Sunteți pe pagina 1din 327

Itay Neeman

The Determinacy of Long Games

Walter de Gruyter
Berlin New York
Author
Itay Neeman
Mathematics Department
University of California
Box 951555, 6334 MSB
LOS ANGELES, CA 90095-1555
USA
e-mail: ineeman@math.ucla.edu
Series Editors
Wilfrid A. Hodges Ronald Jensen
School of Mathematical Sciences Institut fr Mathematik
Queen Mary and Westfield College Humboldt-Universitt
University of London Unter den Linden 6
Mile End Road 10099 Berlin
London E1 4NS Germany
United Kingdom
Menachem Magidor
Institute of Mathematics
The Hebrew University
Givat Ram
91904 Jerusalem
Israel

Mathematics Subject Classification 2000: 03E60, 03E55, 03E47

Keywords: determinacy, large cardinals, infinte games

E
P Printed on acid-free paper which falls within the guidelines
of the ANSI to ensure permanence and durability

Library of Congress  Cataloging-in-Publication Data

Neeman, Itay, 1972


The determinacy of long games / by Itay Neeman.
p. cm.  (De Gruyter series in logic and its applications; 7)
Includes bibliographical references and index.
ISBN 3-11-018341-2 (cloth : alk. paper)
1. Game theory. 2. Determinants. 3. Logic, Symbolic and
mathematical. I. Title. II. Series.
QA269.N44 2004
519.3dc22 2004021609

ISBN 3-11-018341-2
ISSN 1438-1893
Bibliographic information published by Die Deutsche Bibliothek
Die Deutsche Bibliothek lists this publication in the Deutsche Nationalbibliografie;
detailed bibliographic data is available in the Internet at http://dnb.ddb.de.
 Copyright 2004 by Walter de Gruyter GmbH & Co. KG, 10785 Berlin, Germany.
All rights reserved, including those of translation into foreign languages. No part of this book may be
reproduced in any form or by any means, electronic or mechanical, including photocopy, recording, or
any information storage and retrieval system, without permission in writing from the publisher.
Printed in Germany.
Typesetting using the Authors TEX files: I. Zimmermann, Freiburg  Printing and binding: Hubert & Co.
GmbH & Co. KG, Gttingen  Cover design: Rainer Engel, Berlin.
To my parents, Avraham and Bruria
Preface

This book presents my research into proofs of determinacy for games of countable
length. (Further developments on long games, dealing with games of length 1 , are
presented in Neeman [33].) It has been a while since I started studying this topic; some
of the methods in Chapters 2 and 3, and certainly my interest in the subject, go all
the way back to my Ph.D. dissertation [30]. I am grateful to Tony Martin, John Steel,
and Hugh Woodin, first for bringing the study of interactions between determinacy and
Woodin cardinals into existence, and second for helping me join it. I am also grateful
to Ronald Jensen, for his work on fine structure which is crucial to other parts of my
research, and, in the context of this book, for introducing me both to large cardinals and
to proofs of determinacy.
My research was supported by several organizations: the University of California in
Los Angeles; the Society of Fellows at Harvard University; the Alexander von Humboldt
Foundation; and the National Science Foundation through grants DMS 98-03292 and
DMS 00-94174 (CAREER). I thank them sincerely.
There is some background on determinacy, and a synopsis of the current work, in
the introduction to this book. Let me here say a few words on the books structure. The
basic components for proofs of determinacy of long games are presented in Chapter 1.
Most important among them are the auxiliary games map associated to a given name,
and the concept of a pivot. Neeman [34] gives an informal view of these concepts,
and it may be useful as a starting point. (The definitions in Chapter 1 are for the map
A discussed in Section 5.1 of [34]. The map discussed in Section 2 of the paper is
essentially a special case.) Chapter 2 presents the first application, to games of fixed
countable length, and Chapter 3 presents a more elaborate application, to games of
continuously coded length. Neeman [34] gives special cases of these applications, and
again it may be a useful starting point.
Chapters 4, 5, and 6 develop tools for handling much longer games. These tools are
applied in Chapter 7 to games ending at 1 in L of the play. The chapter is written so as
to use only the end results in Chapters 5 and 6, and the relevant concepts. These results
and most of the concepts appear in Sections 4A, 4B, 4D (4), 4E (7), 5G, 6A, and 6G. In
a first reading it may be useful to skim through Chapters 4, 5, and 6, concentrating on
these particular sections, and then continue to Chapter 7.
As far as prerequisites go, the work here should be accessible to any reader with
a knowledge of basic set theory, say from Jech [10] or Kunen [15], some familiarity
with Silver indiscernibles, say from Jech [10, 18] or Kanamori [11, 9], and some
familiarity with extenders and iteration trees, say from MartinSteel [18], [19]. For the
background knowledge of iteration trees the reader may also consult Appendix A, but
the exposition there is very brief.
Let me now leave you with the book. I hope you find it both useful and pleasant.

Los Angeles, California, October 2004 Itay Neeman


Contents

Preface vii

Introduction 1

1 Basic components 15
1A The auxiliary games map . . . . . . . . . . . . . . . . . . . . . . . . 16
1A (1) Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1A (2) The rules of the game . . . . . . . . . . . . . . . . . . . . . . 20
1B Generic runs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1C Pivots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
1C (1) The game Apiv [x] . . . . . . . . . . . . . . . . . . . . . . . 28
1C (2) Constructing piv [, x] . . . . . . . . . . . . . . . . . . . . . 30
1C (3) Properties of piv . . . . . . . . . . . . . . . . . . . . . . . . 35
1D Mirror images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
1E Sample application . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
1F Mixed pivots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
1F (1) The game . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
1F (2) The strategy mix [, x] . . . . . . . . . . . . . . . . . . . . . 48
1F (3) Properties of mix . . . . . . . . . . . . . . . . . . . . . . . . 49

2 Games of fixed countable length 51


2A General games and iteration games . . . . . . . . . . . . . . . . . . . 51
2B Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2B (1) The basic definitions . . . . . . . . . . . . . . . . . . . . . . 57
2B (2) I wins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2B (3) II wins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2B (4) The third case . . . . . . . . . . . . . . . . . . . . . . . . . . 66
2B (5) Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2C Successors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2D Limits again . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2D (1) II wins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
2D (2) I wins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
2D (3) Determinacy . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2E Universally Baire sets . . . . . . . . . . . . . . . . . . . . . . . . . . 82

3 Games of continuously coded length 87


3A Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3B First determinacy result, part I . . . . . . . . . . . . . . . . . . . . . 89
3B (1) Names . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3B (2) The basic step . . . . . . . . . . . . . . . . . . . . . . . . . . 93
x Contents

3B (3) I wins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
3B (4) Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
3B (5) Internal ultrapowers vs. copying . . . . . . . . . . . . . . . . 106
3C First determinacy result, part II . . . . . . . . . . . . . . . . . . . . . 109
3C (1) II wins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
3C (2) Determinacy . . . . . . . . . . . . . . . . . . . . . . . . . . 112
3D A slight improvement . . . . . . . . . . . . . . . . . . . . . . . . . . 119
3D (1)  02 functions . . . . . . . . . . . . . . . . . . . . . . . . . . 120
3D (2)  01+ functions . . . . . . . . . . . . . . . . . . . . . . . . . 124
3E Variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
3E (1) A sketch of the proof . . . . . . . . . . . . . . . . . . . . . . 129

