon distance at recombination, manage to have temperatures agreeing to better
413
T. Shanks et aI. (eds.). Observational Tests o/Cosmologicallnf/ation. 413-418.
© 1991 Kluwer Academic Publishers.
414

than 1 part in 104 •


In order to further improve constraints upon the CMB anisotropy, and ultimately to
detect fluctuations, very sensitive observations at a variety of frequencies are required, made
from the ground, balloons or satellites. Each has its own advantages and disadvantages.
Balloon measurements have relatively short integration times, but can use high frequency
bolometer systems with a wide effective bandwidth, thus providing good sensitivity. The
high frequency, however, means that contamination is present both from galactic dust and
possible residual fluctuations in the atmosphere, particularly ozone (see e.g. Meyer et al.,
1991). Both of these contaminants can be difficult to model or subtract. Satellite mea-
surements can be made at frequencies designed to minimize both galactic synchrotron and
dust emission (e.g. the two outer COBE channels at 31 and 90 GHz bracket a minimum in
galactic emission - Smoot et al., 1991) and give whole sky coverage. The receivers however
tend to be much less sensitive than equivalent ground or balloon ones, and spreading ob-
servations over the whole sky means that less sensitivity is obtained for individual patches.
Also galactic emission, while minimized, is still present, and must be subtracted, and since
in the COBE case the two outer channels are affected by different contaminants, with
quite different spectra, this means that supplementary high and low frequency information
must be used. There are thus some advantages to working at lower frequencies, where only
galactic synchrotron and free-free emission need be considered, and from the ground, where
equipment may be run stably over periods of years, and very good sensitivity built up over
significant sky areas. We discuss here continuing ground-based observations made using two
switched beam twin-hom systems at 10 and 15 GHz sited in Tenerife, which are currently
setting probably the most stringent constraints on degree-scale anisotropy, and compare
these results with theoretical predictions. In the final section some new developments in
ground based CMB interferomtery from Cambridge are described.

2 The Tenerife Observations and Results

The observations are made using twin beam systems sited at Teide Observatory, Tenerife,
at an altitude of 2300 In. The originall0A GHz system (Davies et al., 19S7) had So FWHM
beams with So separation, further switched through So by a rocking mirror to give a triple
beam pattern on the sky. The horns were then lengthened to give 5° FWHM beams but
with the same switching geometry and at the same frequency (see e.g. Watson et al.,
19S9). This system is now running in parallel with a similar system at 15 GHz, which
has the same beam sizes and switching geometry, but due to increased bandwidth and
improved amplifiers, is about a factor of two more sensitive. Daily scans are taken at
a constant declination determined by the angle to the horizontal of the rocking mirror.
Results to be discussed here relate to the scans taken in the quietest Galactic region at
Dec=40°, along which we have greatest sensitivity. Fig. 1(a) shows the stacked results from
the 5° FWHM 10 GHz scans. At 10 GHz we also have scans at neighbouring declinations of
35°, 37.5°, 42.5° and 45°. These were used in a maximum entropy solution to reconstruct
the microwave sky and attempt to separate its contribution from that of noise and long-
timescale (> Sh) variations which are present in the daily switched data and are fitted
for as separate channels. This sky reconstruction, when reconvolved with the beam and
415

(E~
- -- - -- - - -- - - --- -
f:E:::::: : : ::::,=3
~

- - - - - - - - - . - -- - - -
~

(:~
- - - - - - - - - -- - - - - - ~

Figure 1: Results from the 10 GHz 5° FWHM experiment. (a): The stacked scan over the
range 120° to 280° with 10- error bars. (b): Reconvolved maximum entropy result. (c) Point
source prediction. (d) Galactic prediction at 10 GHz using both the 408 and 1420 MHz sky
surveys. (e) Galactic prediction using the 408 MHz survey alone, and a constant spectral
index of2.8.

switching geometry, is shown in Fig. l(b). In Fig. l(c) we show the expected point source
contribution at 10 GHz and in Figs. l(d) and (e) show the expected Galactic contribution,
first using both the 408 and 1420 MHz surveys (Haslam et 41., 1982; Reich & Reich, 1988)
and in (e) using the 408 MHz survey alone, with an assumed spectral index of a = 2.8. It
can be seen that over the region where we have best signal/noise, 200°-250° in RA, there
fOoJ

is a reasonably good match between the combined Galactic and point source predictions,
and the MEM reconstruction, provided one uses the Galactic prediction in Fig. 1(d). In
our original publication (Davies et 41., 1987) only the 408 MHz survey was avaiable to us,
which does not predict the feature seen in our data at RA = 225°. The reliability of
the 408 MHz/1420 MHz extrapolation is open to question since systematic scanning effects
leading to spurious arcs and ridges are present in both surveys (Davies et al., in preparation),
but our own work, together with that of Banday & Wolfendale (1990), does make it seem
416

...
~ ~~~~~~~~~~~~~~~~~~~~~~

i-.:1}I I I IrWi~IJI"l l pI~IDI~P~Il IJfl lnIQql l plI lhl l hI I IDl hlidIlI Iq1qfdJI I I I I 1 1 \lpl l l fl l l lm~'II
_ _ _ _ _ _ _ m ~ _ _ _ _ _ _ _ _

f:~ - _ - _ _ _ _ _ ~ _ _ m _ _ _ _ _

LE::::,::::::::::.::: : : :.: :::::::"~.:=J


_ _ _ _ _ _ _ _ ~ _ _ m _ _ _ _ _

i+:=~'~~~:>=: : : :3
- - - - - - - - - - - - - -- - ~

il~=':'~:~:~~=::::::::::::::::::":
- - - -- - - - - ---- =-: -::]
-- ~
Figure 2: Results from the 15 GHz 5° FWBM experiment. (a)-(e) as for Fig.I.

likely that the features seen at 10 GHz should be ascribed to the Galaxy rather than the
CMB. This interpretation is borne out when considering the new 15 GHz data, for which
equivalent plots are shown in Figs. 2(a)-(e). Even though this represents < 3 months data,
it can be seen to have a much smaller scatter than the 10 GHz data (which represents about
90 days integration time on Dec 40° itself). Considering the best signal/noise region again
(N 200°-280°), and bearing in mind that in this case there are no neighbouring declination
scans to guide the maximum. entropy solution, which is therefore less reliable here, one sees
that the results are consistent with a much reduced Galactic and point source contribution,
and no CMB fluctuations. The two bumps at RA '" 172° and 1880 , which intruigingly
seem to be apparent in the 10 GHz reconstruction as well, are not stable under division of
the data into two independent sets, and thus should be regarded with suspicion, and are
in any case in regions of poorer signal/noise ratio. A maximum likelihood analysis of the
region 161 0 to 2300 in RA, with proper account taken of the errors, yields no evidence for
a detection, and a Bayesian integration gives an upper limit of'" 2.5 X 10-6 in AT /T on a
coherence scale of '" 40 •
This result can be compared with a limit of '" 3.8 X 10-6 on 10 scales at 90 GHz from
417

the recently reanalyzed UCSB South Pole experiment (Vittorio et al., 1991) and limits from
'" 2 X 10-& to 4 X 10-& on scales from 3° to 22° from the MIT anisotropy balloon experiment
(Meyer et al., 1991). In the latter case, however, the authors state that there is clearly
evidence for some sort of anisotropy in the data at a level of AT IT ~ 1-3 X 10-& on scales
~ 4°, presumably due to Galactic dust or atmospheric contamination at the high frequency
employed (170 GHz). Comparing with theoretical predictions, standard Cold Dark Matter
models with adiabatic perturbations predict peak power on angular scales closer to that
of the UCSB experiment rather than ours, and hence despite our slightly better upper
limit, the Vittorio et al. limit is more restrictive. For isocurvature perturbations however,
and in particular the reionized baryonic isocurvature model of Peebles, our new results
will be particularly restrictive. For example, taking the old 10 GHz detection levels as
equivalent upper limits, Sugiyama et al. (1990), find that even with the inclusion of a
cosmological constant (which loosens the constraints from the CMB), Peebles models with
a scale invariant spectrum are excluded for all values of O. Several other models which
Sugiyama et al. find are just allowed on the basis of the old Tenerife result, also become
excluded using our new 15 GHz limit. Integrations at 15 GHz will continue over the coming
year, during which time we hope to be on course for reaching the 1 X 10-5 level for CMB
upper limits.

3 CMB Interferometry at Cambridge

At Cambridge work is proceeding on construction of a three-element interferometer to make


observations of the CMB on angular coherence scales from 10' to 45', where power from most
models is expected to peak. This system, the Cosmic Anisotropy Telescope (CAT), is being
built in order to try to reach detection levels of a few X 10-6 in AT IT in integration times of
a few months, via a very careful design which tries to eliminate most sources of systematic
error, whilst still achieving a good filling factor, and hence temperature sensitivity for the
interferometer. The telescope will run at three frequencies over the band 12 to 18 GHz, so
that Galactic synchrotron and free-free emission can be separated from CMB fluctuations
via their quite different spectral signatures. Smallest baselines are approximately 1 m. The
CAT, which is expected to be in operation in 1993, is seen as a preliminary stage for a 'larger'
instrument, called the VSA (Very Small Array), with of order 20 elements, able to provide
high quality images of the CMB on angular scales from 10' to 2°, and with temperature
sensitivity at least as good as CAT. Again measurement over a range of frequencies is
planned, in order to separate out the CMB contribution to the anisotropies measured.
The Ryle Telescope at Cambridge (an enhanced version of the old 5-km) is now in
operation at 5 GHz and will shortly move to 15 GHz. This telescope consists of 8 13 m
Cassegrain antennae on an E- W baseline, with 4 fixed and 4 moveable elements. High
sensitivity to CMB anisotropies has been provided by cooling all the receivers, increasing
the bandwidth to 350 MHz and implementing a 'dense-pack' configuration for 5 of the
antennae, with baselines as small as 18 m. This is to be compared with minimum baselines
of 30 m and 45 m at the Australia Telescope and VLA respectively. The current main use for
the telescope, occupying approximately half its time, is for study of the Sunyaev-Zel'dovich
effect in clusters of galaxies (see e.g. Jones, 1991), but it will also be very useful for studies
418

of primordial perturbations. The first two-dimensional maps of the S-Z decrement in several
clusters are expected this year, and there is great interest in combining this information
with X-ray data in order to put bounds on Ho.

References

Banday, A.J., & Wolfendale, A.W., 1990. Mon. Not. R. G8tr. Soc, 243,576.
Davies, R.D., Lasenby, A.N., Watson, R.A., Daintree E.J., Hopkins, J., Beckman, J.E.,
Sanchez-Almeida, J., & Rebolo., R., 1987. Nature, 326,462.
Haslam, C.G.T., Salter, C.J., Stoffel, H., & Wilson, W.E., 1982. Astr. A6trophY6. Suppl., 47,1.
Jones, M.E., 1991. in Interferometry: Theory, Techniques and Applications, ed. Cornwell, T.,
NRAO, in press.
Meyer, S.S., Cheng, E.S., & Page, L.A., 1991. Preprint.
Reich, P., & Reich, W., 1988. Astr. Astrophys. Suppl., 74, 7.
Smoot, G.F., et al., 1991. Astrophys. J. Lett, 371, L1.
Sugiyama, N., Gouda, N., & Sasaki, M., 1990. A6trophY6. J., 365, 432.
Vittorio, N., Meinhold, P., Muciaccia, P.F., Lubin, P., & Silk, J., 1991. Preprint.
Watson, R.A., Rebolo, R., Beckman, J.E., Davies, R.D., & Lasenby, A.N., 1989. in Large scale
structures and motion in the Universe, p133, eds. Mezzetti, M., Giulicin, G., Mardirossian,
F., & Ramella, M., Kluwer Academic.
FOREGROUND EFFECTS AND THE SEARCH
FOR FLUCTUATIONS IN THE CMB RADIATION

A.J.Bandayl, M.Giler2 , B.Szabelskas , J.Szabelskis & A.W.Wolfendale l


1 Phy.ie. DeptJrCmenC, Univer.iC" Durham, U.K.
:I1ndifute 0/ Phy.ie., Univer.if" L6d4, Poland
s 1",CiCuCe 0/ Nuclear Scudiu, L6dl, Poland

ABSTRACT. Precise measurement8 of anisotropie8 in the Cosmic Microwave Background radiation


are of great importance to modern cosmology. Of special interest here are the recent measurements
at .... 10GH. where some evidence for a detection of cosmological ftuctuations has been claimed.
However, background effects, particularly these due to the Galactic Synchrotron Radiation and
infra-red emission from dust, can contaminate the genuine signal. We examine these local fields
in some detail and conclude that predicted fluctuations in the GSR are of the same magnitude as
the observed ones at 10GH•. Future searches for CMB fluctuations will be possible at frequencies
greater than 20GH. provided that the nature of the Galactic signal is better understood.

1 Introduction.

Studies of 8uduations in the Cosmic Microwave Background can, in principle, provide infor-
mation about early phases of the Universe, especially the era of recombination at a redshift
z ,...1500, and also about the clumping
of matter leading later to the formation 10-19 .--------:--r-----r1
of galaxies and galaxy clusters. Such /
Buctuations should be seen on a vari- /
ety of angular scales and their magni- I
tude, expressed as ~l, predicted by the Fv /1
favoured cosmological models should (Wm-2 -21 /1
be of the order 10-8 - 10-6 on a scale Hi1 (A
of a few degrees. Of special impor- sr-1 ) -22
I! dust
tance here are the measurements of the " , II II
Jodrell-IAC-Cambridge group (Davies
et al., 1987, Watson, 1989, Watson -23 'I)
I '
et al. , 1989, Rebolo et al. , 1989) at I
a frequency of 10.46 GHz which uti- -24 L-....L.-.L._ _-'-_----"L......L_--'-_--'
lizes the twin hom drift scan system 10 7 10 11
10
of the Tenerife apparatus. The authors v(Hz)
have claimed a probable detection of
¥ = 3.7.10- 11 over an angular scale Fig.I. CMB spectrum and other contaminating
of 8° (Davies et al. , 1987). radiation fields for galactic latitude b=900.
419
T. Shanks et al. (eds.J, Observational Tests o[Cosmological Inj/ation, 419-423.
© 1991 Kluwer Academic Publishers.
420

However, in this frequency region, the contaminating effects of Galactic Synchrotron


Radiation due to cosmic ray electrons gyrating in the local magnetic field become very
important. Fluctuations in the cosmic ray background are connected with the changes of
the magnetic field value Crom point to point and the variations of the electron density caused
by SNR shocks passing through the Interstellar Medium (Szabelska et aI. , 1991), or they
are related to the large scale distribution of cosmic rays. To evaluate the angular scale and
magnitude of these variations we are examining Faraday rotation data, atomic hydrogen
column densities N(HI) as a function of galactic coordinates I, b and also the theory of
small-scale CR irregularities. The work is being extended to higher frequencies, where
infra-red emission from dust becomes important. In what follows we shall consider the
region ofspace at high Galactic latitudes (b-400) where fiuctuations in the contaminating
backgrounds are supposed to be smallest. Fig.l shows the CMB spectrum together with
other contaminating radiation fields for b = 90°.

2 Cosmic Ray Background.

The GSR spectrum measurements extend to -5GHz only and we have to extrapolate
from lower frequencies to predict the shape of the spectrum at higher frequencies. For an

Fig.2. Comparison of observed


and predicted second differences
at 10.46 GHz. P: predicted by
USj R: Rebolo et aI. (1989), WI:
Watson (1989), W2: Watson et
al. (1989). The points are aver-
ages over 10 degrees in R.A.

,-$ a.n·,-
.......; ..,"'.:-.,, ,""", "
..••!.!'.. .:-:. ,:,:"" ... , ..
.................."":.:",' ...... ...... ,
...... ,... ,
\ '"
...............
••. 't.u-~~~"""'"--':~~~.~ .....::--:/''""'"--':~~~~'""'""':!~~
m m m ~ m m ~ ~ * m ~
R. A. Idegrees I
421

ambient electron spectrum of the form N(E) = K· E-"I, the synchrotron flux will vary as
1(11) ex ~-1112-'. where B is the magnetic field intensity and fJ = ('1+ 3)/2. The electron
spectrum steepens with energy due to increasing energy losses, hence fJ also increases with
frequency. However, the analysis of detailed radio-maps at frequencies 38, 408, 820 and
1420MHz (Lawson et AI. , 1987) shows that there is considerable variation of the spec-
tral index fJ from place to place on the sky with a dispersion that grows with increasing
frequency.
In Fig.2 we give a comparison of the observed second differences at 10.46 G Hz with those
predicted from radio data. What we may conclude is that the amplitude of the observed
and expected second differences is similar, although there is no point-to-point agreement.
The discrepancies are probably due to the variations of fJ and to experimental noise (with
the latter probably being more important).
1·0
10 0 "-
(191 "-
,,
-1 ...... ....
~ 0·5 \
T \ /
-2 \ I
\ I

-) 0·0
\
--
...
\
/
-- _.-._.-
,_/ .....
""-, ,
/\
/ \
-~
\/
/ \ ,,
"-
_---
-0·5
-5 "... .....