4 Pullbacks 135
4A Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
4B Woodins extender algebra . . . . . . . . . . . . . . . . . . . . . . . 141
4C Names, part I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4C (1) Removing obstructions . . . . . . . . . . . . . . . . . . . . . 155
4C (2) Relative successors . . . . . . . . . . . . . . . . . . . . . . . 157
4C (3) Relative limits and records . . . . . . . . . . . . . . . . . . . 158
4D Names, part II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
4D (1) Woodin limits of Woodin cardinals . . . . . . . . . . . . . . . 162
4D (2) Relative successors . . . . . . . . . . . . . . . . . . . . . . . 165
4D (3) Compositions . . . . . . . . . . . . . . . . . . . . . . . . . . 166
4D (4) Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
4E Mirror images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
4E (1) Removing obstructions . . . . . . . . . . . . . . . . . . . . . 167
4E (2) Mirrored successor game . . . . . . . . . . . . . . . . . . . . 168
4E (3) Relative limits and records . . . . . . . . . . . . . . . . . . . 169
4E (4) Woodin limits of Woodin cardinals . . . . . . . . . . . . . . . 172
4E (5) Relative successors . . . . . . . . . . . . . . . . . . . . . . . 173
4E (6) Compositions . . . . . . . . . . . . . . . . . . . . . . . . . . 174
4E (7) Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

5 When both players lose 175


5A Saturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
5B Successors, basic step . . . . . . . . . . . . . . . . . . . . . . . . . . 181
5C Relative limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
5C (1) Basic step . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
5C (2) Construction . . . . . . . . . . . . . . . . . . . . . . . . . . 187
5D Woodin limits of Woodin cardinals . . . . . . . . . . . . . . . . . . . 194
5D (1) Basic step . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
5D (2) Impossibility . . . . . . . . . . . . . . . . . . . . . . . . . . 197
5D (3) Another impossibility . . . . . . . . . . . . . . . . . . . . . . 199
5D (4) Construction . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Contents xi

5E Relative successors . . . . . . . . . . . . . . . . . . . . . . . . . . . 204


5F Compositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
5G Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

6 Along a single branch 209


6A The game . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
6A (1) How to leap . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
6A (2) The skipping game . . . . . . . . . . . . . . . . . . . . . . . 218
6B Successors, basic step . . . . . . . . . . . . . . . . . . . . . . . . . . 220
6C Relative limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
6C (1) Finite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
6C (2) Infinite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
6C (3) Construction . . . . . . . . . . . . . . . . . . . . . . . . . . 228
6D Woodin limits of Woodin cardinals . . . . . . . . . . . . . . . . . . . 235
6D (1) Hopeful . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
6D (2) Not hopeful . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
6D (3) Combined . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
6E Relative successors and compositions . . . . . . . . . . . . . . . . . 244
6F Skips . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
6F (1) Routes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
6F (2) Extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
6F (3) Closing skips . . . . . . . . . . . . . . . . . . . . . . . . . . 258
6F (4) Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
6F (5) Safe positions . . . . . . . . . . . . . . . . . . . . . . . . . . 262
6F (6) The strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
6G Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266

7 Games which reach local cardinals 268


7A Shifted payoff . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
7B Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
7B (1) Extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
7C Basic step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
7C (1) Obstruction free . . . . . . . . . . . . . . . . . . . . . . . . 280
7C (2) I-acceptably obstructed . . . . . . . . . . . . . . . . . . . . . 281
7C (3) Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
7D Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
7E The main theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
7F Determinacy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292

A Extenders, generic extensions, and iterability 301

Bibliography 309

Index 313
Introduction

The determinacy of infinite games has been central to the development of modern set
theory. From its humble, anecdotal beginnings in the 1930s and 50s, the subject of
determinacy has grown to provide the dividing line between the realm of definable sets
of reals and the realm of the axiom of choice. It has had conceptual influences on the
study of forcing, and substantial concrete influences on the study of large cardinals.
The basic definitions are quite simple. Let C , that is let C be a set of infinite
sequences of natural numbers. Define G (C), the length game with payoff set C,
to be played as follows: Players I and II collaborate to produce an infinite sequence
x = $x(i) | i < % of natural numbers. They take turns as in Diagram 1, I picking x(i)
for even i and II picking x(i) for odd i. If at the end the sequence x they produce belongs
to C then player I wins; and otherwise player II wins. The game G (C), or any other
game for that matter, is determined if one of the two players has a winning strategy,
namely a strategy for the game that wins against all possible plays by the opponent.
The set C is said to be determined if the corresponding game G (C) is determined.

I x(0) x(2) ......


II x(1) x(3) ......

Diagram 1. The game G (C).

For illustrative purposes it is helpful to express determinacy by means of logical


quantifiers. The statement that player I has a winning strategy in G (C) is the natural
interpretation of the quantifier string:

x(0) x(1) x(2) x(3) . . . . . . . . . $x(i) | i < % C (1)

where the quantifiers range over natural numbers. The statement that II has a winning
strategy on the other hand is the natural interpretation of the string:

x(0) x(1) x(2) x(3) . . . . . . . . . $x(i) | i < %  C (2)

where again the quantifiers range over natural numbers. Determinacy then is the state-
ment that one of the equations holds, in other words that equation (2) is the negation of
equation (1). This would simply be a matter of logic if the quantifier strings were finite.
But the strings here are of length , and determinacy is far from trivial. Indeed, one
cannot expect determinacy to hold for all sets: using the axiom of choice, more specif-
ically using a wellordering of the real line, a straightforward construction produces a
non-determined set C. Surprisingly, it has turned out that one can expect determinacy
to hold for concrete sets of reals, meaning sets which are definable, over the real line
and in fact more.
2 Introduction

Following standard usage let < denote the set of finite sequences of natural
numbers. For s < let Ns = {x | x extends s}. The sets Ns (s < ) are
the basic open neighborhoods in , and a set C is open if it is a union of basic
open neighborhoods. Similar definitions can be made for the product spaces ( )k
(though strictly speaking there is no need for this, since these spaces are isomorphic
to ). A set D is analytic if it is the projection of the complement of an
open set. In other words D is analytic if there is an open C ( )2 so that
x D (y )$x, y%  C. The class of analytic sets begins a hierarchy
of definability, from open sets, through applications of complementations and real
quantifiers. (A real quantifier here means a quantifier over . The reason for the
terminology is that is isomorphic to the real line minus the rational numbers. In
descriptive set theory R often means , for the same reason.) In general a set D
is said to be  1n if there is an open set C ( )n+1 so that:

(y1 )(y2 ) . . . . . . (yn )$x, y1 , . . . , yn %  C if n is odd; and
x D
(y1 )(y2 ) . . . . . . (yn )$x, y1 , . . . , yn % C if n is even,

where the quantifiers in both cases range over . The analytic sets are thus the  11
sets. A set is said to be 1n if its complement is  1n , and said to be 1n if it is both  1n
and 1n . (The lightface versions n1 , 1n , and 1n are defined similarly, only replacing
open, which means an arbitrary union of basic neighborhoods, by its lightface parallel,
namely a recursively enumerable union of basic neighborhoods.) A set is projective if
it is  1n for some n < , in other words if it can be obtained from open and closed sets
through the application of finitely many real quantifiers.
The point of introducing here the definitions of open, analytic, and projective sets
is that these sets are somehow more concrete than the general sets one can obtain from
the axiom of choice, and the kind of argument that uses a wellordering of the reals to
contradict determinacy should not apply to them. These sets could, conceivably, be
determined.
At the start of the projective hierarchy stand the open sets, and in 1953 GaleStewart
[8] proved that they are determined. The proof uses the kind of quantifier reasoning that
works for finite games, and adds to it a convergence argument. Fourteen years after its
publication, Blackwell [2] used the determinacy of open games to give a new proof of
Kuratowskis reduction theorem for analytic sets [16], a theorem which states that the
intersection of any two analytic sets A, B R can be presented as the intersection of
two analytic sets A A and B  B, such that A B  = R. Inspired by the method of
his proof, Martin [20] and AddisonMoschovakis [1] proved a similar theorem for 31
sets, assuming the determinacy of 12 sets. Moreover they obtained for the 13 level the
same structural property of the 11 level which is used to prove Kuratowskis theorem
for 11 sets, and in fact propagated this property through all the odd projective levels
assuming projective determinacy.
There have been earlier results on sets of reals proved using infinite games, see
Oxtoby [36], Davis [4], and MycielskiSwierczkowski [28]. These papers derived
regularity properties, Lebesgue measurability in the case of MycielskiSwierczkowski
Introduction 3

[28], for sets of reals assuming determinacy. But the work of Martin and Addison
Moschovakis was the first to apply determinacy to a structural property of a class of
definable sets. It gave a substantial push to the doctrinean elaboration on the proposal
in MycielskiSteinhaus [27]of assuming determinacy in the study of definable sets
of reals. Additional research following this doctrine gradually established a very rich
structure theory for classes of definable sets of reals, see for example Moschovakis
book [26].
To actually bring this theory to bear on a particular level of definability one had to
prove the relevant determinacy. The open games determinacy of GaleStewart [8] and
the somewhat stronger results of Wolfe [43] and Davis [4] on low levels of the Borel
hierarchy were only enough to bring the structure theory to bear on the level of analytic
sets, where it was anyway known classically. Even full Borel determinacy (which would
be proved in ZFC by Martin [22] some years later) would not lift the theory further.
Applying the theory to levels beyond analytic required proving determinacy beyond
Borel games.
There were some vague indications at the time that determinacy may be connected
to large cardinal axiomsaxioms which state the existence of cardinals with various
reflection properties not provable in ZFC. Solovay had shown that determinacy for all
subsets of (a statement which contradicts the axiom of choice) implies that 1 carries
a total, countably complete, 2-valued measure. Cardinals which carry such measures
are called measurable, and in the context of the axiom of choice their existence is a
large cardinal axiom. In the other direction Solovay had proposed that large cardinal
axioms (the ones he had in mind were much stronger than measurable) should imply
projective determinacy.
It came as a very pleasant surprise when, not long afterwards, Martin [21] proved
11 determinacy from a measurable cardinal. This was the first, and at the time only,
proof of determinacy from a large cardinal axiom. Its ideas gave rise to several impor-
tant concepts and methods, including the notion of homogeneously Suslin sets which is
now useful both in proofs of determinacy and in the study of consequences of determi-
nacy. Moreover it turned out through work of Martin and Harrington that the proof is
essentially optimal. A large cardinal axiom of roughly the magnitude of a measurable
cardinal is necessary for a proof of 11 determinacy, see Harrington [9].
11 determinacy allowed extending some of the regularity results previously known
for analytic sets, to the level of  12 sets. But an extension of the deeper structural results
to levels beyond analytic required at least the determinacy of games with 12 payoffthe
kind of games used in the results of Martin [20] and AddisonMoschovakis [1] on the
13 leveland an extension to all projective levels required projective determinacy. The
first proof of 12 , and in fact 12 determinacy was found in 1978 by Martin [23]. Some
six years later Woodin (1984) proved projective determinacy. But there was something
missing. Both proofs used extremely strong large cardinal axioms. It was expected,
certainly at the time of Martins proof and for a few brief months after Woodins proof,
that these axioms should prove necessary, just as the axioms used by Martin to prove
11 determinacy proved necessary through Harrington [9]. But no one was able to show
4 Introduction

this, and in fact it turned out that much weaker axioms suffice.
The appropriate axioms for projective determinacy were discovered in 1984 through
a now legendary process, involving a very fast succession of results obtained by Fore-
man, Magidor, Shelah, and Woodin during a time of close communications.
The process was initiated by the work of ForemanMagidorShelah [6], which
among other things dealt with generic elementary embeddings. These embeddings are
generated by forcing with the poset P ()/I , where is an infinite cardinal and I is a non-
trivial ideal over . A generic filter for this poset gives rise to an ultrafilter over , which
in turn gives rise to an ultrapower M of the universe V and an elementary embedding
j : V M. Properties of the ideal I translate to properties of the embedding j and
its target M. For example if I is -complete then the critical point of j is precisely .
If I is also + -saturated (meaning that P ()/I has no antichains of size + ) then the
ultrapower M is closed under sequences of length in the generic extension, and in
particular it is wellfounded.
For the discussion below let ideal mean 1 -complete, non-trivial ideal. The dis-
cussion centers on the additional property of saturation. Prior to the work of Foreman
MagidorShelah [6] it was known that the existence of an 2 -saturated ideal over 1
can be forced assuming a large cardinal axiom called huge. The forcing, due to
Kunen [14], involved making a huge cardinal of V into 1 of the forcing extension
(in particular collapsing 1V ), and extending a hugeness embedding from V to the kind
of ultrapower embedding one would get from an 2 -saturated ideal over the new 1 .
ForemanMagidorShelah [6, 1,2] improved on this in several ways. They reduced
the large cardinal assumption substantially, from huge to the existence of a cardinal
known as supercompact; they managed to deal specifically with NS 1 , the ideal of non-
stationary subsets of 1 ; and most remarkably they found a way to force and make
this ideal 2 -saturated without collapsing 1V . It followed that a supercompact cardi-
nal outright implies the existence of generic elementary embeddings with critical point
1V and wellfounded targets (reached by first passing to the ForemanMagidorShelah
extension, and then forcing with P (1 )/NS 1 ). These were embeddings of a substan-
tially new nature, obtained not as extensions of large cardinal embeddings from smaller
models, but through a true use of forcing.
There had been results, see Foreman [7] and Magidor [17], connecting generic ele-
mentary embeddings with regularity properties for definable sets of reals. Magidor [17]
had proved that all  13 sets are Lebesgue measurable using large cardinals and a generic
elementary embedding with critical point 1V and a wellfounded target. His argument
combined with the results described above to prove Lebesgue measurability for  13 sets,
from a supercompact cardinal. The connection with higher levels of definability was
made by Woodin. Through a modification of his earlier work [44, Theorems 58] he
showed that all projective sets are Lebesgue measurable, assuming the existence of a
supercompact cardinal, and using the ForemanMagidorShelah theorem that one can
force to obtain an 2 -saturated ideal over 1 without collapsing 1 (in fact just the
consequence that there are generic elementary embeddings with critical point 1V and
wellfounded targets). He also noted that an easier argument [6, pp. 2728] would show
that all subsets of in L(R) are Lebesgue measurable, if only there was a forcing
Introduction 5

which obtains saturation without adding realsa stronger demand than that of not col-
lapsing 1 . Shelah at the same time realized that techniques from his book [38] could be
used to adjust the construction of [6, 1,2] and produce such a forcing [6, Theorem 21,
Proposition 22]. Thus it was shown from a supercompact cardinal that not only all
projective sets, but moreover all subsets of in L(R), are Lebesgue measurable.
The fact that supercompact cardinals imply Lebesgue measurability for all projec-
tive sets hinted that the strength of projective determinacy need not pass a supercompact,
much less than the axioms used by Martin and Woodin for 12 and projective deter-
minacy. Moreover it was clear that even less than a supercompact should suffice for
the proof of Lebesgue measurability. Woodin and Shelah began taking turns weak-
ening the hypothesis. First Woodin noticed that a superstrong cardinal, rather than a
supercompact, is enough. Then Shelah isolated a weaker concept now known as a She-
lah cardinal, and showed that the existence of n + 1 Shelah cardinals (still much less
than a single superstrong cardinal) implies that all  1n+2 sets are Lebesgue measurable.
Shelahs proof of Lebesgue measurability used a certain forcing property, rather than
directly using a large cardinal property. This property followed from the existence of
n + 1 Shelah cardinals, but it was clear that even less would do. The precise nature
of the end assumption was discovered by Woodin. By inverting quantifiers in Shelahs
definition he isolated a weaker concept now known as a Woodin cardinal. He then
showed that the weaker concept is still sufficient for a proof of Lebesgue measurability.
The end result, published in ShelahWoodin [37], was that the existence of n Woodin
cardinals and a measurable cardinal above them implies that all  1n+2 sets are Lebesgue
measurable. It was obtained during a conversation in Jerusalem, just a couple of weeks
after the original impetus of [6].
The work of Foreman, Magidor, Shelah, and Woodin described above led to many
important advances on forcing relative to large cardinals. From our perspective though
the most crucial discoveries reached through this work were first the isolation of the
concept of a Woodin cardinal (see Definition 1A.11), and second the ShelahWoodin
theorem that the existence of n Woodin cardinals and a measurable cardinal above them
implies the kind of Lebesgue measurability that would follow from 1n+1 determinacy.
Equipped with an expectation created by these two discoveries, Martin and Steel
set out to prove 1n+1 determinacy from n Woodin cardinals and a measurable cardinal
above them. Their approach was motivated by considerations from inner model the-
ory. Inner models are generalizations of Gdels L, and like L they have an internal
wellordering of their reals. In the case of L the wellordering is 12 , and from this it
follows that in L not all  12 sets are Lebesgue measurable. In richer inner models the
internal wellorderings are more complex, but in all inner models considered up to the
work of MartinSteel [18], [19] the complexity was at most 13 . In particular these
inner models could not satisfy the statement all  13 sets are Lebesgue measurable. In
light of the ShelahWoodin result the inner models were limited to at most a Woodin
cardinal, a much lower level than previously expected.
Martin and Steel began their investigation trying to understand the reason for this
limitation. They discovered (though this was done after the proof of projective deter-
minacy) that it had to do with the linear nature of the iterated ultrapowers used in inner
6 Introduction