10-6
1 10 100 -1·0
vlGHzl
o· 10· 20· 30·
9-
Fig.3. Predicted GSR anisotropy ver- Fig.4. Comparison oC CMB and GSR correla-
sus frequency (Cor FWHM 5.6°). The tion functions C(O). Broken lines are extremes
values relate to the rms of 2nd differ- from the sample oC 5 correlation functions com-
ence fiT, divided by T=2.7K, denoted puted from cuts across one realisation oC the
as If and arise Crom an extrapolation CMB in the isocurvature model oC Bond and
of the measured T 408 and T 1420 maps. ECstathiou. eLSA' denotes their large scale av-
The lower curve allows for the steepen- erage. Chain lines represent extremes Crom the
ing of the spectra with increasing fre- sample oC 7 GSR correlation functions for fre-
quency. The data refer to 6 = 40°, quency " = 408 MHz.
RA: 180° - 250°. Also shown is the
signal given by Davies et AI. (1987)
and the upper limit from Rebolo et
AI. (1989). The CMB prediction for the
isocurvature model is also shown.
422

In Fig.3 we show the predicted GSR anisotropy as a function of frequency together with
the experimental results and expected range of CMB anisotropy for a typical model, the
isocurvature model of Bond and Efstathiou (1987). It can be easily seen that the predicted
fluctuations in GSR are ofsimilar magnitude as the observed ones at 10GHz, but the role
of GSR effects decreases quickly with increasing frequency, so future searches for CMB
anisotropies should be easier at higher frequencies.
We shall now examine the problem of whether or not the correlation function C(B) can
be used in order to distinguish the cosmological fluctuations from the fluctuations arising
in the GSR field (see Banday et al. , 1990). In Fig.4, correlation functions computed from
cuts across one realisation of the CMB in the isocurvature model are presented (these
should be compared with the large scale average correlation function also shown) together
with correlation functions for GSR at a frequency v = 408 MHz for different values of 6
bordering 6 = 40°. It is apparent that both GSR and CMB correlation functions have
similar shapes (at least for the smoothing inherent in the present study). The anisotropies
of GSR and CMB are seen to have similar angular scales and it will therefore not be possible
to distinguish them on the basis of the correlation function, if its shape does not change at
higher frequencies.

3 The Situation at Higher Frequencies - the Role of Emission from Dust.

In order to improve the prospects for the


detection of CMB anisotropies, searches
must be made at higher frequencies where
the contaminating role of GSR diminishes.
However, for frequencies above "" 20 G Hz
emission from dust heated by the interstel-
lar radiation field in our Galaxy becomes
important (Fig.1). High latitude infra-red
radiation from dust is strongly correlated
with the atomic and molecular hydrogen -5
distribution (Boulanger and Perault, 1988) "- "-
"-
I
-_/
and should thus be amenable to study. The
emission from gas and dust should on aver-
age vary as cosec(b), but the scatter around
10. 6
the mean cosecant law is rather large indi-
cating that the distribution of interstellar
10 10 " fj12

vlHzl
matter is clumpy. Dust-to-gas ratio varia-
tions, as well as uncertainties in the amount Fig.5. Second differences calculated for
of molecular hydrogen will cause fluctua- foreground-quiet region for beam throw
tions from point to point in the IR flux 5.60 • The solid curve denoted DC refers
which will be difficult to quantify. Com- to an average level offoreground, while
prehensive data on radiation from dust are the broken line (lAC') represents fluc-
available 80 far only at lOOp and problems tuations intrinsic to each of the fore-
arising with extrapolation back to lower grounds.
423

frequencies are similar to the problems with extrapolation forward for synchrotron radiation.
(Here, the m flux is assumed to vary as a Planck spectrum of characteristic temperature
Tdust modified by a factor vQ). Some variations of the IR background will also be caused
by changes in the general interstellar radiation field for different regions of the Galaxy. Our
estimate for the likely fluctuations is denoted 'AC', in Fig.5. The frequency dependence of
a genuine fluctuation at the 10-5 level is also shown.

4 Conclusions.

We have shown that the predicted anisotropies in GSR at 10 GHz are of the same magnitude
as those measured by the Jodrell Bank-IAC experiment and could in principle mimic the
cosmological signal. To avoid these contaminating effects future experiments must go to
higher frequencies. The best region for future searches lies around 60 GHz where the level
of background fluctuations falls to '" 4· 10-6 • More detailed studies of the fluctuations
in the Galactic foregrounds at a level '" 10-6 are required before detections of genuine
cosmological anisotropies can be claimed. .

Acknowledgements.AJB thanks the SERe for the provision of a research studentship. MG, BS
and JS thank the University of Durham and the Royal Society for hospitality. The authors thank
Dr A.N. Lasenby, Dr J.L. Osborne and Dr R.A. Watson for useful discussions.

References

Banday, A.J., Giler, M., Szabelska, B., Szabelski, J. & Wolfendale, A.W., 1991.Astrophys. J., in
press.
Bond, J.R. & Efstathiou, G., 1987. Mon. Not. R. astr. Soc, 226, 655.
Boulanger, F. & Perault, M., 1988. Al/trophYI/. J., 330, 964.
Davies, R.D., Lasenby, A.N., Watson, R.A., Daintree, E.J., Hopkins, J., Beckman, J., Sanchez-
Almeida, J. & Rebolo, R., 1987. Nature, 326, 462.
Lawson, K.D., Mayer, C.J., Osborne, J.L. & Parkinson, M.L., 1987. Mon. Not. R. astr. Soc, 225,
307.
Rebolo, R., Watson, R. & Beckman, J.E., 1989. Astrophys. Space Sci., 157, 333.
Szabelska, B., Szabelski, J. & Wolfendale, A.W., 1991. Journal of Physics G, in press.
Watson, R.A., 1989. Ph.D. Thesis, University of Manchester.
Watson, R.A., Rebolo, R., Deckman, J.E., Davies, R.D. & Lasenby, A.N., 1989. in Large Scale
Structure and Motions in the Universe, p. 133,00. Mezzetti et al. , Kluwer Academic.
MICROWAVE BACKGROUND ANISOTROPIES AND
LARGE SCALE STRUCTURE IN THE UNIVERSE

G. Efstathiou
Department of Physics,
University of Oxford, England.

ABSTRACT: Recent observations show that the Universe contains more structure
on scales ~ lOh- 1 Mpc* than expected in the standard cold dark matter (CDM)
theory of structure formation. This may indicate that the initial fluctuations were not
scale-invariant. I investigate how this would affect microwave background anisotropies
if the Universe is spatially flat with n = 1, as seems most natural if the Universe
went through an inflationary phase. If we are allowed arbitrary freedom in choosing
the shape of the initial fluctuation spectrum, then it is possible to construct models
which can account for observed structure in the galaxy distribution but which result in
small (l:!..T /T ;S 10- 6 ) anisotropies in the microwave background radiation on angular
scales ~ 50. However, it is very difficult to avoid producing temperature fluctuations
of order 10- 5 on angular scales of rv 10 , even if the Universe were reionised. If the
extra large-scale power detected in galaxy surveys really reflects clustering in the
mass distribution, then experiments at fJ rv 10 should soon see fluctuations in the
microwave background.

1. Power-spectra and Filter Functions

Figure 1 shows the power-spectrum Cl of the microwave background temperature


fluctuations expected in the standard scale-invariant cold dark matter model (Bond &
Efstathiou 1987). In the figure f2Cl is plotted against log(f) and I have superimposed
approximate filter functions, Fl, appropriate to a number of recent experiments. This
type of figure provides a very clear illustration of the angular scales probed by various
experiments and how well matched they are to the predictions of theoretical models.
Dick Bond and I are particularly fond of this way of looking at the temperature
fluctuations, but it may seem obscure to others so it is worthwhile trying to explain
it in some detail.
The temperature pattern on the celestial sphere can be expanded in spherical
harmonics

l:!..T/T= I>i}T(fJ, 4». (1)


l,m

* h is Hubble's constant in units of 100 kms- 1 Mpc- 1


425
T. Shanks et al. (eds.), Observational Tests ofCosmologicallnf/ation, 425-436.
© 1991 Kluwer Academic Publishers.
426

1-,
0.75 , ,
.
I \

:
I •

,..... SP: ~ OVRO


,..... II •I
0 I
'-' /
()
I")
'-' , I
I

"...
()
0.5
,
,
I
I
,

,
,

~ ,I •

,,
I
I

,,
I

0.25 ,,
TEN ,
\
,,
,,
I ,
,
\
\

O~~~~~~~~~~--~~~~--~~"~~~
10 100 1000 104
l

Figure 1. The dotted line shows f 2 Ct for the standard cold dark mat-
ter matter model with h = 1/2 and baryon density S1B = 0.1 plotted
against log(f). The power-spectrum has been divided by 3C(0), where
C(0)1/2 is the rms value of the temperature fluctuations and is equal to
3.6 x 1O- 5 /b for this model. The solid lines show the filter functions, Fi
for several recent experiments. The curve marked COBE shows a Gaussian
filter, Fi = texp(-(Osf)2) with Os = 7°, appropriate to the COBE DMR
experiment (Smoot et.al. 1991). The curve labelled TEN models the Tener-
ife experiment of Davies et.al. (1987) as a three-beam experiment (equ 7)
with a beam width Os = 4° and a beam throw 0 = 8°. The curve labelled
SP shows the approximate filter function of the South Pole experiment of
Meinhold & Lubin (1991) modelled as a three-beam experiment with a beam
width 08 = 13' and a throw of 0 = 1°. The curve labelled OVRO models
the Owens Valley experiment of Readhead et.al. (1989) as a three-beam
experiment with a beam width of 0.72' and a throw of 7' .

The power-spectrum is defined by the mean square value of the coefficients aT


(2)

and is related to the expectation value of the temperature autocorrelation function


427

by

The radiation correlation function is given exactly by equation (3). However, for
e ~ 1, the Legendre polynomials tend to Besselfunctions, Pt( cos 9) --+ J o ((e + 1/2) 9),
so (3) can be approximated as

(4)

i. e. for high order harmonics (small angles) we can approximate the celestial sphere
as a flat surface so the temperature autocorrelation function can be written as the
two-dimensional Fourier transform of the power spectrum Ct.
Consider now a two-beam experiment in which the mean-square temperature
difference is measured between two beams of width 98 separated by angle 9. The
expectation value of the mean square fluctuations can be written in terms of the
correlation function (3)

(5)

and so we can write (5) as a sum over the power-spectrum

(6)

where the filter function is

(7)

if the beams are approximated as Gaussians with width 9•. For a three-beam exper-
iment in which the temperature difference is measured between a field point Tf and
the mean temperature in two patches separated by an angle 9 on either side, the filter
function is more complicated,

Ft = [3/4 - Pt(cos9) + 1/4Pt(cos29)]exp( -9~e2). (8)


Generally, then, we can estimate the expectated mean-square temperature dif-
ferences for any particular experimental configuration by evaluating the sum (6) with
the appropriate filter function Ft. This is why Figure 1 is so useful, since from (6)

so multiplying l,2CI. by FI. and summing the area in Figure 1 gives a measure of the
expected mean-square temperature difference for the experiment. We can therefore
see immediately from Figure 1 that the Meinhold & Lubin (1991) experiment is
428

particularly well matched to the standard CDM model because it samples the large
peak in l2Cl at l'" 200. The COBE and Tenerife experiments sample regions of the
power spectrum where l2Cl ~ constant, i.e. angular scales where the temperature
pattern is scale-invariant and the anisotropies are caused by primordial potential
fluctuations (the Sachs-Wolfe (1967) effect). The Owens Valley experiment samples
very small angular scales where the power-spectrum has a low amplitude because of
the finite width of the last scattering surface.
Another way of seeing how different experiments sample the temperature pattern
is illustrated in Figure 2 which shows the expected rms temperature fluctuations for
two- and three-beam experiments with a beam throw of twice the beam width. This
shows that the signal peaks for beam throws of '" 30' - 40', i. e. on scales similar to
those of the Meinhold & Lubin experiment.

0.1

(8. /orcmin)

Figure 2. RMS temperature fluctuations for the CDM power-spectrum


shown in Figure 1 plotted against Gaussian beam width 911 for two-beam
(solid line) and three-beam experiments (dotted line) with a beam throw
() = 2(}•. The dashed line shows an amplitude of 10-5 lb.

We can use these figures to see how various experiments sample physical length
scales. We specialize to spatially flat models with n = 1 and define a conformal time
T = J dtla, where a is the cosmological scale factor. In the matter dominated era T
is related to redshift by T = 2(aoHo)-1(1 +z)-1/2. The physical wavenumber defined
by the Hubble radius (ct) at the time of recombination (z '" 1000) is given by

Arec '" (k rec I ao )-1 -_ -; Ho 1 + Zrec )-1/2


1 ( 2c ) ( (9)
429

where k ree is defined by kreeTree = 1r. The l'th harmonic in the temperature power-
spectrum samples wavenumbers of about k '" lITo, thus the physical scale sampled
on an angular scale 8 '" i-I radians is roughly

(10)

One should be cautious in applying equation (10), however, because wavelengths


smaller than ,xl do contribute to the anisotropies on angular scales i-I because the
temperature power-spectrum involves a projection onto the celestial sphere of the
three-dimensional plane-waves describing perturbations in the Universe. (In Section
3 an example is given where the correspondence between spatial and angular scales
implied by equation (10) fails). Nevertheless, if we compare equations (9) and (10)
with Figures 1 & 2, we see that the 1 0 experiment samples scales about equal to
the Hubble radius at recombination (if this occurs at the usual redshift of z '" 1000)
which are comparable to the largest scales on which structure has been observed in the
Universe ('" 50h- 1 Mpc see Section 2). The temperature anisotropies on these and
smaller scales depend on physical processes occuring at the time of recombination
(Thomson scattering from moving electrons in the case of the peak at l '" 200 in
Figure I) and so their amplitude depends on the matter content of the Universe (e.g.
the baryon density) and the history of recombination. If we want to sample potential
fluctuations directly then we must look on larger angular scales, () ~ 50 (l;S 10). On
these scales, the temperature power-spectrum is given by

(11)

where h is the trace of the metric perturbations. If the primordial fluctuation spec-
trum is a power law Ihl 2 ex: k n (n = 1 for scale-invariant fluctuations) equation (11)
gives,

(12)

if n < 3 and so a fit to observations of the low-order spherical harmonics could give a
direct measure of the spectral index n and its amplitude. To get a feel for the numbers
involved, the standard CDM model (i.e. scale-invariant adiabatic fluctuations) gives
Q = (5C2/41r}1/2 ~ 4 X 1O-6 h- l /b, where b is a 'biasing' factor which is thought to
lie in the range 1- 2 (see Kaiser & Lahav, 1989). (This factor is defined as b= l/ui,
where ui is the variance in the mass distribution at the present epoch in spheres
of radius 8h- 1 Mpc; b = 1 if galaxies are clustered like the mass distribution and
b > 1 if galaxies are more strongly clustered than the mass). The RELICT 1 limit for
scale-invariant spectra is Q < 1.5 X 10-5 at the 95% confidence level (Klypin et.al.
1987) and the present limits from COBE on [(21+ I}Cd41r]l/2 are about 5 X 10-5 for
l;S 20. These limits are above the theoretical predictions even if b '" 1 and h '" 1/2,
though not by much. It is easy to see why the theory predicts numbers of order 10-5 ,
for on large scales t1T IT tv (t1<jJ h.J c2 , where (t1<jJ h are the potential fluctuations on
scales ,xj if the potential fluctuations are scale invariant (t1<jJ = constant), from the
v;
scales of galaxy clusters to arbitrarily large scales, then t1TIT '" I c2 '" 10-5 since
the typical internal velocity dispersions of clusters are Ve tv 1000 km s-l.
430

However, the temparature fluctuations on angular scales ~ 5° sample length


scales of ~ 600h- 1 Mpc where we have no direct ob"ervational evidence for any large-
scale structure in the Universe. H we allow the primordial fluctuation spectrum to
have an arbitrary shape, the temperature anisotropies on large angular scales could be
considerably smaller than '" 10- 5 without conflicting with the idea that structure in
the galaxy distribution arose by gravitational instability. This point will be discussed
in more detail in Section 3.
Experiments probing angular scales :5 2° sample perturbations on physical length-
scales where density perturbations in the Universe are known to exist. Bond et.al.
(1991) have shown that the Meinhold & Lubin and Owens Valley experiments provide
interesting constraints on standard CDM models (assuming no reionization) and in
fact rule out models with OB ~ 0.2 if b :5 1 and h '" 1/2. However, recent observa-
tions of large-scale structure (Maddox et.al. 1990, Efstathiou et.al. 1990, Saunders
et.al. 1991) indicate more structure on large-scales than expected in the standard
CDM model. We investigate some of the implications of these results in the next
Section.