O
O
MJ 7

MO 6 

o 7 M5 
o
ooo 
MO 4 

 ? M3
j2,4

MO 2  j1,3

oo7
j0,2 M1
oo
o j0,1
M0

Diagram 2. A sample iteration tree, with the tree order 0 T 1, 0 T 2, 1 T 3, 1 T 7, . . . .

model theory at the time. Inner model theory at the level of a Woodin cardinal and be-
yond is now known to require non-linear iterated ultrapowers, which Martin and Steel
called iteration trees. Such non-linear iterations involve some choices at limits, and it
is the need for these choices that adds complexity to the internal wellorderings of the
reals. The added complexity in the case of finitely many Woodin cardinals corresponds
precisely to the ShelahWoodin theorem: MartinSteel [19] produced models with n
Woodin cardinals and a 1n+2 wellordering of their reals; ShelahWoodin [37] showed
that if the models were to contain a bit more, namely an extra measurable cardinal above
the n Woodin cardinals, they would not have a 1n+2 wellordering of their reals.
A precise, though brief, definition of iteration trees is given in Appendix A, and
more details may be found in MartinSteel [18, 3]. For the sake of this introduction,
only iteration trees of length are needed. Roughly speaking an iteration tree of length
on a model M consists of:
a tree order T on (see Section 7B for the definition of a tree order);
a sequence of models $Mk | k < % with M0 = M; and
embeddings jk,l : Mk Ml for k T l.
Each individual step along an iteration tree is created through an ultrapower by an
extender (see Appendix A for a brief definition and further references). This is illustrated
in Diagram 3. Ml+1 is generated as an ultrapower by an extender El Ml . El is
applied to a model Mk , for a k l. The T -predecessor of l + 1 is set equal to k,
and jk,l+1 : Mk Ml+1 is set equal to the ultrapower embedding by El . In a linear
iteration El would be applied to Ml , in other words k would have to always equal l.
Introduction 7

Ml+1 = Ult(Mk , El )
O

jk,l+1 El Ml

Mk

Diagram 3. Forming Ml+1 .

But in an iteration tree k < l is allowed too, provided that Mk and Ml are in sufficient
agreement that the ultrapower Ult(Mk , El ) makes sense.
A cofinal branch through an iteration tree of length is an infinite set b which
is linearly ordered by T . For example the sample iteration tree of Diagram 2 has an
even branchthe branch consisting of {0, 2, 4, 6, . . .}. Given a cofinal branch b through
T let Mb denote the direct limit of the models and embeddings of T along b, more
precisely the direct limit of the system $Mk , jk,l | k T l b%.
Martin and Steel had not yet developed the theory of iteration trees in the spring
of 1985. But Steel had realized that non-linear iterations may help in the comparison
process for inner models, and began to use them in a limited way. Among the various
forms of non-linearity one should be prepared to tackle in developing inner model theory,
perhaps the simplest is that of an alternating chains: an iteration tree with precisely two
branches, an even branch consisting of {0, 2, 4, 6, . . .}, and an odd branch consisting of
{0, 1, 3, 5, . . .}. It seemed natural to conjecture that in an alternating chain on V at least
one of the branches should have a wellfounded direct limit (in fact something of the sort
seemed necessary for any development of inner model theory). During a conversation
on this in the department lounge at UCLA, Martin and Steel came to suspect that this
matter of getting the wellfoundedness of one direct limit from the illfoundedness of
the other may potentially be related to the problem of finding a homogeneous Suslin
representation for the complement of a given homogeneously Suslin set. They then
focused their attention on this problem.
Homogeneous Suslin representations were abstracted in the late 70s by Kechris [13]
and independently Martin, and trace back to Martins earlier proof of 11 determinacy
[21]. That proof can be broken into two parts:
(1) a proof that homogeneously Suslin sets are determined; and
(2) a proof from measurable cardinals that 11 sets are homogeneously Suslin.
If one could somehow show that all projective sets are homogenously Suslin then by
(1) projective determinacy would follow. The intuition that alternating chains could
potentially be useful for this proved accurate, and by the end of the summer Martin
and Steel constructed homogeneous Suslin representations for  11 sets, namely for the
complements of the sets covered by (2), using alternating chains. More generally they
found a construction which propagated homogeneous Suslin representations along the
projective hierarchy, using iteration trees. At each level of the hierarchy propagation
8 Introduction

to the next level required a Woodin cardinal, used to supply enough extenders for the
creation of the iteration trees. Combining the propagation with (1) and (2) Martin
and Steel then obtained 1n+1 determinacy from n Woodin cardinals and a measurable
cardinal above them, precisely the result they hoped for given the ShelahWoodin
theorem.
The proof of projective determinacy was published in MartinSteel [18], and the
subsequent development of the general theory of iteration trees was published in Martin
Steel [19]. The two papers had a dramatic effect on the study of large cardinals, opening
the doors to new avenues of research. Particularly notable for the purpose of this book
is a result of Woodin, reached as part of a proof of the consistency of 12 determinacy
from the consistency of one Woodin cardinal (avoiding the measurable cardinal used
by MartinSteel [18]). It was motivated partly by descriptive set theory, partly by
the developing study of iteration trees and inner models for Woodin cardinals, and,
curiously, by a classical theorem of Vopenka.
Recall that HOD denotes the class of sets which are hereditarily ordinal definable.
Vopenkas theorem states that every real of V is generic over HOD. Relativized to L[x],
where x is a real, the theorem implies in particular that x is generic over HODL[x] .
Assuming 12 determinacy Woodin expected that on a cone of x, meaning for all x of
sufficiently large Turing degree, HODL[x] should be a model with one Woodin cardinal,
equal to 2 of L[x]. (This he later proved true, see BurkeSchimmerling [3].) In fact it
seemed reasonable to expect that HODL[x] should be an iterate of M1 on a cone, where
M1 is the minimal iterable class model for one Woodin cardinal. This would explain the
fact that under 12 determinacy the theory of HODL[x] in parameter 2 L[x] is constant
on a cone (it would be equal always to the theory of M1 in parameter , where is the
Woodin cardinal of M1 ) and also allow proving 12 determinacy from the existence of
M1 , since having the theory of HODL[x] in parameter 2 L[x] constant on a cone implies
12 determinacy, assuming 1 is inaccessible to reals. If this appealing connection
between HODL[x] and M1 were to hold, then by Vopenka, x would be generic over an
iterate of M1 .
Inspired by this chain of thoughts, Woodin [45] made a surprisingly general discov-
ery. Given any model M with a Woodin cardinal, he defined a poset P M which can,
through an iteration, be made to admit any real as generic. The poset is now known as
Woodins extender algebra, and the process by which it is made to admit a given real as
generic is known as Woodins genericity iteration. The genericity iteration shares some
qualities with the comparison process for inner models, and just like the development
of inner models, it requires non-linearity.
Woodins extender algebra and genericity iterations proved extremely useful in the
study of large cardinals. Their original use in [45] helped settle the consistency strength
of 12 determinacy (though not quite through the route described above, and the precise
connection between HODL[x] and M1 is still open). Later uses involved proofs of
generic absoluteness, results connecting large cardinals with descriptive set theory, and
most recently Woodins core model induction, an intricate method for extracting models
with large cardinals from various strong statements about V. In fact Woodins extender
algebra is used also in this text, as we shall see below.
Introduction 9