2. Non-Scale Invariant Initial Spectra.

In the standard CDM picture, the present day power-spectrum of the mass fluc-
tuations in linear perturbation theory is given by

P(k) oc kl[l + (ak + (bk)3/2 + (ck)2)"]2/", (13)

if OB ~ 0 0 , where a = 6.4(rioh 2 )-1 Mpc, b = 3.0(Ooh2)-1 Mpc, c = 1.7(Ooh2)-1 Mpc


and v = 1.13 (Bond & Efstathiou 1984). Note that if the primordial fluctuation
spectrum was scale invariant, the shape of the power-spectrum depends only on the
value of Oh 2 • An extensive series of numerical simulations of the CDM model has
been carried out by Davis et.al. (1985) and others showing that the theory can
provide a good match to observed structure on scales:5 lOh- 1 Mpc. On larger scales,
the theory apparently conflicts with observations of galaxy clustering, most notably
with the angular correlation function measured in the APM Galaxy Survey (Maddox
et.al. 1990).
One way out of this difficulty might be to drop the CDM model, replacing it with
a quite different type of theory (e.g. cosmic textures and hot or cold dark matter,
see Turok 1989). However, the basic ideas behind the CDM model have proved so
attractive that it seems reasonable to think of ways in which the standard model can
be modified to explain the new data. There seems to be no shortage of possibilities.
For example, Efstathiou et.al. (1990) have shown that low-density CDM models in
a spatially flat universe can provide an excellent description of galaxy clustering and
peculiar velocity fields if Oh '" 0.15. These models are compatible with two key
features of inflationary theories: the Universe is made spatially flat by invoking a
positive cosmological constant Ao = 3(1 - Oo)H~ (see Peebles 1984) and the initial
fluctuations are assumed to be scale-invariant. Of course one pays a price, there is
no reason why the cosmological constant should take such a special and theoretically
unnatural value ('" 2 X 10-120 m!,).Bond & Efstathiou (1991) have shown that CDM
models with a 17 keV neutrino (Simpson 1985, Hime & Jelley 1991) can produce
excess large-scale power if they decay on a timescale T d '" 1 year. Other long shots
include CDM models with a high baryon density, OB '" 0.2, and models with one or
more light (OeV)neutrinos.
431

Carlberg & Couchman (1991) have suggested that the standard CDM model can
be made to work if the biasing factor is low b ,..., 0.8. With such a small biasing
factor, the matter fluctuations on scales,..., 10 - 40h- l Mpc are of order unity and
they suggest that non-linear gravitational clustering alters the power-spectrum from
the form. of equation (13) to provide an acceptable fit to the observations of galaxy
clustering. They argue that the peculiar velocities between galaxies display 'velocity
bias', i.e. they are significantly lower than the relative peculiar velocities of the mass
distribution and so are compatible with observations. It is extremely important to
check whether their model can work, for it would provide a neat solution to many
of the problems posed by the observations. It would also imply that we ·should see a
detection of microwave background anisotropies if the experiments can be improved
by a factor of two or so in sensitivity.
Another possibile explanation of the observations might be to retain standard
CDM and simply argue that the biasing parameter is scale dependent. Nobody has
yet proposed a compelling (or even reasonably plausible) physical model in which
this could occur on scales as large as '" 40h- l Mpc. Nevertheless, we should remain
acutely aware of this possibility and strive to improve methods of measuring the mass
fluctuations directly, e.g. large-scale streaming motions (see Rubin and Coyne 1989
and references therein), Il.TIT, and the promising new idea of Blandford d.al. (1991)
and Miralda-Escude (1991) who suggest measuring the correlated elongations of faint
distant galaxies caused by gravitational lensing by intervening matter.
In the rest of this Section, I will explore another possibility in some detail. The
'natural' prediction of inflationary models is a scale-invariant initial spectrum of adi-
abatic fluctuations (see e.g. Bardeen et.al. 1983). However, by invoking multiple
scalar fields, it is possible to produce non-scale invariant spectra of a variety of differ-
ent shapes by fine-tuning the parameters (see e.g. Kofman, 1986; Kofman & Linde,
1987; Salopek et.al., 1990). Suppose then, that we retain all of the basic features
of the standard CDM model but allow deviations from a scale-invariant initial fluc-
tuations. We already have an idea of one shape of spectrum that provides a good
description of the observations, namely equation (13) with (}h = O.lS. To get this
shape of spectrum in an (} = 1 universe with h = 1/2 we need to multiply a scale-
invariant spectrum by a function of A: -

(14)

The function F( A:) is plotted as the solid line in Figure 3. This requires a model for
generating initial fluctuations which produces a step to higher amplitudes on length-
scales ..\ '" A:-l ~ SOh- l Mpc. The initial fluctuation spectrum is scale-invariant
on small scales and is approximately scale-invariant on large scales, though with a
higher amplitude. As we will show below, this would have drastic consequences for
microwave background anisotropies on large angular scales.
We have already remarked that there is little observational evidence * for struc-
ture in the Universe on scales ~ SOh-lMpc so perhaps the function F(A:) is overkill.
Suppose we truncate the fluctuation spectrum, i.e. P(A:) = 0 for A: :5 kc • The maxi-
mum value of kc that is compatible with observations of galaxy clustering turns out
to be k;l = SOh-lMpc. This is illustrated in Figure 4 which shows the angular
correlation functions (w( 8» from the APM Galaxy Survey scaled to the depth of the

* Though see Tully 1986, 1987, for tantilising evidence of structure in the distri-
bution of rich clusters of galaxies on ultra-large scales,..., 300h- l Mpc.
432

12

10 10· (b)

8 1000 /'
/'
"......,
~
"......,
~
/
u.. 6 a.. 100 /
/ COM
4 10
:8
2

0 0.1. 3
10-· 10-3 0.01 0.1 10 10- 10- 0.01 0.1 10
k/h Mpc- 1 k/h Mpc- 1

Figure 3. In Figure 3(a) the solid line labelled (A) shows the function F(k)
of equation (14) which gives a non-scale invariant spectrum that can repro-
duce observations of large-scale structure in the galaxy distribution. The
dotted line marked (B) shows the maximum value of the cut-off wavenum-
ber k;l = 50h- 1 Mpc compatible with the observations (see Fig. 4). The
power-spectra of models (A) and (B) are compared with standard CDM
(0 = 1, h = 1/2) in Figure 3b. The power-spectra are normalized so that
O'~ = 1.

Lick catalogue (see Maddox et.al. 1990 for further details). The curve marked CDM
shows w( 8) for the standard CDM model computed from the linear theory power-
spectrum normalised so that 0'1 = 1 (shown as the dashed curve in Figure 3b). The
solid curve marked (A) shows what happens if we multiply the standard CDM power
spectrum by the full ramp F2(k) and the dotted curve labelled (B) shows the result
of truncating the ramp on scales ~ k;l = 50h- 1 Mpc. Models A and B provide
acceptable fits to the APM result (and also to the IRAS redshift survey). If we
truncate F(k) on scales smaller than k;l = 50h- 1 Mpc, w(8) falls off too sharply
at 8 ~ 5° and so the fit to the observations become unnacceptable. The power-
spectra of models A and B are compared to standard CDM in Figure 3b. Model
B is a reasonable representation of what we know about large-scale structure in the
galaxy distribution and shows clearly that there i" no direct ob8ervational evidence for
power on 8cale8 k- l ~ 50h- 1 Mpc. In fact, the sharp cut-off of model B is unrealistic
because gravitational clustering guarantees that the power-spectrum cannot fall off
more steeply than P(k) oc k4 (Peebles, 1974) but this effect is unimportant in what
follows.
433

0.1

" \
0.01 \
\
COM \
\
\
\

10
9 (degrees)

Figure 4. The dots show estimates of the angular correlation function


w(B) for galaxies in the APM Galaxy Survey (Maddox et.al. 1990). These
estimates have been scaled to the depth of the Lick catalogue where 10
corresponds to a spatial scale of '" 5h- 1 Mpc. The dashed line shows the
linear theory prediction for standard CDM. The other two curves show w( B)
for the two non-scale invariant spectra (models A and B) plotted in Figure
3.

3. Microwave Background Anisotropies from Non-Scale Invariant Spectra.

Table 1 lists values for the rms temperature fluctuations for the experimental
configurations described in Section 1. It also lists the rms temperature fluctuations
..jC(0). We have assumed h = 1/2 for each model and the numbers in brackets list
the baryon density {lB.
As expected, Model A exceeds the quadrupole and the Meinhold & Lubin (1991)
limits at B = 10 unless b > 1. Model B produces much lower values for fluctuations at
B ~ 50, down by factors of'" 30 from the values for Model A and well below the COBE
upper limits. This is a direct result of the lack of large-scale power caused by the
cut-off in P(k) at k :::; kc • In models B, the low-order coefficients of the temperature
power-spectrum are no longer given by equation (12). Instead, Ct ~ constant for
l ;S 100, i.e. the large-angle experiments are simply measuring the shot noise from
lumps in the microwave background with sizes'" 0.5 0 and so the signal predicted
is down by a factor of '" VN, where N is the number of lumps in the beam. The
amplitudes for these large angle experiments are thus determined by smaller scale
structure and the entries for Model B in Table 1 are close to the minimum possible
434

values if structure arose by gravitational instability.

Table 1: Computations of b(t:1T/T)

Expt () CDM(O.l) A(O.l) B(O.l) B(0.03) B(O.Ol)

vtc(O) 0 4 x 10-5 9 x 10-5 5 x 10-5 3 x 10-5 3x 10-5


OVRO 7' 8 x 10-6 8 X 10- 6 8 X 10-6 6 X 10-6 4X 10-6
SP 1° 2 x 10-5 3 x 10-5 3 x 10-5 2 x 10-5 16 X 10-6
TEN 8° 4 x 10-6 1 x 10-5 2 X 10-6 1 X 10-6 1X 10-6
COBE 7° 1 x 10-5 3 x 10-5 1 X 10-6 7 X 10-7 6X 10- 7
Q 1r/2 4 x 10-6 2 x 10-5 6 X 10-7 3 X 10-7 3X 10- 7

However, the most striking results in Table 1 are the consistently high amplitudes
predicted for the Meinhold & Lubin experiment at () = 1° . The predictions for
Model B with OB = 0.1 are just about compatible with the present upper limit
(t:1T /T .:s 2 x 10-5 ). The anisotropies on angular scales .:s 1° can be reduced by
lowering OB, so Model B is acceptable if OB .:s 0.1 and the biasing parameter is b ~ 1.
The situation is not much improved if the Universe were reionized - if OB = 0.03 and
the inter-galactic medium never recombined, Model B gives b( t:1T /T)rmB = 8 x 10- 6 •
From this analysis we conclude that if we allow the primordial fluctuation spec-
trum to have an arbitrary shape we can produce models which predict relatively
small anisotropies in the microwave background on scales ~ 5°. If COBE and other
large-angle experiments fail to detect fluctuations even at the few x 10- 7 level, we
are not forced to abandon gravitational instability as a mechanism for forming large-
scale structure in the Universe though we would have to discard the idea of a scale-
invariant spectrum in a spatially flat universe. However, if the large-scale structure
seen in galaxy surveys reflects real clustering in the mass distribution we must expect
fluctuations of order 10-5 on angular scales of '" 1°. The existing limits already
force us to a low baryon density OB .:s 0.1. The sensitivities of intermediate angle
experiments are likely to improve by significant factors over the next year or two -
and so they should detect primordial fluctuations. If they fail to do so, we may face
a real crisis in Cosmology.

Acknowledgements: I thank Dick Bond for many interesting discussions over the
last few years and Will Sutherland for suggesting that I check my normalizations.
This article was stimulated by conversations with Dick Bond, Phil Lubin and Stephen
Meyer on Venice Beach during a conference organised by UCLA. I am grateful to
UCLA for providing travel support.

REFERENCES
Bardeen, J.M., Steinhardt, P.J. & Turner, M.S., 1983, PhY3. Rev. D, 28,679.
Bond, J .R. & Efstathiou, G., 1984, Ap. J., 285, L45.
Bond, J.R. & Efstathiou, G., 1987, M.N.R.A.S., 226,665.
Bond, J .R. & Efstathiou, G., 1991, Phys. Lett. B, submitted.
Bond, J.R., Efstathiou, G., Lubin, P.M. & Meinhold, P.R., 1991, Phys. Rev. Lett.,
in press.
435

Blandford, R.D., Saust, A.B., Brainerd, T.G & Villumsen, J.V., 1991, preprint
Carlberg, R.G. & Couchman, H.M.P., 1991, preprint.
Davies, R.D., Lasenby, A.N., Watson, R.A., Daintree, E.J., Hopkins, J., Beckman,
J., Sanchez-Almeida, J. & Rebolo, R., 1987, Nature, 326, 462.
Davis, M., Efstathiou, G., Frenk, C.S. & White, S.D.M., 1985, Ap. J., 292, 371.
Efstathiou, G., Sutherland, W.J. & Maddox, S.J., 1990, Nature, 348, 705.
Efstathiou, G., Kaiser, N., Saunders, W., Lawrence, A., Rowan-Robinson, M. Ellis,
R.S. & Frenk, C.S., 1990, M.N.R.A.S., 247, lOp.
Hime, A. & Jelley, N.A. 1991, preprint.
Kaiser, N. & Labav, 0., 1989, M.N.R.A.S., 237,129.
Klypin, A.A., Sazhin, M.V., Strukov, LA. and Skulachev, D.P. 1987, Soviet. Astr.
Lett., 13, 104.
Kofman, L.A., 1986, Phys. Lett., 173B, 400.
Kofman, L.A. & Linde, A.D., 1987, Nucl. Phys., B282, 555.
Maddox, S.J., Efstathiou, G., Sutherland, W.J. & Loveday, J., 1990, M.N.R.A.S.,
242, 43p.
Meinhold, P.R. & Lubin, P.M., 1991, Ap. J. Lett. submitted.
Miralda-Escude, J., 1991, Ap. J. submitted.
Peebles, P.J.E., 1974, A. & A., 32, 391.
Peebles, P.J.E., 1984, Ap. J., 284,439.
Readhead, A.C.S., Lawrence, C.R., Myers, S.T., Sargent, W.L.W., Hardebeck, H.E.
& Moffet, A.T., 1989, Ap.J., 346, 566.
Rubin, V.C. & Coyne, G.V., 1988, Large-Scale Motions in the Universe, Princeton
University Press, Princeton.
Sachs, R.K. & Wolfe, A.M., 1967, Ap. J., 147,73.
Salopek, D.S., Bond, J .R. & Bardeen, J .M., 1990, Phys. Rev. D, 40, 1753.
Saunders, W., Frenk, C.S., Rowan-Robinson, M., Efstathiou, G., Lawrence, A.,
Kaiser, N., Ellis, R.S., Crawford, J., Xia, X-Y., & Parry, I., 1991, Nature,
349,32.
Simpson, J.J. 1985, Phys. Rev. Lett. 54, 1891.
Smoot, G. et. al. 1990, COBE preprint No. 90-07.
Tully, R.B., 1986, Ap. J., 303, 25.
Tully, R.B., 1987, Ap. J., 323, 1.
Turok, N., 1989, Phys. Rev. Lett., 63, 2625.
436

DISCUSSION:

Starobinsky: The maximal relative height of your truncated power spectrum ( '" 3.5 times
in terms of amplitude of perturbations, as I see from your figure) is in essential agreement
with the spectra I proposed in my Monday talk which all have the maximal relative height
3 - 3.5 times as compared to the standard CDM spectrum at small scales.
Efstathiou: I wish I had been able to attend your talk on Monday! It is clear that we
can construct models which produce non-scale invariant spectra. My own feeling, however,
is that these models are "unnatural" because they require fine tuning. Clearly, we need a
deeper understanding of the physics of the early universe.
Sanz: I would like to comment that the generation of anisotropies due to non-linear
perturbations could be important. For an isolated structure like the Shapley concentration
you can have ¥ '" 10-5 due to the integrated effect if one assumes M ~ 10 17 M0. For
models of galaxy formation, we are currently involved in the calculation of such effects to
see to what extent they might dominate the linear ones (i.e. Sachs-Wolfe and the Doppler
Shift due to the last scattering).
Efstathiou: Your mass estimate seems high, but I agree with your general point. I am
predicting rms values using linear perturbation theory. If we assume further that the linear
fluctuations obey Gaussian statistics, then we have fully specified the temperature pattern
in linear theory. However, there will also be anisotropies caused by non-linear structures -
in addition to the effect you mention, there will be (non-Gaussian) anisotropies caused by
the Sunyaev-Zeldovich effect, discrete radio sources, dust etc.
DISCOVERY OF THE SMALL SCALE SKY ANISOTROPY AT 2.7cm:
RADIO SOURCES OR RELIC EMISSION?

Yu.N.PARIJSKIJ, B.L.ERUKHIMOV, M.G. MINGALIEV, A.B. BERLIN,


N.N. BURSOV, N.A. NIZHELSKIJ, M.N.NAUGOLNAJA, V.N. CHERNENKOV,
O. V. VERKHODANOV, A. V. CHEPURNOV, A.A. STAROBINSKY
Special Astrophysical Observatory
957147, Niznij Arkhyz, Stavropol Territory
USSR

ABSTRACT. We present new limits on the small scale 3K emission anisotropy. At the scales
of 1° -:- 3.°3 all 7 harmonics have at A = 7.6cm amplitudes below 10- 5 in f1T/T. At I from
6.1 .103 to 17.5 .103 (rad- 1 ) scale this limit is about 2.8 .10- 4 at a wavelength of 2.7cm and may
be fully explained by a discrete source contribution. The low baryonic content of the Universe and
present limit on f1T/T at the horizon scale suggest that (f1M/M)gravitating « (f1M/M)baryonic.
We speculate that 3K anisotropy measurements demonstrate a) the existence of the dark matter,
b) the nonbaryonic nature of this dark matter and c) that the spatial distribution of this matter is
much more homogeneous than that of the visible one.

1 Introduction

All our attempts (from 1968 onwards) to observe the theoretically predicted anisotropy in
the 3K emission are summarized in Parijskij et al. 1986. In this paper we present new obser-
vations, using specially developed multi-frequency cryogenic receivers and data acquisition
system at the RATAN-600 radio telescope.