Let us return now to the matter of proofs of determinacy. There are two possible
directions for extensions of the projective determinacy of MartinSteel [18]. The first
involves keeping the length of the games at , and allowing payoff sets more compli-
cated than projective. Working in this direction Woodin (1985) proved from Woodin
cardinals and a measurable cardinal above them that all subsets of in L(R) are de-
termined. He had already proved earlier that these, and even more complicated subsets
of , are weakly homogeneously Suslin (see Woodin [46] for the relevant definition
and the proof) assuming the existence of a supercompact cardinal. For sets in L(R)
he had reduced the assumption to infinitely many Woodin cardinals and a measurable
cardinal above them. As for the conclusion, a direct application of the MartinSteel
propagation [18] allows obtaining homogeneous Suslin representations, and hence de-
terminacy, from weakly homogenous Suslin representations. Woodins theorem then
gives determinacy for all subsets of in L(R). More generally, to each model M with
infinitely many Woodin cardinals, and each limit of Woodin cardinals in M, Woodin
associated a certain pointclass  and showed that all subsets of in L(R , ) are
determined, where R are the reals of a symmetric collapse of over M. This is the
definitive result in the direction of games of length with increasing payoff complexity.
The second way to progress beyond projective determinacy involves keeping the pay-
off at the level of projective sets (11 sets even), and allowing games of lengths greater
than . It is this second direction, the direction of long games, that is treated in this book.
The current methods for proofs of determinacy of long games trace back to the
work of Neeman [29], [32] on optimal proofs of projective determinacy. The result
of MartinSteel [18], that 1n+1 determinacy follows from the existence of n Woodin
cardinals and a measurable cardinal above them, is in spirit optimal; the construction
of models with n Woodin cardinals and a 1n+2 wellordering of their reals in Martin
Steel [19] shows that n Woodin cardinals by themselves are not sufficient for a proof of
1n+1 determinacy. Still, the extra measurable cardinal used in MartinSteel [18] is an
overkill, and one should be able to do with less. Let ()n denote the statement that over
every real there exists a countable, iterable model with a sharp for n Woodin cardinals.
(Sharps are weakenings of measures, introduced by Silver [39], [40]. Iterability has to
do with iteration trees, see below for more details.) ()n is the natural weakening of the
assumption used in MartinSteel [18], to a statement that should be outright equivalent
to 1n+1 determinacy. Woodin (1989) proved the equivalence for odd n. For even
n > 0 the direction 1n+1 determinacy implies ()n was proved by Woodin (1995).
The direction ()n implies 1n+1 determinacy was established in 1994 by Neeman [29],
[32] (the work there applies to all n > 0, both even and odd) using techniques which
build on the determinacy proof of MartinSteel [18].
Neeman [29], [32] showed that ()n in fact implies determinacy for a pointclass
larger than 1n+1 . Combined with Woodins result that 1n+1 implies ()n this led
to the NeemanWoodin transfer theorem on determinacy (see [32]), generalizing the
transfer theorem [12, Theorem 4] of KechrisWoodin. But for the purpose of this book
most important is the fact that Neeman [29], [32] began to connect moves in long games
on with moves in iteration games.
10 Introduction

Iteration games, introduced by MartinSteel [19], are games in which two players,
good and bad, collaborate to construct iteration trees and branches through them,
starting from a given model M. Bad plays the trees, and good picks the branches.
The payoff condition involves wellfoundedness. If ever good picks a branch which
leads to an illfounded direct limit, she loses. If she manages to always maintain well-
foundedness, she wins. M is said to be iterable if good has a winning strategy in the
iteration game starting from M.
The precise definitions of the iteration games relevant to the developments in this
book are included in Appendix A. For now a very restricted version suffices. Given
a model M and a number n < define W Gn (M) to be the following game between
the players good and bad: In mega-round k bad plays a length iteration tree
Tk on Mk (on M0 = M if k = 0). Good then plays a cofinal branch bk through Tk .
Mk+1 is set equal to the direct limit along this branch, allowing the players to proceed
to mega-round k + 1. They continue this way through mega-round n 1, producing
an iteration of the kind displayed in Diagram 4. If Mn , the final model they reach, is
illfounded, then bad wins; if Mn is wellfounded, good wins.

kk b0 kk b1 k
/ Mn1 kSkSkSkS bn1
M0 kSk
Sk Sk
Sk
SSS / M1 k SSS / M2 S
/ Mn
T0 T1 Tn1

Diagram 4. A weak iteration of length n.

Define M to be n-iterable if good has a winning strategy in W Gn (M). n-iterability


is then a restricted version of the full iterability informally discussed above. Let ()n be
the statement that over every real there exists a countable, n-iterable model with a sharp
for n Woodin cardinals. Recall that the statement of ()n above is identical except that
it demands full iterability. Certainly then ()n implies ()n .
Working with n < and a set C ( )n define Gn (C) to be played as follows:
Players I and II alternate playing natural numbers according to the format in Diagram 5,
producing reals yk = $yk (i) | i < % for k < n. If $yk | k < n% C then player I
wins; and otherwise player II wins.

I y0 (0) y0 (2) . . . . . . . . . . . . yn1 (0) yn1 (2) . . .


II y0 (1) ... ......... yn1 (1) ...
     
moves to produce y0 moves to produce yn1
  
n repetitions

Diagram 5. The game Gn (C).

It is easy to see that the determinacy of G(n+1) (C) for all C in 11 implies the
determinacy of G (D) for all D in 1n+1 ; this is because the n outermost quantifiers
Introduction 11

used in the 1n+1 definition of D can be simulated through the moves for y1 , . . . , yn in
G(n+1) (C).
Neeman [29], [32] reaches 1n+1 determinacy from ()n essentially by proving the
determinacy of G(n+1) (C) for every C in 11 , from ()n . The proof is structured as
an induction on n, and never literally handles G(n+1) (C); rather it handles the first
moves in this game and then appeals to induction for the part involving the remaining
n moves. But by joining the inductive steps together one can see that in essence
G(n+1) (C) is reduced to the iteration game W Gn (M) on the model M given by ()n .
The reduction either matches player I in G(n+1) (C) to good in W Gn (M), and in
this case player I ends up winning G(n+1) (C), or else it matches II to good, and
in this case II ends up winning. In either case it is ultimately a strategy for good in
the iteration game that is converted into a strategyeither for I or for IIin the game
G(n+1) (C).
There are thus two factors that go into the determinacy proof. First is a construction
which reduces G(n+1) (C) to the iteration game W Gn (M), and this is where the n
Woodin cardinals of M are used. Second is the appeal to a strategy for good in
W Gn (M), and this is where the n-iterability of M comes in.
The iterability is derived from the results of MartinSteel [19]. It is shown there
that if M embeds into a rank initial segment of V then it is n-iterable for every n,
and in fact more. (In particular then, the existence of a sharp for n Woodin cardinals
in V is certainly enough to secure ()n .) The main tool for proving iterability in
MartinSteel [19] is quoted as Theorem 12 in Appendix A, and the appendix draws
two specific consequences of this theorem: weak iterability and mild iterability (both
defined precisely in the appendix) for elementary substructures of rank initial segments
of V. The appendix also quotes a theorem of Neeman [31] which secures the existence
of a fully iterable model with a sharp for a Woodin limit of Woodin cardinals, provided
that in V there is a sharp for such a cardinal. These forms of iterability are sufficient for
all the determinacy proofs in this book, and the matter of proving iterability is therefore
not discussed any further.
It is the first factor mentioned above, the construction which allows reducing
G(n+1) (C) to W Gn (M), that is the starting point for this book. The book begins
by breaking this construction into its essential blocks. Subsequent developments re-
assemble the blocks in various ways. These developments again lead to reductions of
long games to iteration games, but apply to much stronger long games than G(n+1) (C).
The matter of breaking the construction of Neeman [29], [32] into basic blocks
is treated in Chapter 1, and a sample application there shows how these blocks can be
reassembled to prove 12 determinacy from ()1 (or more precisely the lightface version
of this).
Chapter 2 begins the exploration of longer games. It defines games of fixed countable
length, and proves the determinacy of these games from weakly iterable models with
an appropriate number of Woodin cardinals. Games of fixed countable length are direct
generalizations of the games Gn (C) (n < ) of Diagram 5, allowing any fixed
countable number of repetitions instead of the finitely many repetitions in Gn (C).
12 Introduction