2 Strategy

We have tried to decrease the problems in the COLD-80 experiment (Parijskij et al., 1986)
due to the confusion effect. The main wavelength in the present COLD-90 experiment was
chosen to be 2.7 cm (instead of the 7.6 cm used earlier) which decreased the confusion
noise by a factor of about 20. To get the appropriate level of thermal sensitivity we chose
a circumpolar region (instead of the equatorial one in the COLD-80 experiment) for the
transit mode of observation.

437
T. Shanks et al. (eels.), Observational Tests of Cosmological Inflation, 437-441.
@ 1991 Kluwer Academic Publishers.
438

3 Receivers

In the COLD-SO experiment 31cm, S.3cm, 7.6cm, 3.9cm, 2.1cm and 1.3Scm receivers, were
utilised with 7.6cm as the main wavelength. In the present COLD-90 experiment we have
improved the 31cm receiver's sensitivity by a factor of2, improved the stability ofthe 7.6cm
receiver, developed a new cryogenic receiver at 2.7cm with sensitivity 5mK/s1 / 2 , improved
the 1.38cm receiver's sensitivity by a factor of 3, and installed a new three-frequencies
receiver block at 6cm, 2.7cm and 1.0cm wavelengths with sensitivities 5,7, and 12 mK/s1 / 2 ,
respectively. At 31cm and 7.6cm only single beam observations were possible; at the other
wavelengths the beam switching mode could be used with some freedom in positioning the
beams on the sky.

4 Data reduction

R.m.s. fluctuations in the average scan (about 60 daily scans) were about OAmK without
any cleaning except for a background subtraction with a window of 200 sec width. We
divided all the scans into two independent groups, averaged over all the scans in each
group, convolved both results with the central beam cross section and looked for the r.m.s.
amplitude of the coherent signal which existed in both groups after cleaning them at some
fixed level (to find the weak coherence signal we used the "Jack-knifing" method of Miller,
1968). Then the result was averaged over different scan groupings. This was done for
different cleaning levels down to the 3u-Ievel of about 1.1mK. If the coherent signal does
not depend on the cut-off level, it might be 3K-anisotropy noise well below the receiver
noise. In the opposite case it may be the confusion noise.
To estimate the confusion noise we used source counts collected by Franceschini et
al. (1989) extrapolated to 2.7cm with the source populations model presented in the same
paper. The computer simulation included generation of two different scans, with different
noise contributions, evaluation of the above described algorithm, and the averaging of the
results over different signal realizations.

5 Results

Here only preliminary results will be discussed connected with the smallest beam spacing
at 2.7cm and with new data reduction of the 7.6cm COLD-SO data.

5.1 7.6cm Observations

Using COLD-80 data (see Parijskij, Korolkov, 19S6 for the details) in the 0°.5 -:- 5° scale
interval we calculated the Fourier spectrum of the coldest part of the strip centered at
SS433 declination after subtraction of the correlated noise at 31cm and at 2.1cm (Galaxy,
standard spectrum sources and water vapour emission). The result is shown in Table 1.
439

TABLE I

Scale, degrees 1.1 1.25 1.43 1.67 2.0 2.5 3.3


Amplitude, !1T/T .106 4 2 7 6 4 8 7

5.2 2.7 Observations

At this wavelength the resolution of RATAN-600 is limited by the proper interval 1/2 . >./D
= 7.5 arcsec with FHPW = 18 arcsec in right ascension and by FHPW = 12 arcmin in
declination. The separation between the two beams in the discussed group of observations
was 40 arcsec. The resulting data discreteness interval was 10 arcsec. We subtracted the
background using a 200 arcsec window to reduce large-scale interference and then convolved
the data with the central beam cross section. All this affected the spectral sensitivity.
It results in the fact that only the information at the scales between I =
6.1 . 103 and
1= 17.5.10 remained (the limits correspond to half-maximum spectral sensitivity).
3
We accumulated data in the transit mode with fixed position of the beam at DEC.
86°14' (1950) and at 2h < R.A. < 22h for two months in the spring of 1990. At this
declination a transit time of 1 arcsec is about 1 sec of time.
We shall summarize the results of 2.7cm experiment at the present stage of data reduc-
tion:

1. The presence of a coherent signal in two independent groups of observations was


established.

2. The amplitude of this signal depends on the cleaning level, being smaller at a lower
cleaning level.

Table II illustrates these statements

TABLE II. Coherent signal and cleaning level

Cleaning Experiment Computer simulation


level signal 1 -(1 error signal 1 -(1 error
mK mK mK mK mK
3.400 0.171 0.027 0.143 0.017
1.600 0.127 0.011 0.115 0.009
1.100 0.088 0.017 0.098 0.012

We see that the experiment shows practically the same level of the coherent signal as
was expected from the simple LogN-LogS model.
Our observations do not contradict the zero level of small scale anisotropy in the re-
sulting spectral window. After careful analysis we estimate an extra noise no more than
about 20 pK in antenna temperature (1-(1 error), that is 7.10- 6 in oT.. /Tbb where T .. -
antenna temperature and Tbb - Black Body temperature 2.78K. This difference between
the experiment and the expected contribution from the discrete sources may be just due to
uncertainties in the source counts. To convert 20pK in oT.. into the brightness temperature
440

fluctuations we should take into account the spectral sensitivity function affected by the
mentioned factors. After this correction we found that our 20jlK in Ta corresponds to 2.8
10- 4 in 6Tbb/Tbb when I is between 6.1.103 and 17.5 .103 • This limit is close to the best
earlier estimate at these scales (Hogan et al., 1989).
The main new result of the first step of realization of our program is the direct check of
the validity of Franceschini at al, 1989 model of the LogN-LogS curve at low flux densities
and short wavelengths.

6 Discussion

We believe that there is no positive detection of anisotropy of the 3K emission at about


the 10-5 level at the few degrees scale, the upper limit being close to the 10- 5 . For CDM
models with standard Harrison-Zeldovich initial spectrum it follows from Table I, that
the biasing parameter, bp > 0.5 at 20- level (this corresponds to e > 4.10- 5 where e 2 c4 =
dP<p/dlnk, see Starobinsky (1987) for details). Inflation suggests that n = 1. Visible matter
gives n = 0.1. Nuclei synthesis suggests n $ 0.035 . h- 2 (h = Ho/100km/s/Mpc); small
scale anisotropy measurements suggest an even smaller value, n - 0.008 for the baryonic
content of the Universe (Fukugita et al., 1989). Visible matter shows the void structure
with I:!.M/M about 0.5 at the scales under consideration. The Sachs-Wolfe effect predicts
I:!. T /Tr:::. I:!.M/M. To be below 10- 5 in I:!. T /T gravitating matter has to be distributed much
more homogeneously than the visible one. Anisotropy measurements demonstrate that all
models with the Universe filled by visible matter contradict the observations. Here we stress
that anisotropy measurements show that the invisible matter should have a nonbaryonic
nature (at least, if reionization is absent) and that it cannot be distributed in the halos of
the visible galaxies (or clusters).
Very small scale anisotropy can tell us about the second ionization epoch and velocity
dispersion of the protogalactic objects.
With the present sensitivity we were able to use our results only to check the behaviour
of the source counts at an unexplored level of flux densities at 2.7 cm. Good agreement
with the model of Franceschini et al. (1989) suggests that the new flat spectrum population
of blue galaxies playa definite role at short cm-wavelengths in the 100jlJy - few mJy region.

References

Franceschini A., Toffolatti L.,Danese L.,& G. De Zotti, 1989. Astrophys. J., 344, 1.
Fukugita M. & Umemt'>ra M., 1989. Astrophys. J., 339, 1I-L4.
Hogan C.J. & Partridge R.B., 1989. Astrophys. J., 341, L29.
Miller R.G., 1968. Ann. Math. Statistics, 39, 567.
Parijskij Y.N.& Korolkov Dv., 1986. Ap. Sp. Sci. Rev., seT.D, 5, 40.
Starobinsky A., 1987. Soobshchenia S.A.O. USSR, 53, 57.
441

DISCUSSION:

Lasenby: How is the "cleaning" carried out which enables the removal of the galactic
contribution, and also sources, in order to set the very low limits on tl.T/T.
Parijskij: We used 6 wavelengths simultaneously, from 31cm to 1.38cm. Quite unexpect-
edly we discovered strong fluctuations on the sky at 7.6, 8.3 and 31cm with non-thermal
spectrum, and 1mK r .m.s. amplitude at 7.6 cm on a scale 1 deg - 2 deg. Using the T -T
plot method we subtracted with the proper weight the 31 cm scan from 7.6 cm scan and
decreased the dispersion by a factor close to 10. The rest of the 7.6 cm. scan was well
correlated with atmospheric emission, which dominated at 2 and 1.38cm. Again we sub-
tracted the 2cm scan from the rest of the 7.6cm and gained another factor of 3. Then we
used a correlation and spectral analysis, see Parijskij et al., Space Sci.Rev. ser.D, vol.S,
p.40, 1986. The confusion level after 20' cleaning was 0.1 mK on a 1 arcmin scale.
BALLOON-BORNE OBSERVATIONS OF CMB ANISOTROPIES AT INTERMEDIATE
ANGULAR SCALES, AT SUBMM AND MM WAVELENGTHS

P. DE BERNARDIS, S. MASI, B. MELCHIORRI* & F. MELCHIORRI


Dept.of Physics, University of Rome * IFA-CNR, Rome
Piazz.le Aldo Moro 2, 00100 Rome, Italy

ABSTRACT. We have perfonned extensive observations of the cosmic microwave background (CMB) at
angular scales of 6-17 degrees in the wavelength range 500 - 3000 micron, by balloon-borne instrumenta-
tion. We have found an upper limit for the square root of the amplitude of the correlation function ranging
from :5 2 x 10- 5 for primordial CDM perturbations with spectral index n=O down to :5 0.8 x 10- 5 for
n=2.9, at 95% of confidence level, after correction for residual atmospheric fluctuations and galactic dust
emission.

1. Introduction

Various authors have claimed in the past marginal detections of CBM anisotropies at angular scales of
1-10 degrees and at level of 2-5 10- 5 (Melchiorri et al., 1981; Davies et al. 1989; Meyer et al., 1990).
If these detections will tum out to be real, they will represent a major step in observational cosmology,
by opening the possibility of observing the primordial universe at scales where the casual processes were
not operative. Unfortunately, other possible sources have been suggested for the observed anisotropies.
Residual atmospheric fluctuations and galactic emissions could be responsible for them.
Since 1980 we have started an observing program to search for CBM anisotropies by balloon-borne
multichannel far infrared photometers. In 6 balloon flights we have carried out a set of experiments
aimed to study the possible sources of spurious anisotropies. We believe to have identified a strategy of
observation suitable to disentangle the true cosmic anisotropies. We have already published partial results
of our experiments (Melchiorri et al., 1981; Ceccarelli et al., 1982; de Bemardis et al., 1984; Dall'Oglio
et al., 1985; de Bernardis et al., 1988) and a paper with a detailed description of our technique is in
preparation (Bottani et al., 1991). In the present pa~r we reasume the results of our analysis.

2. Observational Strategy

The basic idea is that of selecting the wavelengths and the bandwidths of operation of two or more
photometric channels following a rule that optimizes the correlation between "near infrared" and "far
infrared" channels. For sake of simplicity, let us assume that the "near infrared" channel is centered on
300 microns with a 20% of bandwidth: the signal S N I R will be constituted of four tenns: the atmospheric

443
T. Shanks et al. (eds.), Observational Tests ofCosmologicallnflotion, 443-446.
e 1991 Kluwer Academic Publishers.
444

residual anisotropy at balloon altitUdeSaNIR, the galactic anisotropy SgNIR, the extragalactic background
anisotropy due to many unresolved extragalactic sources in the field of view SexNIR, and, finally, the
instrumental offset, generally variable in time SofJ NIR.

SNIR = SaNIR + SgNIR + SexNIR + SofJNIR


The far infrared channel can be written down in a similar way, but adding a term due to the cosmic
anisotropies SeBM, i.e.,

SFIR = SaFIR + SgFIR + SexFIR + SeBM + SofJFIR


This being the situation our problem is that of selecting the wavelength and the bandwidth of the FIR
channel that minimize the difference SFIR -SNIR. The solution of the problem is unique once the spectra
of galactic, atmospheric and extra-galactic backgrounds are given. After several attempts it has led to the
choice of a large-bandwidth FIR channel (Bottani et al. 1991).

3. Observations

We will discuss here the observations obtained in a balloon experiment where a mutichannel photometer
has been employed (a detailed description of this experiment is in preparation).
The modulation amplitude was 6 degrees in the sky and the beam size of 5 degrees.

200

E-t o
<I

-200

150 200 250


Right Ascension (deg)

Figure 1. Plot of far infrared signals after subtraction for near infrared signals optimized to reduce the
variance of the data. The best fit of the residuals is consistent with a O.2mK dipole gradient, in agreement
both in direction and amplitude with COBE-DMR results, when projected along our line of sigth.
445

From about 20 hours of data we have extracted those referring to the cleanest regions in the sky, as
suggested by the lowest standard deviation of the recorded signals. When the instrument was pointing a
fixed direction in the sky we detected atmospheric fluctuations as signals changing with time scales of 1-5
minutes.

100

50

-....
.-...

'0
rn
as
::s
0
rn
Q)

-~
~
.........
-50
.........
E-<
<l
-100
150 200 250
Right Ascension (deg)

Figure 2. Plot of the residuals of figure 1 after subtraction for the dipole term. The error bars are
the standard deviations evaluated over three scans of the same sky zone. The temperature gradients are
thermodynamic

We observed galactic signals when the instrument was pointing toward infrared cirrus. These data allowed
to check for the correlation between NIR and FIR channels. The minimized difference between the two
channels is shown in figure 1, where the gradient due to the dipole anisotropy is clearly visible. We have
analized the residuals after subtraction for the dipole gradient to search for true cosmic anisotropies (figure
2). The result of our analysis is shown in figure 3, where we ploned the upper limits at 95 % of confidence
level for various spectral indexes of primordial perturbations.
446

I)
-2 o 2
spectral index n

Figure 3. Upper limits to CMB anisotropies produced by primordial fluctuations having the spectral index
given in abscissa (¥ )pr.m.s. refers to the r.m.s. fluctuations as measured by a pencil beam over the
sky. They are related to the observed ¥ via the correlation function.

4. Aknowledgments

This work has been made possible by the generous financial support of the Department of Physics of
University of Rome (MPI 60% Faculty Budget) and of IFA-CNR.

5. References

Bottani, S., de Bernardis P., Masi, S., Melchiorri, B., Melchiorri, F., 1991, in preparation
Ceccarelli, C., Dall'Oglio, G., de Bernardis, P., Masi, S., Melchiorri B., Melchiorri, F., Moreno, G.,
Pietranera, L., 1983, ApJ. (Letters), 275, L39.
Dall'Oglio, G., de Bernardis, P., Masi, S., Melchiorri, F., Moreno, G., 'frabalza, R., 1985, Ap. J.,289,
609.
Davies, R.,D., Watson, R., Daintree, E.;., Hopkins, J., Lasenby, AN., Beckman, J., Sanchez-Almeida, J.,
Rebolo, R., 1987, Nature, 326, 6112
de Bernardis, P., Masi., S., Melchiorri, B., Melchiorri, F., Moreno, G., 1984, Ap. J., 278, 150.
de Bernardis, P., Masi, S., Melchiorri, F., Moreno, G., Vannoni, R., Aiello, S., 1988, ApJ., 326, 941.
Melchiorri, F., Olivo, B., Ceccarelli, C., Pietranera, L.,1981Ap. J. (Letters},227, L129.
Meyer, S.,S., Cheng, E.,S., Page,L.,A., 1990. ApJ. (Letters), Oct. 1990
THE DURHAM/UKST GALAXY REDSHIFT SURVEY

Broadbent, A.t, Hale-Sutton, D.I, Shanks, T.t, Watson, F.G.2,4,


Oates, A.P.3, Fong, R.I, Collins, C.A.2, MacGillivray, H.T.2,
Nichol, R.2 & Parker, Q.A.2
1 Physics Dept., University of Durham, South Road, Durham, DHl 3LE.
2 Royal Observatory, Blackford Hill, Edinburgh, EH9 3HJ.
3 Royal Greenwich Observatory, Madingly Road, Cambridge, CB3 OHA.
4 A nglo-A ustralian Observatory, Coonabarabran, NSW2357, Australia.

ABSTRACT. We are currently engaged in a long-term project to make a red shift survey
of'" 4000 galaxies based on the Durham/ROE Southern Galaxy Catalogue (Collins et al.,
1988) which is a COSMOS survey of the 2D distribution of hi < 20.sm galaxies over 60
UK Schmidt Telescope (UKST) fields. The current programme makes further use of the
UKST in conjunction with FLAIR, a fibre-coupled spectroscopy system. The ultimate aim
is to map the 3D distribution of galaxies over all 60 high galactic latitude fields by making
a 1-in-3 random sampled redshift survey of bi.!S 16.75m galaxies. We are currently able to
observe 61 galaxies simultaneously per 5° X 5° field using the Durham dual CCD camera
coupled to 2 FLAIR spectrographs.
When complete, the galaxy redshift survey will enable us to map, out to a distance of
300h- 1 Mpc and in a continuous volume of 4 X 106 Mpc3, the distribution of superclusters,
filaments, sheets and voids whose presence has been claimed in other redshift surveys. The
high sampling rate over such a large volume will enable us to investigate the distribution
over a wide range of galaxy separations, 1 < r < 300h- 1 Mpc.
In Figures 1 and 2 are shown preliminary results from data taken earlier this year.

References

Collins, C.A., Heydon-Dumbleton, N., MacGillivray, H.T. & Shanks, T., 1988. in Large Scale Struc-
tures in the Universe, [AU Symp. No.130, p. 125, ed. Audouze, J. et al., Reidel.
Da Costa, L.N., Pellegrini, P.S., Sargent, W.L.W., Tonry, J., Davis, M., Meiksin, A., Latham, D.W.,
Menzies, J.W. & Coulson, I.A., 1988. Astrophys. J., 321, 544.

447
T. Shanks et al. (eds.), Observational Tests of Cosmological Inflation, 447-448.
© 1991 Kluwer Academic Publishers.
448

,
. ,... .•
-, ..-.. .. .. I.·,
\

.. .. ..
.-
\
: .. e'-:.:r:: ..
..• .• ... ... ....
\

:.. ,. \
21'3v"O'

20000 30000
velocity ( kms-' )

Figure 1. A coneplot of the galaxy distribution in UKST fields 466-471 and 409. Redshift lies in the
horizontal direction and RA in the vertical direction. The Dec of the galaxies are in the range -27.5° to
-32.5°. The dotted arc shows approximately the extent in redshift of The Southern Sky Redshift Survey
of Da Costa et 01. (1988) which covers this same region of sky.
RCI ~--r-r-"""' I , I ' I , , I ' .-r-.,-,--,-. r.-,,..-,'-r-T..-rr·

..,-

40

~ 30
z

40000

Figure 2. A histogram plot showing the distribution of velocities measured from survey galaxies lying in
fields 466-471 and 409.
TIME EVOLUTION OF LENSED IMAGE SEPARATIONS

Tom Broadhurst & Sebastian Oliver


Mathematics Department
Queen Mary fJ Westfield College
London UK

Abstract: Cosmic expansion results in a slow decrease in the angular separation of gravita-
tionally lensed images, which, if monitored, may provide a measure of the global geometry.
We calculate the time dependence of this effect for the simple case of a double image pro-
duced by a point mass in FRW universes. \Ve note that peculiar velocities of source and lens
have a negligible effect on the separation and we briefly discuss the feasibility of observing
this effect.

1 Introduction

The strong evolution of source populations has increasingly made clear the difficulty of using
the classical cosmological tests for placing interesting constraints on the global curvature.
For example, the predicted optical number counts - interpreted in the context of the FRW
models - differ negligibly for the range 0 < n < 1 (with A = 0) by comparison with the large
excess of objects observed at faint magnitudes. Prospects of a consensus on the value of
local expansion rate seem poor at present, and the classical tests are of course independent
of Ho.
A completely alternative approach to the cosmological problem is to monitor the time
evolution of redshiftl. The expansion of the universe produces a very slow decrease of
cosmological redshift with time which is directly related to the global geometry and in-
dependent of any evolution intrinsic to the source. The magnitude of this effect is given
by2

dz ~
dt = Ho(1 + z)(1- y(1 + z)) (1)

for the case n = 1 with A = O. This produces a 'drift' of ~ 1O- 7 .4.yr- 1 at optical
wavelengths for a source at z = 1 with Ho in the usual range.
Unfortunately this change is small compared with the broadening of a spectrum by
internal velocity dispersion, which is typically of order a few .4., thus requiring the centroid
of a spectral feature to be defined to better than one part in 10 7 • To measure this drift,
using current technology, sources would have to be monitored for a minimum of 103 yrs!

449
T. Shanks et al. (eds.J, Observational Tests of Cosmological Inflation, 449-451.
© 1991 Kluwer Academic Publishers.
450

2 Lensed Image Evolution

Cosmic 'drift' may be more readily detected by its effect on the separation of multiply
lensed images of a point source. In an expanding universe a gradual decrease of the angular
separation between images is produced by the steady increase of cosmological distances.
This method is analogous to monitoring the cosmic 'drift', but for the case oflensed QSO's,
the images are bright and point-like so observing this effect is more feasible. To calculate
the time dependence of angular separation between images we express affine distance,A, in
terms of redshift and relate this using eqn 1 to time. The dependence of affine distance on
=
redshift, for the case n 1, is given by 2,

2
A = -«1 + z), - 1)(1 + z)T
6 -5
(2)
5
The characteristic size of the lens a can be defined in terms of the affine distances to
source >-'s, and lens >-'L,

4GMHo (1 + ZL)(>-'S - AL)AL


a= c1 As
(3)

The angular separation between images can then be expressed as

0s 4)'
1

~0 = (b/a)2 + (4)
(b/a)2
where b is the impact parameter of the 'unperturbed' beam and 8s the corresponding
angle between source and lens.
Using equations 1-4 the partial differential equations needed to determine the time
dependence of ~0 can be determined. Figure 1 shows the magnitude of this effect with
values of b/a and the lens mass consistent with the well studied lens 0957+47. The source
is placed at Z. = 1.0 and the the curves indicate the dependance of the yearly change of
separation for the pair oflensed images for a lens placed in the range 0 < ZL < 1.0. The cases
of n = 0.1 and n = 1.0 (filled and empty beam) with A = 0 and Ho = 75Kms- 1 Mpc- 1
are plotted. Note that we require the lensing object to be detected so that, b and 0s can
be determined.

3 Discussion

Peculiar velocities have only a negligible effect on the predicted rates. Radial peculiar
velocities for high redshift sources are small compared with the recession velocity and we
find that tangential motions of source and lens of order 1000[(ms- 1 have an effect at only
a 10-3 level on the angular change.
As yet we have only considered this simple case of a point mass lens. However, we
expect that the mass distribution of the lensing object is unimportant to first order since
the rate of change of angle is only dependent on the square root of the mass and in any
case one may obtain a constraint by considering more than one lens.
451

Clearly such small yearly changes in angular separation are not currently resolvable,
VLBI techniques achieve at very best micro arc second resolution and thus an increase of
baseline to at least 10 3 earth diameters would be required. For a given baseline interfer-
ometry at optical wavelengths produces a factor 106 improvement in resolution over VLBI.
Optical interferometry with a haseline of only lOKm may be sufficient to detect this effect
over a period of 1 year for the brightest lensed images.

References

Lake K., 1981. Astrophys. J., 247, 17.


Turner E.L., Ostriker J.P., Gott J .R., 1984. Astrophy.1. J., 284, 1.

-10

,....
>. -8 dol:- empty beam. 0= 1
III
()

'"
III
solid:- filled beam. 0= 1
...
()

2,'"
0 -6 dash:- 0=00( 1+z). 0 0=0 1
"-
~

'iii
()

OJ)
0
'0 -4
E
III
0
()

--'
-0
"-
CD -2
<l
-0
.~
.... ~

OL-~~~~ __ ~~~~~~~J-~~~~~J-~~_ _~

o .2 .4 .6 .8
z,
Change in image separation with
time as a function of lens reds hi ft.

Source located at z= 1
DEEP GALACTIC SURVEYS AS PROBES OF THE LARGE SCALE
STRUCTURE OF THE UNIVERSE

O.E.Buryak 1, M.Demianski 2, & A. G. Doroshkevich 1.


t In.,titute of Applied Mathematics, Academy of Science., USSR, Moscow, USSR
2 Institute for Theoretical Physics, Univer6ity of Warsaw, War3aw, Poland

ABSTRACT. We report a new method of analysis of deep galactic surveys and apply it to the
available surveys. We consider a few model distributions which allows us to obtain detailed informa-
tion on the one dimensional characteristics of voids. We show that the one dimensional distribution
of the elements of large scale structure (LSS) is almost Poissonian and we obtain an estimate of the
typical size of voids. We confirm the reality of super LSS (of an approximate size of 80h- 1 Mpc)
the observational evidences of which perhaps are explained as the existence of a population of large
voids. This result is also confirmed by analysis of the CIV doublet distribution in the absorption
spectra of quasars. Full presentation of this method and our results are given in the paper (Buryak
et al. 1991)

1 Introduction

Analyzing the distribution of galaxies in deep galactic surveys it is convenient to treat the
problem as one dimensional, and for characterizing the position of a galaxy to use only
the distance along the line of sight and disregard the angular distribution. To apply our
method we use cluster analysis of a one dimensional distribution of galaxies.
In practice to describe results of the cluster analysis we introduce the following functions
of the radius of clusterization R: the filling factor peR) (fraction of the total length of the
survey occupied by clusters) and mean distance between clusters I(R) These quantities can
be also easily calculated from the observational data.
It is possible to show that for a Poisson distribution of points I(R) = 10 = const. A
more realistic "random empty regions" model consists of a random distribution of empty
regions and a Poissonian distribution of points out of them. This model describes all the
characteristic features of the observational curve l(p).
At present there are three deep galactic surveys which can be analyzed by our method.
There are six plates of the 3 square degrees (KOSS survey) which contain 281 radial ve-
locities of galaxies up to 16m .5 (Kirshner et al. 1983) and five plates of size 13 square
degrees of the English-Australian survey (DARS survey) containing 342 radial velocities
of galaxies up to 17m (Peterson et al. 1986). There is also the English-Australian survey

453
T. Shanks et al. (eds.). Observational Tests a/Cosmological Inflation. 453-455.
Ii:) 1991 Kluwer Academic Publishers.
454

(DAFS) consisting of five plates of diameter approximately 20 arcmin (Broadhurst et al.


1988) containing 188 redshifts of galaxies with 20 m .5 < b < 21 m .5.

2 Results

For all samples (KOSS, DARS, DAFS) the function I(p) in the middle region of p has a
characteristic plateau which indicates that the distribution of complexes is Poissonian and
allows us to determine 10• The value of 10 depends on the size of the plate and increases
with decreasing (J. Therefore the spatial distribution of galaxies in the direction transverse
to the cone is strongly non-uniform. For the DARS survey, the dependence of the total
number of galaxies in the plates and the value 10 on the aperture of the cone (J can be
described by an empirical relation

N((J) = (380 ± 50)((J/40), 10((J) = dV/1 + ~(J


<T >
, (1)

where d = 1150km/ s, TO = 700km/ sand < T(J > denotes the averaged diameter of the
observational wne. The value of TO is close to the comoving correlation radius of galaxies.
For the KOSS survey (J = 1.6 degree and according to (3.1b) 10 = 1730km/ s instead of
a calculated value 2350km/s. Therefore we see that the difference between the KOSS and
DARS surveys is mainly connected with the difference of the angular size of the surveys.
There is a characteristic triangular shape for the function I (p) in the region p '" 0.7
which is similar to that for the random empty regions model. From our analysis it follows
that the mean diameter and fraction of voids, Ie and 0, are approximately given by

lKoss'" 0.8 .104 km/s, [DARS '" 0.55 .104km/s, OKOSS '" 0.4, ODARS '" 0.35, (2)

Typical parameters characterizing the DAFS survey are 10 = 2600km/s, which is rea-
sonably close to the estimation 10 = (2100 - 2200)km/ s, following from (1).

3 Conclusions

The main result of our analysis of deep galactic surveys is the demonstration of the Pois-
sonian character of distribution of the basic elements of the LSS, and estimation of the
mean distance between non-correlated complexes of gala..xies along a line of sight which we
identify with d - an asymptotic value of 10 for large angle (J

lo( (J > 20 ) = d", 1100km/ s. (3)

The fact that 10 depends on the average diameter of the sample < r(J > explains why
the value of 10 is different for different surveys.
For the KOSS and DARS surveys we observe the characteristic triangular shape on
the curves I(p) which is typical also for our model. These results are interesting as a
confirmation and perhaps explanation of results obtained by Broadhurst, Ellis, Koo and
Szalay (1990)(BEKS).
455

It is interesting to compare the BEKS results with an analysis of the absorption spectra
of quasars. The simplest tracers of the large scale structure of the Universe in the spectra of
quasars are the absorption systems which probably are connected with the separate galaxies
or possibly with some groups of galaxies. Usually such absorption systems contain lines of
ions of heavy elements. At present there is a list of 45 quasars containing redshifts of 220
doublets of elY (Sargent, Steidel and Boksenberg, 1988) between z = 1.27 and z = 3.226.
This list allows us to obtain 175 distances between doublets elY. We can form a composite
model of the distribution of points using invariant distances between real doublets of elY.
Such an analysis shows that the distribution of these points is Poissonian and l(p) has a
plateau at the value 1 = 7800km/ s for p :$ 0.9. We consider this result as an independent
confirmation of the results of BEKS.

References

Buryak, O.E., Demianski, M. & Doroshkevich, A.G., 1991. preprint. Broadhurst, T.J., Ellis,
R.S. & Shanks, T., 1988. Mon. Not. R. astr. Soc, 235, 827.
Broadhurst, T.J., Ellis, R.S., Koo, D.C. & Szalay, A.S., 1990. Nature, 343, 726.
Kirshner, R.P., Demler, A.J., Schecter, P.L. & Schectman, A.S., 1983. Astron. J., 88, 1285.
Peterson, B.A., Ellis, R.S. & Efstathiou, G., 1986. Mon. Not. R. astr. Soc, 221, 233.
INTERGALACTIC ABSORPTION IN THE SPECTRA OF mGH-REDSHIFT QSOs

S. CRISTIANI 1 & E. GIALLONG0 2


1 Dipartimento di Astronomia, Universita di Padova, Italy
20sservatorio Astronomico di Roma, Italy

ABSTRACT. Average flux decrements in the Lya jOTest have been computed from the observed
spectra of seven quasars. Comparing all the available data with simulations of the average absorp-
tions in QSO spectra, we suggest that, within the observational and statistical uncertainties, there
is evidence neither for a dramatic change in the evolution of the absorbers, nor for the presence of
a new diffuse component of neutral hydrogen in the spectra of very high redshift (z > 4) quasars.

1. INTRODUCTION

Information about the cosmological distribution of the absorbing clouds giving origin to the Lya
forest is obtained from the statistical analysis of the numerous systems observed in samples of
high redshift quasars. To a first approximation the number of absorption lines per unit interval
of absorption redshlft z and intrinsic equivalent widths W (= Wob ./(1 + z)) is ex N*(z)e- w / w '
where W* is a characteristic equivalent width independent of z, and N*(z) represents the redshift
evolution of the line number density. For a uniform cloud distribution the evolution takes the form
=
of a power-law in redshift with index ~ 0.5 - 1, for qo 0.5 - 0 respectively.
At low resolution this "forest" of absorptions appears as a steepening of the continuum slope
in the region shortward the Lya emission as compared to the slope defined on the red side of Lya.
Since accurate line statistics have been performed only in a limited range of redshifts (z = 1.5 to
3.8), where high resolution spectroscopy is available, it is possible to investigate the properties of
the intergalactic absorbers at lower and higher redshifts by measuring the average depressions in
well defined regions shortwards of the Lya emission.
The average flux decrements DA, between Lya and Lyf3 emissions, and DB, between LY{1
and the Lyman limit have been introduced, in this respect, to study the average properties of
the absorption (Oke and Korycansky, 1982). These quantities are, within relatively large limits,
independent of the instrumental resolution, at least as far as the true continuum can be reliably
extrapolated from the unabsorbed (red) region.
Schneider, Schmidt & Gunn (1989), on the basis oftheir sample of quasars with Zen> > 4, have
recently shown that the depression observed shortwards of the Lya emission exceeds that predicted
from the Lyman clouds statistics. As a possible explanation of this excess they suggested a dramatic
change in the increasing number of clouds for z > 3, or an evolution in the physical properties of
the individual clouds, or the presence of a new distributed component of neutral hydrogen, the so
called Gunn-Peterson effect.
On the basis of new observations and of synthetic spectra we explore the consistency between
clouds statistics and the evolution of the average depressions in the quasar spectra.

2. OBSERVATIONS AND NUMERICAL SIMULATION

The observations of the quasars have been carried out at the 3.6m and the 2.2m ESO telescopes
at La Silla. The resolution of the spectra ranged between 15 and 20 A. The seven quasars chosen
for low-resolution spectroscopy have been selected on the basis of their redshift (in the interval
2.8 < Zen> < 3.8) and apparent magnitude (brighter objects have been preferred) and are by no
means a complete sample.
We use also a synthetic QSO absorption spectrum to compute the average flux decrements in
= =
the regions A (~A 1040 - 1190 A) and B (~A 912 - 1020 A). In our simulation the following

457
T. Shanks et al. (eds.). Observational Tests o/Cosmologicallnf/ation. 457-459.
e 1991 Kluwer Academic Publishers.
458

classes of absorbers are taken into account: the Lya lines not associated with met,al-lines systems
(Hunstead et al. 1988; Carswell et al. 1984); metal-lines systems (Sargent et al. 1980); damped Lya
systems (Wolfe et al. 1986; Sargent et al. 1989). We extract a set of random absorption redshifts
and equivalent widths according to the appropriate statistical properties of the various classes and
we compute, for each quasar emission redshift, 100 synthetic absorption spectra using Voigt profiles
with natural damping only.

3. RESULTS

In Fig. 1 the distribution in redshift of the depression D A for the available data is shown. We
estimate that for the region A discrepancies among the various authors are mainly due to different
choices of the spectral range. In spite of this, our data show that intrinsic dispersions are as large
as systematic differences of the quasar samples.
Dispersions obtained by binning the data of Fig. 1 at Zem ~ 3 are consistent with the intrinsic
=
ones obtained from Monte Carlo simulations, (jD A ~ 0.05 and the average in the bin az 3.1- 3.3
with (z) = 3.2 is (DA) =
0.35 ± 0.05 in agreement with the corresponding value of the simulations
(D A )= 0.34 ± 0.05. Considering the nine quasars at (zem) = 4.2, we have an observed value
= =
(D A ) 0.56±0.07, to be compared with (D A ) 0.54±0.04 from the simulations shown in Fig. 1.