The results of Chapter 2 are thus direct generalizations of projective determinacy. The
additional complexity which separates these results from projective determinacy is the
matter of propagation through limits. Every determinacy proof for long games involves
propagating a construction of a run of the game. The planning for this propagation
starts from the end really, since the construction has a definite goal, namely the payoff
set. In the case of Gn (C) (n < ) every stage of the construction except the initial
one has an immediate predecessor, and one can easily plan from the end backwards.
But in the general case of fixed countable length one has to somehow deal with limits.
The problem is solved in Chapter 2 by breaking the construction in a limit stage
into partsthis can be done naturally since the methods of Chapter 1 give rise to
constructions carried in roundsand spreading these parts over stages cofinal in .
Beyond games of fixed countable length, yet still below games of fixed length 1 ,
there lies a very rich hierarchy of games of variable countable length. These are long
games where the length of a run, while always countable, depends on the moves played
during that run. Chapter 3 deals with one particular class of games of this kind, the class
of games of continuously coded length. The notion of continuously coded length was
first introduced by Steel [41]. Roughly speaking his games can be described as follows:
Players I and II alternate playing natural numbers to produce reals y = $y (i) | i < %.
The reals are produced in the manner of Diagram 5, except that there is no set limit
on the number of repetitions. Instead, player I is asked to play a natural number n at
the start of the -th repetition, with the proviso that she is not allowed to play the same
number twice. Sooner or later she must run out of natural numbers, and at that point
the game ends. Since there are just countably many natural numbers I can play, the end
of the game must arrive at a countable stage. But player I can stretch the game to make
it as long as she wishes below 1 . Determinacy for these games is therefore stronger
than determinacy for games of fixed length for every < 1 , yet weaker than
determinacy for games of length 1 .
Our own games in Chapter 3 are somewhat stronger than those of Steel described
above. Instead of having n played at the start of the -th repetition, we fix some
function f : R and set n = f (y ) for each , so that n depends uniformly on
the moves made during the -th repetition. Chapter 3 proves determinacy for games
of this kind. Remember that a determinacy proof for long games always must involve
advance planning for propagation through limits. The planning in Chapter 3 is more
difficult than that in Chapter 2, since it is not known from the start how many limits to
plan for. Roughly speaking this problem is solved by recycling: taking moves that are
part of the current limit stage, and pushing them upward using some extender which
overlaps a Woodin cardinal, so that they can be used again in a future limit stage.
Without going into the details of the construction it is worthwhile looking at the kind
of iterations involved in such a process. A typical iteration for Chapter 3 is illustrated
in Diagram 6. j, +1 in the diagram is the ultrapower embedding by the extender F ,
which is taken from Q and applied to M . Notice that the iteration in Diagram 6 is
similar to that in Diagram 4, except for the addition of the extenders F . It is these
extenders which do the pushing mentioned above, shifting the moves of stage upward,
so that they can be used again in later stages.
Introduction 13

j0,1 j1,2

kkk # kkk #
Sk
Sk Sk
Sk
F0 F1
k SSSS k / Q0 k SSSS k / Q1 /
 
M0 M1 M2
0 1
T0 T1

j, +1

kkk #
Sk
kk F
/ M S SSSS k / Q 
M +1 /

T

Diagram 6. Illustration of the iterations used in Chapter 3.

This idea of pushing moves upward using extenders which overlap Woodin cardinals
turns out to be the seed for a method of advance planning that allows for proofs of
substantially stronger determinacy than that in Chapter 3. The method makes runs
of long games generic for Woodins extender algebra. It uses a particular format of
the extender algebra, see Section 4B, involving only extenders which overlap Woodin
cardinals. The advance planning for the determinacy construction is made to fit with the
extender algebra, in such a way that the use of extenders overlapping Woodin cardinals in
the determinacy construction doubles as a use of these extenders in a genericity iteration.
A typical stage of the resulting iteration is illustrated in Diagram 7. It is similar to stage
in Diagram 6, except that F is now applied to M , where may be smaller than . This
is in line with the non-linear dictates of Woodins genericity iteration, where extenders
are always applied to the earliest possible model. F continues also to serve for pushing
moves upward so that they can be used in later stages. In the settings of Diagram 7 it
takes moves from stage of the construction, and shifts them upward for recycled use.

j, +1

kkk / '
/ M _ _ _ _ _ _ _ M SSkSSS
kk Q  F
M +1 /
S k
T

Diagram 7. Stage , with non-linearity.

The precise definitions involved in the fitting of the advance planning for deter-
minacy proofs with Woodins extender algebra are given in Chapter 4. The relevant
constructions are then presented in Chapters 5, 6, and 7. It is in Chapter 7 that iterations
of the kind mentioned in the previous paragraph, and illustrated partially in Diagram 7,
show up. To facilitate the transition from the linear construction in Chapter 3 to the
non-linear construction in Chapter 7, the linear strands of the latter construction are iso-
14 Introduction

lated and dealt with in Chapter 6. Typical stages on such a linear strand have the format
illustrated in Diagram 8, and at least in one part (Section 6C) this makes the construction
of Chapter 6 similar to the constructions of Chapter 3 (illustrated in Diagram 6).

j, +1

kkk '
/ M kSk
SkSSSS k / Q _ _ _ _ _ _ _ Q  F
M +1 /

T

Diagram 8. From the perspective of stage .