.8
0 Steidel & Sargent (1987)

cOke & Korycansky (1982) · I
.6
:-f'!
r
• Schneider et al. (1989a,b)

~
.4
I!>. O'Brien et al. (1988)
t.i I r
0

I.J·
0
0
.2
III I .0

0
1 2 3 4 5
z
Figure 1. D A values as a function of redshift. Crosses represent average values from the Monte
Carlo simulations.

The corresponding similar analysis of the average absorption DB in the B region is complicated by
459

differences in the S/ N ratio of the various observations which leads to different choices of the short
wavelength linut.
A similar comparison for the average ratio DB/DA confirms early results (Giallongo et al.
1990) concerning the intrinsic shape of the QSO continuum. In fact, observed values greater than
one are reproduced by simulations for any resonable value of the Doppler line parameter, with a
decreasing trend toward higher z, as expected for blending effects of numerous lines. Thus, there is
not evidence for an intrinsic steepening of QSO spectra up to the Lyman linut even at low redshifts.
An optical- UV slope of the intrinsic continuum extending toward the soft X-ray band constrains the
models for the QSO enussion.
Within the observational and statistical uncertainties, the number density of damped lines
extrapolated to z ;;:: 4 is roughly consistent with the number of damped candidates present in the list
of Schneider, Schnudt & Gunn (1989). We conclude that the discrepancy between the cosmological
evolution derived from line statistics and the cosmological evolution of the average absorption D A
could be reconciled including a fiat distribution in equivalent widths of stronger absorption lines
associated with metal line systems up to the damped Lyman systems (Giallongo & Cristiani, 1990).
Such a statement implies the cosmological persistence at high z of non-luminous structures
in the comoving space, in contrast with the apparent decline in the number of luminous quasars.

References

Carswell, R. F., Morton, D. C., Smith, M. G., Stockton, A. N., Turnshek, D. A. & Weymann, R.
J., 1984. Astrophys. J., 278, 486.
Giallongo, E., & Cristiani, S., 1990. Mon. Not. R. astr. Soc., 247, 696.
Giallongo, E., Gratton, R. & Trevese, D., 1990. Mon. Not. R. astr. Soc., 244, 450.
Hunstead, R. W., Murdoch, H. S., Pettini, M. & Blades, J. C., 1988. Astrophys. J., 329, 527.
O'Brien, P. T., Gondhalekar, P. M. & Wilson, R., 1988. Mon. Not. R. astr. Soc., 233, 801.
Oke, J. B. & Korycansky, D. G., 1982. Astrophys. J., 255, 11.
Sargent, W. L. W., Young, P. J., Boksenberg, A. & Tytler, D., 1980. Astrophys. J. Suppl., 42, 41.
Sargent, W. L. W., Steidel, C. C. & Boksenberg, A. 1989. Astrophys. J. Suppl., 69, 703.
Schneider, D. P., Schmidt, M. & Gunn, J. E., 1989. Astron. J., 98, 1951.
Wolfe, A. M., Turnshek, D. A., Smith, H. E. & Cohen, R. D., 1986. Astrophys. J. Suppl., 61, 249.
A COMPLETE QUASAR SAMPLE AT INTERMEDIATE REDSHIFT

F. LA FRANCA1,2, S. CRISTIANI1
C. BARBIERI1,3, R. G. CLOWES 4 , A. IOVIN0 5
1. Department of Astronomy, Padova, Italy
2. Institute of Radio Astronomy, CNR, Bologna, Italy
3. Astronomical Observatory, Padova, Italy
4. Royal Observatory, Edinburgh, Scotland
5. Brem Astronomical Observatory, Milano, Italy

ABSTRACT. A search for intermediate-redshift quasars has been carried out with slitless spec-
troscopy in the central 21.07 deg 2 of the SA 94.

1. The selection
The field of the SA 94 has been since some years our test-field for a number of investi-
gations about quasars, providing samples which are the product of three independent and
complementary techniques (multi-color, objective prism and variability) and allowing us to
quantify with unprecedent accuracy the biases affecting each selection method (Barbieri
and Cristiani, 1986; Cristiani et al. 1989, 1990).
We have carried out a slitless survey based on the AQD (Clowes 1986, and references
therein) plus a template matching technique in the central 21.07 deg 2 ofthe SA 94 (Cristiani
et al. 1990). Using the QSO composite spectrum b) of Cristiani and Vio (1990) as a
template spectrum, a grid ofredshifted QSO spectra has been obtained in the range 1. 7 ~
z ~ 3.4, in which the slitless technique is most effective. Each objective prism spectrum has
then been compared with this grid of simulations and the "best-guess" redshift has been
finally derived by least-squares non-linear fitting. The corresponding RMS of the observed
spectrum with respect to the simulated one has been calculated and assumed as indicative
of the quality of the candidate in function of the magnitude. 50 candidates have been
observed with slit spectroscopy, confirming 34 quasars and 2 HI! galaxies.

2. The quasar number counts and evolution


The completeness of this survey as a function of magnitude and redshift has been analysed
and an effective area of 16.9 deg 2 has been evaluated. The following corrections have been
taken into account:
a) a random incompleteness of 20% , fundamentally due to the overlapped spectra;
b) a differential galactic extinction LlAB = -0.15 with respect to the galactic pole (see
Burstein and Heiles, 1982). Mter this correction the magnitudes are indicated as b =
B - 0.15.
The estimated quasar surface density for QSOs with 1.8 < z < 3.2 and b < 19.25 is
1.4 ± 0.3, in agreement with other surveys in the literature (Crampton et al. 1989, Foltz et
al. 1989). All this information at intermediate redshift can therefore be confidently used to
address the issue of the evolution of the quasar volume density and luminosity function. This

461
T. Shanks et al. (ells.), Observational Tests of Cosmological Inflation, 461-462.
© 1991 Kluwer Academic Publishers.
462

turns out to be easy if we restrict the investigation to quasars with -27.35 < MB < -28.5.
In this case relevant data besides our survey can be obtained from the PG (Schmidt and
Green, 1983), the MBQS (Mitchell et aI. 1984), the APM (Foltz et aI. 1989), the CFHT
(Crampton et aI. 1989) and the MMB (Mitchell, Miller and Boyle, 1990) samples.
It is apparent (Fig.1) that the volume density of bright quasars such those studied in
the present survey is strongly increasing in the past: about a factor 103 from 7 to 11 Gyr.
0.2 Redshifl 1 2 34
....
0
,,~

-1 1 1-
.,
~

I Ho=50 qo=0.5

..
0
p. -27 .35<MB<-28.5
::iii
I
....
0

t-
'ol
.,1:1 " PG
.,
Q

....
El

1
~
.E0
>-
....
0
~
0 2 4 6 8 10 12
Time (Gyr)
Fig. 1. The evolution of the comoving density of quasars
From the present data, no firm sign of a decrease in the volume density at early epochs
can be found. If any, it has to take place within the first 1.5-2 Gyr after the Big Bang. The
MMB data might show some evidence for a flattening, but the statistics is extremely poor
for that redshift range. A better assessment of this issue will be possible when the MMB
and other high redshift surveys will be compieted.
Fainter surveys of the type carried out by Crampton et aI. (1989), will allow to study
the volume density evolution of less luminous objects, while large-area surveys are needed
to fill the gap between the MBQS and the PG survey. This is precisely the aim of an ESO
Key-Programme we have recently undertaken (Barbieri et aI. 1989).

References
Barbieri, C., and Cristiani, S., 1986, Astron. Astrophys. Suppl. 63, 1
Barbieri, C., Cristiani, S., Andreani, P., Clowes, R.G., Gemmo, A., Gouiffes, C., Iovino, A., La
Franca, F., Savage, A., Vio, R., 1989, ESO Messenger, 58, 22
Burstein, D., Heiles, C., 1982, Astron. J. 87, 1165
Clowes, R.G., 1986, Mon. Not. R. astr. Soc. 218, 139.
Crampton, D., Cowley, A.P., Hartwick, F.D.A., 1989, Astrophys. J. 345, 59
Cristiani, S., Barbieri, C., Iovino, A., La Franca, F., and Nota, A., 1989, Astron. Astrophys. Suppl.
77, 161
Cristiani, S., and Vio, R., 1990, Astron. Astrophys. 227, 385
Cristiani, S., La Franca F., Barbieri, C., Clowes, R.G., Iovino, A., 1990, Mon. Not. R. astr. Soc.
in press
Foltz, C.B., Chaffee, F.H., Hewett, P.C., Weymann, R.J., Anderson, S.F., MacAlpine, G.M., 1989,
Astron. J.98, 1959
Mitchell, P.S., Miller, L., Boyle, B. J., 1990, Mon. Not. R. astr. Soc. 244, 1
Mitchell, K.J., Warnock, A., Usher, P.D., 1984, Astrophys. J. 287, L3
Schmidt, M., and Green, R.F., 1983, Astrophys. J. 269, 35£
RADIO-LUMINOSITY DEPENDENCE OF THE IR-RADIO ALIGN-
MENT EFFECT IN HIGH-z RADIO GALAXIES

J.S. DUNLOP
Department of Physics & Astronomy,
Lancashire Polytechnic,
Preston PR1 2TQ.

J.A. PEACOCK
Royal Observatory,
Blackford Hill,
Edinburgh EH9 3Hl.

ABSTRACT. We present initial results from a K-band imaging study of a subsample of radio
galaxies from the S2.7GHz > 0.1 Jy Parkes Selected Regions (PSR), along with a comparison sample
of 3CR galaxies which are a factor ~ 10 more radio luminous. Our K-band images of 3CR galaxies
at z ~ 1 show that i) the established optical-radio alignment effect is still sufficiently apparent at K
to have been easily discovered in the near-IR, and ii) our IRCAM images are certainly deep enough
to reveal comparable distortions in the PSR galaxies, should they exist.
However, we find no evidence for a comparable IR-radio alignment effect in t.he PSR sample,
and also find the PSR galaxies to be generally less elongated and have redder R-K colours than their
3CR counterparts. The only significant a priori difference between the two subsamples of objects
appears to be radio luminosity.
These preliminary results indicate that, at least at K, radio-induced contamination of the light
of the underlying galaxy may only be a significant problem for the most extreme radio luminosities.
This suggests that radio galaxies may still be relevant probes of galaxy evolution in general, provided
one avoids the most extreme radio luminosities found in the 3CR gala.'Cies at z ~ 1.

1 Introduction

We present preliminary results from a near-infrared study of radio galaxies from the Parkes
Selected Regions (PSR) sample l , along with a control sample of the more powerful 3CR
galaxies. At a given redshift, the PSR galaxies have radio luminosities a factor rv 10
lower than the powerful 3CR radio galaxies, and hence provide an opportunity to separate
radio luminosity effects from genuine epoch-dependent effects which might apply to massive
galaxies in general. Of particular interest is the investigation of the prevalence and possible
radio-luminosity dependence of elongation and alignment with the radio axis at near-IR
wavelengths. We now possess IRCAM K-band images of 38 PSR galaxies and 19 3CR
galaxies, with integration times of between 27 and 54 minutes.

463
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 46~.
© 1991 Kluwer Academic Publishers.
464

2 Infrared-radio alignments in the 3CR sample

The first question to address is whether or not the well-established tendency for the optical
images of 3CR galaxies with z ~ 0.8 to be elongated and aligned with the radio axis is
also prevalent at 2.2 p.m. Optical position angles for 15 of the 3CR galaxies in our sample
are either given by McCarthy ct al. (1987)2, or can be easily determined from published
images 3 ,4.5.6. For these objects we show (Figure 1) the optical-radio alignment histogram
compared with the IR-radio alignment histogram derived from our IRCAM data (to avoid
subjectivity, the infrared position allgles were determined automatically from our K-band
images using a moment analysis). Although the extent of the elongation is generally slightly
less than in the optical, this diagram shows that the alignment effect is still sufficiently clear
to have been 'discoverable' at K (in both cases, comparison with a uniform distribution
using the KS test gives a significance level p < < 0.01). This indicates that, even at K, a
significant fraction of the light from these galaxies is being influenced by the passage of the
radio source (most probably through stimulation of large-scale star-formation activity).

3 Lack of infrared-radio alignments in the PSR sample

In contrast to the 3CR sources, we find no clear evidence for an IR-radio alignment effect
in our sample of PSR galaxies (comparison with a uniform distribution gives p > 0.2).
However, whereas our 3CR sample is very strongly peaked around z '" 1, the PSR sample
contains a much wider range of redshifts, raising the worry that any alignment effect could
be diluted. For this reason, and also to remove any confusing epoch dependence, the
comparison of the two samples is clarified by confining attention to two 'matched' galaxy
subsamples with 0.8 :::; z :::; 1.3.

4 Comparison of 3CR and PSR subsamples at z '" 1

Selecting only those objects with redshifts in the range 0.8:::; z :::; 1.3 produces two 'matched'
subsamples of galaxies - 19 3CR galaxies and 15 PSR galaxies (the PSR redshifts have to
be estimated from their K magnitudes, but this is still a reasonably reliable procedure at z
'" 1 - see e.g. refs. 1 & 7). These subsamples are statistically indistinguishable in redshift
distribution, distribution of radio angular size, and distribution ofradio spectral index. This
leaves radio luminosity as the only significant a priori difference between the subsamples,
but three other important differences are revealed by our K-band imaging, namely: i)
virtually all the PSR galaxies are less elongated than those in the 3CR subsample, ii)
the 3CR subsample shows clear evidence of an IR-radio alignment effect, while the PSR
subsample shows none (Figure 2), iii) the PSR galaxies at z '" 1 generally have redder R-K
colours than their 3CR counterparts.
465

-
o

to
-
0

to -
I

co co :- -
z Z
"<t' I- -
N I- -
o ~~~~~~~~~~
0
I I I I

o 20 40 60 80 0 20 40 60 80
II PA/degrees (optical-radio) II PA/degrees (IR-radio)
Figure 1 Comparison of the optical-radio (LH plot) and K-band-radio (RH plot) alignment
histograms for 15 3CR galaxies with 0.8 < z < 1.3.

-
0

to
-
0

to
3CR PSR
CO CO
Z Z
"<t' "<t'

N N

0 0
0 20 40 60 80 0 20 40 60 80
II PA/degrees (IR-radio) II PA/degrees (IR-radio)

Figure 2 Comparison of the K-band-radio alignment histogram for the z '" 1 3CR sub-
sample with the K-band-radio 'non-alignment' histogram for the z '" 1 PSR subsample.
466

5 Conclusion

These initial results imply that, at least at K, pollution of the light of the underlying galaxy
by radio-induced effects may well be only significant for the most extreme radio luminosities.
This suggests that, despite some scaremongering in the literature, radio galaxies may still
be relevant probes of galaxy evolution in general, provided one steps down from the most
extreme radio luminosities found in the 3CR sample at z rv 1.

References

1. Dunlop, J.S. et al., 1990. Mon. Not. R. astr. Soc, 238,1171.


2. McCarthy, P.J. et al., 1987. Astrophys. J., 321, L29.
3. McCarthy, P.J., 1988. PhD thesis, University of California.
4. O. Le Fevre, F. Hammer & J. Jones, 1988. Astrophys. J., 331, L73.
5. O. Le Fevre & F. Hammer, 1988. Astrophys. J., 333, L37.
6. F. Hammer & O. Le Fevre, 1990. Astrophys. J., 357, 38.
7. Lilly, S.J., Longair, M.S. & Allington-Smith, J.R., 1985. Mon. Not. R. astr. Soc, 215, 37.
DENSITY AND PECULIAR VELOCITY FIELDS IN THE
REGION OF DRESSLER'S SUPERGALACTIC PLANE SURVEY

MICHAEL J. HUDSON
In&titute of A.dronomy,
Cambridge Univer,ity,
Madingley Road,
Cambridge CB3 OHA.

Summary

The dynamics of the 'Great Attractor' region are of considerable interest. Dressler has
recently completed the Supergalactic Plane Survey (SPS), a redshift survey of ESO galaxies
between -35 :::; b :::; 45 and 290 :::; I :::; 350. We calculate the density fields and peculiar
velocity fields for the SPS, and use the SSRS as a 'fair sample of the universe' to calculate
the diameter functions and the average density.
Survey completeness as a function of apparent diameter S( 0) and metric diameter
functions ¢(D) (analogous to luminosity functions) are discussed elsewhere (Hudson &
Lynden-Bell 1991, in preparation). A smoothed density field is obtained by convolving the
weighted set of survey galaxies with a top hat window of radius 600 km/s. The weight
assigned to each galaxy in the redshift survey at distance r with metric diameter D = rO is
w = D2 / C( Omin r), where the fraction of the total D2 ('light') which is seen by the redshift
survey at distance r is C(Ominr) = JOo:,;n r D2S(D/r)¢(D)dD/ Jo D2¢(D)dD. For regions
outside the SPS area or with r ;::: 8000 km/s, the density is set equal to the average SSRS
density, which is also assumed to be the average density of the universe. The density in the
galactic plane (-10:::; b:::; 10) is a linear interpolation between the -15 :::; b:::; -10 strip
and the 10 :::; b :::; 15 strip. Initially, Local Group frame velocities are used in place of r.
Using the smoothed density field and setting no. 6 /b = 004, peculiar velocities are calculated
for each galaxy. These are used to obtain improved estimates of r. Weights and the density
field are recalculated and the procedure is iterated until convergence.
Figure 1 shows the resulting density and peculiar velocity fields for the SPS. It can be
seen that whereas the Centaurus cluster is at the densest region (l '" 300, b '" 20, cz '" 3000),
there is a secondary peak at (l '" 325, b '" 35, cz '" 7000) and a bridge connecting the two.
Thus the center of overdensity of the supercluster is beyond the Centaurus cluster, and so
the Centaurus cluster experiences an outflow. Figure 6 of Lynden-Bell (these proceedings)
shows in detail the density field in the 'new' supergalactic plane.

467
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 467-468.
@ 1991 Kluwer Academic Publishers.
468

a) d)

b) e)

c) f)

Figure 1. Wedge diagrams in galactic coordinates for Dressler's SPS redshift survey. The
contours for the smoothed (500 km/s) density field are shown for the mid-plane of the
wedge and the contour levels are multiples of the average SSRS D2 density, with the first
(narrowest) contour at the average density. One third of the galaxies in the survey are
shown with the tips of the vectors at the redshift of the galaxy and the circles at the model
distances. a) 25 ::; b ::; 45; b) 10 ::; b ::; 25; c) -35 ::; b ::; -10; d) 290 ::; l ::; 310;
e) 310 ::; l ::; 330; f) 330 ::; l ::; 350.
SCALE INVARIANCE INDUCED BY NON-LINEAR GROWTH OF
DENSITY FLUCTUATIONS

F. MOUTARDE & J.-M. ALIMI


L.A.E.C.. ObsertJatoire de Meudon,
98195 MEUDON Cedez. FRANCE.

F.R. BOUCHET
Institut d'Astroph!lsique de Paris,
75014 PARIS. FRANCE.

R. PELLAT
Centre de Ph!lsique Theorique. Ecole Pol!ltechnique,
91188 PALAISEAU Cedez. FRANCE.

ABSTRACT. We investigate analytically and numerically the formation of smooth dark matter
halos in an 0 = 1 universe. We have proved analytically in the spherical case that a power-law
density profile is expected before shell-crossing and give a general formula relating the slope to
the initial overdensity. Performing high resolution particle-mesh numerical simulations (up to 1283
particles in a 1283 grid) we have shown that a self-similar regime takes place soon after shell-crossing
in various 3D and 2D collapses (spherical, cylindrical or superposition of 2 or 3 caustics). We have
compared this regime in some details with the semi-analytic approximate model of Fillmore and
Goldreich (1984). We suggest that the afore-mentioned scale invariance at collapse provides a
natural starting point for the self-similar solutions, which was the missing link in the semi-analytic
studies.

1 Scale-invariance at collapse
In Moutarde et al. (1991a) we have pointed out the formation of power-law density profiles
at the collapse time during the formation of halos from various initial smooth overdensities.
In the case of spherical symmetry exact analytic treatment is possible until the collapse
time (see Peebles 1980, §19). Using the parametric expressions for 8M/M and r given
in the above reference, we have shown that (1 + 8M/M) <X r-/3 for 8M/M » 1 at
the collapse time. The obtained power-law index is related to the initial overdensity by
(3 = 6a/(2a + 3) where a is defined by the generalized development of (8M/M); around
r; ~ 0: (8M/M); ~ 60[1 - (r;/ro)aj. The details of the calculations will be given in a
forthcoming paper (Moutarde et al. 1991b). We have checked this result by computing
analytically the density profiles at collapse for various spherical initial perturbations (see
figure 1). We have also ran numerical simulations with the same initial conditions. The
good agreement obtained with the analytic results, even for very high densities and small
scales, incidently provide a test of the reliability of the code.

469
T. Slumks et al. (eds.J. Observational Tests o[Cosmologicallnflation. 469-470.
© 1991 Kluwer Academic Publishers.
470

5,-----,----,.-----r----,---,

Figure 1: density profiles at collapse for three -


A
different spherically symmetric illitial over- ::I
V
densities. The lines correspond to the analytic ~
computations and the symbols to the numeri- '0
cal simulations. The values for the parameter +
...
Q (see text) are 1, 2,4 and the profiles are in 'l
good agreement with the predicted power-laws .2
of respective index -6/5, -12/7, and -24/11.
Ol.--_L..-_L..-_L..-_Lao-.l
-2 o
2 Self-similarity after shell-crossing
After shell-crossing, the evolution can be determined only by numerical simulations. We
find for all our 2D and 3D simulations that the evolution quickly becomes self-similar,
as was already pointed out in Alimi et 4/. (1990) in the particular case of 2 superposed
caustics. This self-similarity is seen on the density profile which becomes a power-law in
space and time, and on the velocity distribution which is invariant when the velocities are
scaled to their r.m.s. value (which itself varies as a power-law of time). Such self-similar
regimes were suspected by Gunn (1977) and studied semi-analytically by Bertschinger
(1985) and Fillmore & Goldreich (1984). We compared in some details our numerical
simulations with the hypothesis, predictions and conjectures of these authors. In particular
we confirm the power-law time dependance of successive apocenters of a given particle, the
existence of an adiabatic invariant along particle trajectories, and a minimum steepness of
the density profile during self-similarity. An interesting point is that the self-similar slope
seems to be the same as the one at collapse if the latter is steeper than the afore-mentioned
minimum. This indicates that the scale invariance at collapse pointed out in §1 prepares
the system for the subsequent self-similar evolution.

Acknowledgements. The computational means we used (Cray2) were made available thanks to
the scientific counsel of the "Centre de Caleul Vectoriel pour la Recherche" (Palaiseau).

References
Alimi J.-M., Bouchet F. R., PeUat R., Sygnet J.-F., & Moutarde F., 1990. Astrophys. J., 854, 3.
Bertschinger E., 1985. Astrophys. J. Su.pp., 58, 39.
Fillmore J.A., & Goldreich P., 1984. Astrophys. J., 281, 1.
Gunn J.E., 1977. Astrophys. J., 218, 592.
Moutarde F., Alimi J.-M., Bouchet F. R., PeUat R., & Ramani A., 1991a. Astrophys. J. Lett., ,
submitted.
Moutarde F., Alimi J.-M., Bouchet F. R., & PeUat R., 1991b. , , in preparation.
Peebles P.J.E, 1980. in The Large-Scale Stru.cture of the Universe, Princeton University Press.
THE POWER SPECTRUM OF GALAXY CLUSTERING

J .A. PEACOCK
Royal Ob,ervatory
Blackford Hill
Edinburgh EH9 !lH], UK.

ABSTRACT. There now exist several datasets with the depth necessary to determine the power
spectrum of galaxy clustering on scales ~ 100 Mpc. Results from redshift surveys and angular
clustering all agree on the form of the function, which may be well approximated for optically-
selected galaxies by ~2(k) = 0.08y4/(1 + y2.4), where y == k/0.034h MpC 1 • At small k, h = 0.5
CDM with J3 normalization gives an amplitude a factor 2.0 smaller. Since normalization should aim
to match spectra on the largest scales, where linear theory is most nearly applicable, it seems that
investigations of the CDM model should therefore use a higher amplitude of fluctuations; an increase
by a factor 2 is probably allowed by present limits on fluctuations in the microwave background.
N-body simulations have not studied such a model in much detail; it may be premature to reject
the n= 1 adiabatic CDM universe as a viable picture for galaxy formation.

1 Introduction

Ever since its inception, the 'standard' CDM model (i.e. a Universe with n = 1 dominated
by Cold Dark Matter and containing scale-invariant adiabatic fluctuations) has suffered
from allegations that it is deficient in power on large scales (~ 10 Mpc). An early worry
was the apparent positive autocorrelation function of Abell clusters out to separations of
nearly 100h- 1 Mpc (Bahcall & Soneira 1983); as usual, h == Ho/100 kms- 1 MpC 1 • This
problem was apparently confirmed by the existence of a peculiar velocity field of amplitude
'" 1000 kms- 1 Mpc-t, coherent over scales ~ 10h-1 Mpc. Possible models with different
power spectra were suggested in response to these difficulties (e.g. Salopek et al. 1989;
Peebles 1987), but none of these were very strongly motivated within the context of inflation.
In any case, doubts about the observations persisted (some well justified: Sutherland 1988),
apparently allowing CDM to survive. However, two new major pieces of evidence have
recently been presented which both claim that the CDM power spectrum is greatly deficient
in fluctuations of wavelength ~ 50h- 1 Mpc. These are the angular two-point autocorrelation
function for galaxies in the APM survey (Maddox et al. 1990) and the counts-in-cells
analysis of the QDOT !RAS redshift survey (Efstathiou et al. 1990). The unanimity of
these latter two pieces of work is impressive, but the consequences for CDM may not yet
be fatal, for the following reasons.

471
T. Shanks et a/. (eds.). Observational Tests ofCosm%gicallnflation. 471-474.
e 1991 Kluwer Academic Publishers.
472

2 The power spectrum

Let us try to reduce the recent pieces of work to a power spectrum. For the APM w( 8)
results, Limber's equation may be expressed in terms of the 3D power spectrum:
w(8) = Jooo y44>2 dy Jooo 'If 1).2(k) Jo(ky8) dkjk 2,
[Jooo y24>dy]2
where 1). 2 = d0'2 j dIn k ex: k 3 PC k) and 4>(Y) is the selection function. Rather than attempt a
noisy deprojection, we can apply some theoretical prejudice and seek a formula for 1).2(k)
which is consistent with scale invariance (1).2 ex: k4 for small k). The following gives a
reasonable fit to the APM data:
1).2(k) = 0.08 y4 . y == kjO.034h Mpc- 1 •
(1 + y2.4) ,
We can attempt to compare this empirical fit with points derived from redshift surveys.
Direct power-spectrum analysis has been performed for two samples: (a) the efA survey
(Baumgart & Fry 1991); (b) a survey of radio galaxies at z < 0.1 (Peacock & Nicholson
1991). Note that it is a much more robust procedure to determine the power spectrum
directly than by transforming the correlation function: in the latter case, conclusions about
large-scale clustering are very sensitive to the assumed mean density.

....
o
a
0.01 0.1
k/h Mpc- 1

Figure 1 The power spectrum in the form 1). 2 == d0'2 j dIn k. The solid line shows a model fit
to w(8); filled points are radio galazies {Peacock fJ Nicholson 1991} with 1).2 reduced by a factor 3;
open circles are IRAS {Efstathiou et al. 1990} with 1).2 increased by a factor 1.2; crosses are from
the CfA {Baumgart fJ Fry 1991}. The fact that points lie below the line for k ~ 0.2 is probably due
to peculiar velocity smearing.
Lastly, results from the IRAS QDOT survey are published in the form of the variance
(0'2) of 0 in cubical cells of side i. For a Poisson spectrum (1).2 ex: k 3 ), we have
0'2 = 1). 2 (k = (2'1f 2)1/3 j i) ,
473

which simply makes the obvious point that the variance is mainly sensitive to waves with
A ~ 21. At the 20% level, this formula works for a wide range of power spectra. Given the
observational uncertainties in the IRAS points, we shall therefore simply plot the IRAS 0"2
points on the 6 2 - k plane as above.
The results are shown in figure 1, after some vertical scaling to allow for the we1l-
established facts that radio galaxies cluster more strongly that optically selected galaxies,
while IRAS galaxies cluster less strongly. There is an impressive degree of unanimity about
these data, and it is clear that a break in the spectrum at A ~ 100h-1 Mpc has been
detected. The fact that the break is at such a large scale shows why it has taken so long to
obtain data of the necessary depth to detect it.
In passing, we note that the value of 6 2 is < 0.1 at the wavelength of 12Sh- 1 Mpc
singled out by Broadhurst et al. (1990); the Universe is reasonably close to homogenous on
these scales.

3 Amplitudes and models

It is useful to characterise these results in some dimensionless way. For scale-invariant


spectra, the obvious choice is the rms metric fluctuation, defined in terms of the power
spectrum of the gravitational potential fluctuations:


2=_ V 31 12/ C4 -_ 49 (Ck)
(21r)341rk ¢>k Ho
-4 2
6 mass (k).
The above power spectrum therefore yields €b = 4.1 X 10- 5 (allowing for bias).

Adiabatic COM:
h=l. 0.5. 0.3

0.01 0.1
k/h Mpc- 1

Figure 2 A comparison of the 'observed' power spectrum with linear CDM. For J 3 normalization,
the CDM curves would crolls the solid line at k ~ 0.1 - producing a great apparent deficit of large-
scale power.
It is interesting to compare this with n = 1 adiabatic CDM models, which generally
give smaller values of €. For J3 normalization the following is a good approximation:
474

fb = 4.1 x 1O-5/(4.3h - 0.13 - 0.65h4 ). For h = 0.5, the result of normalizing to A2 at


the largest wavelengths is a spectrum with an amplitude about a factor of 2 higher than is
conventional. This would not be ruled out by current limits on the microwave background,
although it could put us very close to detection if the optical clustering amplitude is nearly
unbiased (Bond et al. 1991). Figure 2 compares the linear CDM power spectra with
observation, renormalized in this way; from this point of view, the problem with CDM is
that it produces too much small-scale power. However, for low bias, the discrepancy now
occurs only at a scale where A '" 1: nonlinear effects could well alter the shape of the
spectrum considerably at this point (Efstathiou et al. 1988).
If a model with linear power spectrum that fits the data is demanded, one could not
do better than isocurvature CDM with h = O.S and eb = S.3 X 10-5 • This gives a rather
sharper bend at the break scale, as the data seem to require. Sadly, this model does appear
to conflict with CBR limits (Efstathiou & Bond 1986).

4 Conclusions

The form of the clustering power spectrum appears to be reasonably well determined at
last, and a break towards scale-invariance is seen on very large scales. This argues for
a higher amplitude of fluctuations in models such as CDM. In the past, this suggestion
would have been rejected as predicting too high small-scale velocities. However, with the
increasing awareness of strong 'velocity bias' effects in high-resolution N-body simulations
(Carlberg et al. 1990), this objection may be invalid.
In short, tests of the CDM model may depend on our ability to make correct predictions
in the most difficult non-linear regimes. Our understanding of these issues is still evolving,
and it may be premature to reject the model.

References

Bahcall, N.A. & Soneira, R., 1983 Astrophys. J., 270, 20.
Baumgart, D.J. & Fry, J.N., 1991. for submission to Astrophys. J.
Bond, J.R., Efstathiou, G., Lubin, P.M. & Meinhold, P.R., 1991. submitted to Phys. Rev. Lett.
Broadhurst, T.J., Ellis, R.S., Koo, D.C. & Szalay, A.S., 1990 Nature, 343, 726.
Carlberg, R.G., Couchman, H.M.P. & Thomas, P.A., 1990 Astrophys. J., 352, L29.
Efstathiou, G. & Bond, J.R., 1986. Mon. Not. R. astr. Soc., 218, 103.
Efstathiou, G., Frenk, C.S., White, S.D.M. & Davis, M., 1988. Mon. Not. R. astr. Soc., 235, 715.
Efstathiou, G., Kaiser, N., Saunders, W., Lawrence, A., Rowan-Robinson, M. Ellis, R.S. & Frenk,
C.S., 1990. Mon. Not. R. astr. Soc., 247, lOP.
Kaiser, N., 1987. Mon. Not. R. astr. Soc., 227, 1.
Maddox, S.J., Efstathiou, G., Sutherland, W.J. & Loveday, J., 1990. Mon. Not. R. astr. Soc., 242,
43P.
Peacock, J.A. & Nicholson, D., 1991. For submission to Mon. Not. R. astr. Soc.
Peebles, P.J.E., 1987. Nature, 327, 210.
Salopek, D., Bond, J.R. & Bardeen, J.M., 1989. Phys. Rev. D., 40,1753.
Sutherland, W.J., 1988. Mon. Not. R. astr. Soc., 234, 159.
HIGHER MOMENTS OF THE IRAS GALAXY DISTRIBUTION

CALEB A. SCHARF,
IMtitute 0/ A.tronomy,
Cambridge University,
Madingley Road,
Cambridge CB3 OHA

Summary

The surface number density of lRAS galaxies is expanded in spherical harmonics. Maps
of the surface number density are then 'reconstructed' using only the first 120 expansion
coefficients. This provides a very 'natural' enhanced, smoothed picture of structure in the
IRAS angular galaxy catalogue.
Using a colour selected lRAS galaxy catalogue (Meurs & Harmon, 1988, Astr. Astro-
phys.206 53) with a flux limit of O.7Jy we obtain tV 11,000 galaxy candidates. The surface
density of these objects is expanded as
m=+1
u(8,¢) =.E .E a,mYim(8,¢) (1)
1=0 m=-I

where Yim are the standard, orthonormal set of spherical harmonics and a'm are estimated
by summing over all objects: a'm = Eff:l Yim(8;, <Pi) • Peebles «1973), Astrophys. J.,185,
Peebles & Hauser (1973), Astrophys. J.,185) discusses the relationship of a'm to correlation
functions and the angular power spectrum. To produce the map in Fig. 1 the a'm with
I = 1,2, ... , 10 were used to reconstruct u with an angular resolution", 7l' /10.
The equal area projections in Fig. 1 have a 'north' pole (left-hand plot) centered on
galactic I = 3070 and b = 90 , the direction of streaming motion as deduced by the Seven
Samurai. The left hand hemisphere is identical to that shown by Lynden-Bell (this volume)
but rotated anti-clockwise by 90 0 • Contours are equally spaced in number density and
dashed contours indicate u to be less than the mean (monopole). The galactic plane is
clearly seen horizontally across Fig. 1, along with part of the lRAS 'missing strip' on
the right-hand plot. The 'Great Attractor' and Perseus-Pisces regions dominate the plots,
along with a strong feature in Puppis. This feature is close to the galactic plane, but occurs
in a region of comparative optical transparency and a new redshift survey in this region
(Kraan-Korteweg, private communication) confirms the existence of nearby galaxies there.
Further work on spherical harmonic analysis is in progress (Scharf, Hoffman, Lahav &
Lynden-Bell, in preparation).