There is, alas, much more to ponder in the development of Chapters 4 through 7
than the structure of the relevant iterations, discussed above. The bulk of the work by
far goes into the definitions and constructions which reduce long games to iteration
games. It is too early to say anything on this now, except for mentioning that it is there
that the large cardinals of the given model are used.
The method of fitting advance planning for determinacy constructions with Woodins
extender algebra, developed in Chapters 4 through 7, is really the pinnacle of this book.
It brings determinacy proofs to the level of long games which run up to a locally
uncountable ordinal. These are games where players I and II produce reals y in the
standard manner, continuing through transfinitely many repetitions until they reach the
first which cannot be collapsed to by any function which is definable, in some given
level of complexity, from the play. (The word locally in the terminology refers to the
fact that is only required to be uncountable with respect to some collection of definable
functions. Typically is still smaller than 1V .) In the first determinacy result we give
in Chapter 7, the restriction of definability is to functions which are constructible from
the play. Later we formulate a more general restriction. Finally we discuss a couple of
applications of our work, due to Woodin. One, which we only mention briefly, leads
to a connection between determinacy and completeness in !-logic. The other leads to
the consistency of determinacy for all ordinal definable games of length 1 .
It turns out that the tools developed in Chapters 4 through 7 can be used further
(see Neeman [33]) to prove determinacy for games which outright run to 1V . But this,
already, is another story.
Chapter 1
Basic components

Fix throughout this chapter a ZFC model M, a Woodin cardinal of M, and some
set X M&. We work with elements of the space X . Our goal is to develop
some Lipschitz continuous method for witnessing that a sequence x X has a given
property. Continuity is of the essence here. Ultimately we shall want to use this
method in proofs of determinacy, where x is presented by installments as the game
proceeds.
Let us be more precise. Fix some g which is col(, )-generic/M. We work in
M and in the generic extension M[g]. Fix A M, a col(, ) name for a subset of
(M&) X .
We intend to associate to A a certain family of auxiliary games AX, [x], defined
for each x X . Since X and are fixed throughout the chapter, we omit the subscript
and refer to the games as A[x]. By a game we oftenand in particular heremean a
game tree, that is a set of rules for valid moves, without any payoff at the end. A[x]
will be a length game with moves taken from M&.
The map x  A[x] will be Lipschitz continuous; the rules for the first n rounds of
A[x] will only depend on xn. For s X < we will thus talk about A[s], a game of
lh(s) many rounds. We will have:

Property 1. Each A[s] belongs to M. Moreover, the map s  A[s] belongs to M.


We refer to this map as A.

We will define the concept of a generic run of A[x] and show that:

Property 2. If a (M&) is a generic run of A[x], then $


a , x%  A[g].

We will define further the concept of a pivot for x. A pivot will be a pair T , a where:

T is an iteration tree of length on M;

T has a distinguished branchcalled the even branchleading to a direct limit


T
model Meven = Meven T : M M
and a direct limit embedding jeven = jeven even ;
and

a is a run of jeven (A)[x].

(For the last item, jeven (A)[x] is simply the union of the games jeven (A)[xn] over
n < .) We refer to the branches of T other than the even branch as odd branches.
Given an odd branch b we use Mb = MbT to denote the direct limit along b, and
jb = jbT : M Mb to denote the direct limit embedding. Most importantly a pivot
for x will satisfy:
16 1 Basic components

Property 3. For every odd branch b of T , there exists some h so that:

(1) h is col(, jb ())-generic/Mb ; and

(2) $
a , x% jb (A)[h].

Condition (2) of Property 3 should be contrasted with Property 2. The connection


between the two properties comes through the game A[x]. With the aid of a strategy
for II in this game we will be able to construct a generic run; with the aid of a strategy
for I we will be able to construct a pivot. It is thus fair to characterize A[x] as a
game where I tries to witness that $ a , x% jb (A)[h] for some shifts jb (A)[h] of A;
while II tries to witness that $
a , x%  A[g]. Intuitively, A = (s  A[s]) serves as
a Lipschitz continuous method for witnessing a particular property of x X . The
property addressed by a run a of A[x] is membership in the a -section of A[g] and its
shifts.

1A The auxiliary games map


We work here to define the map A associated to the name A and obtain Property 1.
Our definition of the games A[x] is an adaptation of definitions from Neeman [32],
which builds on MartinSteel [18]. We start with the basic definitions of types. These
definitions follow the exposition of [32, Section 3], with some cosmetic changes. The
changes are pointed out in Remark 1A.4. Then in Section 1A (2) we describe the rules
of the game A[x]. In that part there is a more serious change from Neeman [32], since
A in [32] only involved X , while here it involves (M&) X .

1A (1) Types. The definitions here form the rudiments for the later definition of the
auxiliary games A[x]. We phrase the definitions working in V, with a distinguished
ordinal . Later on we shall apply them in M with the ordinal , and in various other
models of ZFC . We work here, as we do throughout the book, in the language L of
Appendix A. By a formula we mean a formula of L .

Definition 1A.1. u is called a (, n)-type, where is a limit ordinal and n < , if u


is a set of formulae involving n free variables v0 . . . vn1 , a constant
, and additional
constants c for each c V {}.

Since is a limit ordinal, finite tuples of elements of V belong to V . It follows


that (, n)-types can be coded as subsets of V . We may therefore think of (, n)-types
as (coded by) elements of V+1 .

Definition 1A.2.

If u is a (, n)-type, we call the domain of u (denoted dom(u)).


1A The auxiliary games map 17

For , and m n, we let

projm c0 , . . . ,
(u) = { (, ck , v0 , . . . , vm1 ) | k N, c0 , . . . , ck V { },
( , ck , v0 , . . . , vn1 ) u, and makes no men-
c0 , . . . ,
tion of vm , . . . , vn1 }.

We use proj (u) to denote projn (u), and projm (u) to denote projm
(u).

Definition 1A.3. We say that the (, n)-type u is realized (relative to ) by x0 , . . . , xn1


in V just in case that
x0 , . . . , xn1 and are elements of V ; and
for any c0 , . . . , ck V {} and any formula ( , c0 , . . . , ck , v0 , . . . , vn1 ),

(, ck , v0 , . . . , vn1 ) u V |= [, c0 , . . . , ck , x0 , . . . , xn1 ].
c0 , . . . ,
(Implicitly we must have > and > .)
We call u the -type of x0 , . . . , xn1 in V (relative to ) if u is the unique (, n)-type
which is realized by x0 , . . . , xn1 in V .
We work relative to throughout this subsection, whether this is mentioned explic-
itly or not. We say that a (, n)-type u is realizable if it is realized by some x0 , . . . , xn1
in some V . Note that if u is realized by x0 , . . . , xn1 in V , then projm
(u) is realized
by x0 , . . . , xm1 in V .
Remark 1A.4. Our definition of types here differs from that of [32] in two ways. First,
we allow constants from V {} in -types, rather than just constants from V . This
will save us from passing to + -types later on. Secondly, we build the parameter
into the definition through the constant
.
Definition 1A.5. If the formula there exists a largest ordinal, and the formula ,
,
v0 , . . . , vn1 V , where is the largest ordinal are both elements of the (, n)-type
u we define

u = { (
, ck , v0 , . . . , vn1 ) | k N, c0 , . . . , ck V {}, and
c0 , . . . ,
the formula V |= [ , ck , v0 , . . . , vn1 ] where
c0 , . . . ,
is the largest ordinal is an element of u}.

If , , x0 , . . . , xn1 V and u is realized by x0 , . . . , xn1 in V+1 then u is


defined and is realized by the same x0 , . . . , xn1 in V .
Definition 1A.6. Let u be a (, n)-type, and let w be a (, m)-type. We say that w is a
subtype of u (and write w < u) if
< ;
m n; and
18 1 Basic components

the formula there is an ordinal and vn , . . . , vm1 V such that w


is realized
by v0 , . . . , vm1 in V is an element of the type u.
Note that w V , since < . Thus the formula listed in the last item above can
literally be an element of the type u. The definition of a subtype makes no mention of
realizability but only stipulates that one particular formula belongs to u. It is immediate
then that the property w < u is absolute for any two models of Set Theory which have
w and u as elements.
Definition 1A.7. We say that a (, m)-type w exceeds the (, n)-type u, if
> ;
m n; and
there exist ordinals , , and x0 , . . . , xm1 V such that
w is realized by x0 , . . . , xm1 in V ,
u is realized by x0 , . . . , xn1 in V , and
+ 1 < .
, , x0 , . . . , xm1 are said to witness that w exceeds u.
The definition of an exceeding type does refer to realizability. Note the essential
difference between w < u and w exceeds u for a type u which is realized by
x0 , . . . , xn1 in V . In both cases w must be a type with at least n variables which is
realized at a rank below . If w is a subtype of u then dom(w) < dom(u), while w can
exceed u only if dom(w) > dom(u).
Definition 1A.8. Let < be ordinals, E a -strong extender with crit(E) = and
u a type with dom(u) = . Let iE : V Ult(V, E) be the ultrapower embedding. We
define StretchE
(u) to be equal to proj (iE (u)).

iE (u) in Definition 1A.8 is a type in Ult(V, E) with domain iE (). iE () is at least


as large as by Fact 2 in Appendix A, since E is -strong. So proj (iE (u)) in the
definition makes sense.
Lemma 1A.9. Let M and N be models of ZFC . Suppose that M& + 1 = N & + 1.
Let u be a type in M with dom(u) = . Let E M and assume that

M |= E is a -strong extender with crit(E) = .