475
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 47~76.
© 1991 Kluwer Academic Publishers.
.j>.
FORNAX/ERIDANUS PERSEUS - PISCES -..l
01

.\
I '", \.

i ~/(' -'~'~~:~~'~:~:~':=~'-
-
\ \ .!\..\' , \ \ I.... ,"
---::>-. --~<-'J-' ....._-
'<.:-:::--::
~
........ ' . . . '-----,
~-
.. ------ ... :--- .............. -. - ---
:::: ~.- -----.- -"--
- ---...:
"::.:- ~~~-:.::~""'''::':-.::::~

VIRGO CENTAURUS/GREAT ATTRACTOR

FIGURE 1
COLLAPSE OF.A PROTOGALACTIC CLOUD

S. YOSHIOKA
Department of Physics,
Tokyo University of Mercantile Marine,
Etchujima, Koto, Tokyo 195, Japan

ABSTRACT. We investigate process of galaxy formation through collapse of a protogalactic cloud


consisting from gas and dark matter using three-dimensional hydrodynamical code.

1 Introduction

Conventional N-body methods can follow only formation of dark halos. In order to inves-
tigate formation process of visible galaxies we must follow evolution of gas destined to be
stars as well as dark matter. Hence, we perform three-dimensional simulations of galaxy
formation through collapse of systems consisting from gas and dark matter.

2 Model and Numerical Method

Dark matter is treated as a fixed dark halo, or as collisionless particles. We assume the
top-hat model: the cloud is spherical and uniform. In addition it rigidly rotates. Gas is
supposed to be isothermal. Calculations are started from states where the clouds have no
radial motion, which corresponds to the maximum expansion. We use the SPH code.

3 Cases that dark matter is treated as a fixed dark halo

Density distribution of a fixed dark halo is given using a solution of an isothermal sphere
cut by a certain radius rcut. Except for the parameters of the dark halo, parameters of the
present calculation are mass M gaa , initial radius Tg,O, temperature Tg , and initial angular
velocity Wi of the protocloud. However using some normalization they are reduced to
two, i.e., 0g,O =
(thermal energy of gas)/\potential energy of gas self-gravity\ and /3g,O =
(rotational kinetic energy of gas)/\potential energy of gas self-gravity\.
In the absence of a dark halo a collapsed gas cloud is known to gravitationally fragment
for small 00 and /30' The existence of the dark halo tends to extend the region of the cloud's
parameters, where it collapses to a stable state, to larger Og,O and /3g,o. In particular the
existence of stable thin disks is made possible. However, for parameters for which the cloud

477
T. Shanks et al. (eds.), Observational Tests o/Cosmological Inflation, 477-478.
© 1991 Kluwer Academic Publishers.
478

fragments in the absence of the dark halo, the stabilization of gas by the dark halo is not
effective, hence the cloud gravitationally fragments.
Next we investigate evolution of a more realistic cloud. Analytic and numerical works
on the origin of galaxy angular momentum based on the tidal torque theory show that
the dimensionless spin parameter oX '" 0.07 if dissipation can be neglected. This value
corresponds to a small f30. Thus, we simulated these small-po cases.
In the present simulation Mha/o, the dark halo mass, is 4.5 X 1011 M0 , the halo tem-
perature is one corresponding to the rotational velocity Vc = 200kms- 1 , the core radius
Tc is 9.0kpc, and Tcut = 5T c • On the other hand Mgaa = 5 X 101OM0' and Tg,O = 85kpc.
Wi is 2.0 X 10- 17 s-l, which corresponds to oX = 0.1. Final configurations for models with
various Tg are shown in Fig. 1 (from left Tg = 2 X 105 , 1 X 105 , and 5 X 104K). They can
be interpreted as the evolutional sequence of a single cloud if angular momentum loss can
be neglected: Soon after collapse, the gas cloud is heated to a high temperature (about the
virial temperature of the dark halo'" 106 K) and has a spheroidal configuration. As the
cloud temperature decreases by cooling, the cloud becomes flat. When the temperature
decreases to ;:;1 X 10 5 K, the cloud fragments.

4 Cases that dark matter is treated as collisionless particles

Initially dark matter particles are supposed to have a similar distribution and motion to gas
except for velocity dispersion ul. For small Tg and ul, the gas cloud fragments soon after
collapse while dark matter is violently relaxing. For large Tg and ul, the collapse of the
dark matter is weak and the gas cloud collapses to a stable disk or spheroid. In the cases
that fragmentation occurs, angular momentum is transferred from gas to dark matter.

20

-20

20 10
10

u
~ 0

-10
-20 -10
Index of Authors

Alimi, J .-M. . .............................. 469 Gutierrez, C. . ............................. 413


Banday, A.J. : ............................. 419 Guzman, R. . .............................. 193
Barbieri, C. . .............................. 461 Hale-Sutton, D. . .......................... 447
Beckman, J.E............................. .413 Hanes, David A. . .......................... 211
Berlin, A.B. . .............................. 437 Heavens, A.F .............................. 379
Bohringer, H. . ............................ 293 Henry, J.P................................. 293
Bouchet, F.R. . ............................ 469 Hoffman, Y. . .............................. 385
Boyle, B.J. . ................................ 309 Hoyle, F. . ................................. 273
Brandenberger, Robert .................... 39 Huchra, John P. . .......................... 315
Briel, U.G.................................. 293 Hudson, Michael J. . ....................... 467
Broadbent, A. . ........................... .447 Iovino, A. . ................................ .461
Broadhurst, Tom .......................... 449 Laflamme, Raymond ........................ 39
Bursov, N.N. . ............................. 437 la Franca, F. . ............................. 461
Buryak, O.E. . ............................. 453 Labav, O. . ................................. 375
Carr, B.J. . ................................. 103 Liddle, Andrew R. . ......................... 23
Carter, D. . ................................ 193 Lilje, P.B. . ................................. 375
Chepurnov, A.V........................... 437 Lilly, Simon J. . ............................ 233
Chernenkov, V.N. . ....................... .437 Lucey, J.R. ................................ 193
Clowes, R.G. . ............................. 461 Lynden-Bell, D. . .......................... 337
Coles, Peter ................................. 75 MacGillivray, H.T........................ .447
Collins, C.A. . ............................. .447 Major, J. . ................................. 205
Cowie, Lennox L. . ......................... 257 Martinez-Gonzalez, E ...................... 47
Criastiani, S. . ........................ 457, 461 Masi, S.................................... .443
Davies, R.D. . .............................. 413 Melchiorri, B. . ............................ 443
de Bernardis, P. . ......................... 443 Melchiorri, F. . ............................ 443
Dekel, Avishai ............................. 365 Melott, Adrian L. . ......................... 389
Deliyannis, Constantine P .................. 111 Metcalfe, N. . .............................. 187
Demarque, Pierre ......................... 111 Mijie, Milan ................................. 39
Demiatiski, M. . ........................... 453 Mingaliev, M.G ............................ 437
Devaney, N. . .............................. 205 Moss, Ian ................................... 53
Doel, P ..................................... 205 Moutarde, F ............................... 469
Doroshkevich, A.G. . ................. 327, 453 Myers, R. . ................................. 205
Dunlop, C. . ............................... 205 Naugolnaja, M.N. . ........................ 437
Dunlop, J.S ................................ 463 Nichol, R. . ................................ 447
Ebeling, H. . ............................... 293 Nizhelskij, N.A. . .......................... 437
Edge, A.C. . ................................ 293 O'Kane, P ................................. 205
Efstathiou, G. . ............................ 425 Oates, A.P. . ............................... 447
Ellis, Richard S. . ........................... 243 Oliver, Sebastian ........................... 449
Erukhimov, B.L ........................... 437 Parijskij, Yu. N. . .......................... 437
Feast, M.W. . .............................. 147 Parker, Q.A................................ 447
Fong, R ..•................................. 447 Peacock, J.A.......................... 463,471
Frenk, Carlos S. . ........................... 355 Peebles, P.J.E ............................... 63
Fukugita, M. . ............................. 267 PeDat, R. . ................................. 469
Georgantopoulos, 1. ....•...•...•..••••••. 309 Pierce, M.J. . .............................. 173
Giallongo, E. . ............................. 457 Plionis, Manolis ............................. 75
Giler, M. . ................................. 419 Primack, J.R. ............................. 375
Griffiths, R.E. . ............................ 309 Pritchet, C.J ............................... 199
Guiderdoni, B............................. 217 Pounds, K. . ............................... 309
Guth, Alan H. . .............................. 1 Rebolo, R. . ................................ 413

479
480

Redfern, M. • ...•.•........••...••..•....•• 205


Ree., M.J •... : •.•....•.......•....••..••••• 375
Rensini, Alvio .....•....................... 131
Rowan-Robinson, Michael ................ 161
Salopek, D.S. . .............................• 81
SaJUI, J.L. . ••......•.••.•...•....•..•••.....• 47
Sarajedini, Ata ..........•....•....•....••• 111
Sarkar, S. . ........•............•...•.....••• 91
Scharf, Caleb A. . ....••.....•...•....•..... 475
Schwarz, R.A. . .•......••..........•....••• 293
Shanks, T. . ...•..•..............• 187, 205, 447
Smoot, G.F ................................ 395
Starobinsky, A.A. . ........................ 437
Stewart, G.C •........•....•........•....... 309
Sutherland, Will ..•..............•...•.... 331
Szabelska, B. . •.................•..•....... 419
Szabelski, J ..•............................. 419
Tammann, G.A............................ 179
Tanvir, N. • •..••.....•.•••••••.......••.•.• 205
Terlevich, R.J. . ............................ 193
Verkhodanov, O.V. . ...•...•.............. 437
Voges, W. . ................................ 293
Watson, F.G............................... 447
Watson, R.A. . .•...•......•.....•..•......• 413
White, Simon D.M. . ....................... 279
Wolfendale, A.W. . .........•.....•........ 419
Yoshii, Y................................... 267
Yoshioka, S................................ 477
Subject Index geometry 258
models 233
timescale 111, 173
IT-keV neutrino 94 cosmology
21cm line widths 181 chaotic 8
quantum (see quantum: cosmology)
A (see also cosmological constant) 233 critical density 1
n (see also density parameter) 337, 355
DR - (1' method 163, 181, 193
Abell clusters 471 dark halo 477
action principle 66 dark matter
absorption spectra 455 halos 469
alignment effect 251, 463 non-baryonic 91,437
APM Galaxy Survey 331 dating techniques 113
axion 91 deceleration parameter 243
decoherence arguments 39
Baade-Wesselink 134, 152 deep galactic surveys 453
baryon-to-photon ratio 273, 276 densityiluctuations 469
bias 379 on the microwave sky 47
factor (b) 337, 365, 429 density parameter 279, 375, 379
biasing 64 baryonic 103
parameter (b) 356, 377 designer potentials 31
binaries 282 dipole measurements 349
black holes 103 distance estimates 133, 212
brown dwarfs 103 distance scale 199
domain walls 98
CDM (see also Cold Dark Matter) 48,369, 377, dwarf galaxies 259
382
linear 473 environmental dependence 193
standard 385,471 Euler Poincare characteristic 77
Cepheid evolution of clusters of galaxies 306
photometry 187 excursion set 75
sequences 187
Cepheids 147 filaments 390, 447
chemical abundances 111
CIV 455 galactic halos 103
cluster distances 180 galaxies at high redshifts 70
clusters of galaxies 69,293 galaxy
clusters 315 clustering 471
COBE 395, 414, 428 clusters 320
Cold Dark Matter (CDM) 105,425 counts 224, 243, 257
model 84 evolution 217,258,267,463
theory 64 formation 131, 413
color-magnitude diagram 112, 132 luminosity evolution 234
colour gradients 212 redshift survey 447
Coma ellipticals 193 galaxy-galaxy interactions 194
correlation function 331 Gaussian
cosmic 'drift' 450 lluctuations 85, 370, 403
Cosmic Microwave Background (see also random fields 76
Microwave Background) 338, 395, 419, 443 statistics 29
cosmic virial theorems 286 genus 75
cosmological globular cluster ages 131
constant 6, 231. 53, 175, 375, 430 globular clusters 111,179, 181,211

481
482

gravitational large-scale
instability 66, 365, 389 structure 68, 293, 304, 315, 327
potential 395 uniformity 1
potential fluctuations 473 lensed QSO's 450
wave background 31 lensed image separations 449
gravity waves 396 gravitino, light 94
Great Attractor 287, 337, 367, 379, 385, 467, photino, light 94
475 Limber's equation 472
great attractors 49 linear theory 376, 430
groups 282 lithium abundance 111
local
Ho (see also Hubble constant) 173, 179,418 distance scale 147
Harrison-Zel'dovich spectrum (see also flow 161
scale-invariant spectrum) 29, 39 Local Group 65, 169, 337
HDM (Hot Dark Matter) 50 luminosity evolution 243
helium luminosity function 211
abundance 148, 154 cluster 112
diffusion of 111 luminosity-line-width 173
Higgs fields 3, 4 Lya forest 457
higgsino, light 94
Hubble M31 147,188
constant 147, 179, 199, 205 M33 147,188
diagram 161, 220, 251 MACHO 103
expansion 1 macrolensing 105
flow 174 Magellanic Clouds 147
Hyades modulus 151 clusters 141
magnetic monopoles 1
image sharpening 205 main sequence fitting 134
inflation 1, 161, 275, 323 main sequence 151
chaotic 7 Malmquist-type bias 162
double 24 mass-to-light ratio 279
double-field 26 massive weakly interacting stable parti-
extended 26,31 cle 96
new 24 merging 217
old 26 metallicity 154
power law 24 microlensing 103, 105
R2 24 Microwave Background 277, 331, 361, 425,
super-Einstein 24 474
infrared background 103 anisotropy 413
initial spectrum 327, 332 correlation function 402, 421, 443
instability strip 147 dipole anisotropy 398
intergalactic absorption 457 isotropy 206
interstellar dust 161 power spectrum 425
intracluster gas 294 quadrupole anisotropy 402
IRAS Sachs-Wolfe effect 428
QDOT survey 472 secondary anisotropies 47
galaxies 169, 337, 355, 365, 379, 385 Sunyaev-Zel'dovich effect 413
galaxy distribution 475 missing mass problem 206
isochrones 112, 132 MMRD relation 199
isophotal magnitude 267 models 322
monopoles 98
K-band 257
neutralino 97
483

neutrino 93 'oscillation' 99
17-keV 94 rich clusters 284
heavy Dirac 98 ROSAT 293, 309
solar 1310 RR Lyrae 152
non-Gaussian distances 133
fluctuations 81, 389
perturbations 31 scalar fields 3
non-linear growth 469 scale invariance 469, 472
novae 181, 199 scale-invariant 39, 355, 365, 431
nucleosynthesis 103, 206 spectrum 23
number-magnitude relations 238 selection effects 179, 270
self-similarity 470
optical number counts 449 separated Hamilton-Jacobi equation 81
sheets 390, 447
pancakes 69, 328 shell-crossing 470
peculiar acceleration 356 sneutrino 97
peculiar motions 193 spectrum amplitude 29
peculiar velocity fields 467 Standard Model 91
peculiar velocities 357, 365, 376, 379, 395, star formation 253
431,471 stellar evolution 112
period-luminosity relation 148, 187 streaming motions 179
phase transitions 7, 91 strings 98
planetary nebulae 161, 182 supercluster 276, 332, 447
PLC relation 148 supernovae 181, 245
'Population III' stars 103 Supersymmetry 92
power spectra 390, 395, 471 surface brightness fiuctuations 161
power spectrum 335, 365, 385
primordial density technibaryon 97
perturbations 39 textures 98
perturbation spectrum 29, 413 timing argument 280
protogalactic cloud 477 topology 75, 360, 389
Tully-Fisher 147, 187
QSOs 309 infrared 161
high-redshift 457
Type Ia supernova 161
quantum
cosmology 53
universal age 206
fluctuations 8, 39 Ursa Major 173
tunnelling
Virgo 161
quasars (see also QSOs) 455
Cluster 173, 179, 199
intermediate-redshift 461
galaxy 205
number counts 461
Virgo centric "infall" 180
evolution, luminosity function 461
'V~ry Massive Objects' (VMOs) 104
evolution, volume density 461
VOids 48, 327, 390, 447, 453

radio galaxies 219, 251 white dwarfs 137


high-z 463
WIMP 103, 139
recombination 428 wormholes 57
redshift surveys 267, 315
relics X-ray background 312
asymmetric 97 X-ray observations 294
'freeze-out' 95
'plasma' 93
'soliton' 98

Você também pode gostar