Then E can be applied to N and StretchE N


(u) (as computed in M) is equal to proj(iE (u)).

Proof sketch. The ultrapower embeddings iEN and iEM agree on subsets of M&, and u
is a subset of M&. #
Definition 1A.10. A (, n)-type u is called elastic just in case that u is defined and u
contains the following two formulae:
1A The auxiliary games map 19


is an inaccessible cardinal.

Let be the largest ordinal. Then for all < there exists a -strong extender
E V& such that crit(E) =
and StretchE
(u ) is realized (relative to
) by
v0 , . . . , vn1 in V .

The definition of elasticity, like the definition of the notion of a subtype, simply
requires certain formulae to belong to u. Elasticity is therefore absolute between any
two models of Set Theory which contain u. The main formula in Definition 1A.10
could not, formally, be an element of any type, since it makes a reference to u which is
not an allowed constant. However u is clearly definable (uniformly over all realizable
types). Strictly speaking we should replace u by its definition, namely the set of
formulae with constants in V { } {
} which are satisfied by v0 , . . . , vn1 in V ,
where is the largest ordinal.
Ordinarily if u is realized by x0 , . . . , xn1 in V+1 then StretchE
(u ) is realized
Ult(V,E)
by iE (x0 ), . . . , iE (xn1 ) in ViE () , and relative to iE (). The requirement that it
must also be realized by x0 , . . . , xn1 in V , and relative to , gives strong additional
information about E. In the terminology of [18], the existence of an elastic type which
is realized by x0 , . . . , xn1 in V+1 and has domain implies that is reflecting
in the parameters x0 , . . . , xn1 relative to . Thus the existence of realizable elastic
types is stronger than the mere existence of extenders. Indeed, to obtain many elastic
realizable types we need a Woodin cardinal.

Definition 1A.11. Let be a cardinal.

Let H V . Let < . An extender E is -strong wrt H if it is -strong and if


in addition iE (H ) V = H V .

A cardinal < is -strong wrt H if it is the critical point of an extender which


is -strong wrt H .

is <-strong wrt H if it is -strong wrt H for each < .

is a Woodin cardinal if for every H V there exists < which is <-strong


wrt H .

Lemma 1A.12 (MartinSteel, see [18], [30]). Assume that is a Woodin cardinal.
Let > be an ordinal, and let x0 , . . . , xn1 be elements of V . Then there exist
unboundedly many < such that the -type (relative to ) of x0 , . . . , xn1 in V+1
is elastic.

Proof sketch. Apply Definition 1A.11 to H = $H | < % where H is the -type of


x0 , . . . , xn1 in V . #

Finally we state the One Step Lemma, which essentially says that if w exceeds u
and u is elastic, then one can stretch u to obtain a supertype of w.
20 1 Basic components

Lemma 1A.13 (One Step Lemma, MartinSteel [18]). Assume that u is an elastic
type, and that w exceeds u. Let = dom(w) and = dom(u). Assume < .
Then there exists a + -strong extender E V& with critical point and such that
w < StretchE + (u).

Proof. Let , , x0 , . . . , xm1 witness that w exceeds u. Note, since u exists, is a


successor ordinal. Say = + 1. Pick E V& so that StretchE
+ (u ) is realized
by x0 , . . . , xn1 in V , and relative to . (This is possible since u is elastic.) Then
w is a subtype of StretchE
+ (u ), as it is realized by x0 , . . . , xn1 , xn , . . . , xm in V
and < . Simple properties of realizable types now imply that w is a subtype of
StretchE + (u). #

Remark 1A.14. Given < we can add the clause E is -strong to the conclusion of
Lemma 1A.13; simply replace StretchE E
+ (u ) in the proof by Stretchmax{ +,} (u ).

1A (2) The rules of the game. Recall that is a Woodin cardinal of M, X is an element
of M&, and A is a name for a subset of (M&) X in M col(,) . Fix x X . Our
immediate goal is to define the auxiliary game AX, [x] = A[x] and obtain Property 1.
Moves in A[x] will include types taken from M&. We shall need some local
indiscernible above as a starting point for the realizations of our types. We make the
following definition:

Definition 1A.15. Let L < H be ordinals greater than . We say that $L , H % is a pair
of local indiscernibles of M relative to just in case that:

(M&L + ) |= [L , c0 , . . . , ck1 ] (M&H + ) |= [H , c0 , . . . , ck1 ]

for any k < , any formula with k+1 free variables, and any c0 , . . . , ck1 M&+.

Local indiscernibles are thus indiscernible in a limited way for formulae with pa-
rameters in M& + . Exactly why we need a local indiscernible in the definition of
A[x] will only become clear in Section 1C. For the time being let us note that there are
local indiscernibles in M. In fact one neednt go very far above to encounter them,
as the next claim demonstrates:

Claim 1A.16. Let jumpM () denote the successor in M of the cardinality in M of


M&( + + 1). Then there are L < H which are indiscernible in the sense of
Definition 1A.15, and are both smaller than jumpM ().

Proof. This is a simple cardinality argument. For each < jumpM () let T be the set
of tuples $k, , c0 , . . . , ck1 % so that k is a natural number, is a formula with k +1 free
variables, c0 , . . . , ck1 belong to M& + , and (M& + ) |= [, c0 , . . . , ck1 ]. T
is a subset of M& +. So there can only be cardM (M& + +1) many possible values
for T . It follows that there exist some L < H below jumpM () so that TL = TH .
$L , H % is then a pair of local indiscernibles of M relative to . #
1A The auxiliary games map 21

Let $, % be the least (lexicographically, minimizing first over ) pair of local


indiscernibles of M relative to . will appear in our definition of A[x], specifically
in rule (2d). will be used in Section 1C.
Define the game A[x] to be played according to Diagram 1.1 and rules (1)(4)
below. We refer to a run of the game as a = $a0 , a1 , . . . % where a0 = $l0 , p0 , u0 , w0 %,
a1 = $l1 , p1 , u1 , w1 %, etc. It is helpful to think of A[x] as a game where I tries to
witness that there exists some h so that $ a , x% belongs A[h]. II tries to keep I honest
by playing dense sets which I must meet. In addition II plays rank functions on Is
attempts to create h, in effect trying to witness that no such h-s exist.

I l0 , p0 , u0 l1 , p1 , u1 l2 , p2 , u2
II w0 w1 w2 ...

Diagram 1.1. The game A[x].

The moves in A[x] take place inside M, and rules (1)(4) should be read relativized
to M. In particular all the types mentioned are elements of M. Realizations are in M
and relative to .

(1) (Rule for I) ln is a natural number smaller than n, or ln = new.1

(2) (Rule for I) If ln = new then:

(a) un is a (n , 4)-type for some n < . If n > 0 we require n > n1 . If


n = 0 we require n > rank(X).
(b) proj3 (un ) is elastic.
(c) pn M&n is a condition in col(, ).
(d) There exist some names a and x, both members of M& + 1, so that un is
realized by A, a , x, in M& + 2.

When I plays ln = new she is indicating that she wishes to start a fresh attempt at
constructing h. This fresh attempt consists of some condition pn , the first in a decreasing
chain which should form the generic h, and names a , x which will later on be forced to
belong to A, and to produce exactly a and x. Instead of directly playing the names, I
merely has to play their type. Note that in a sense this is easier for I; a single type may
be realized by many different names, and I is not asked to pick among them.

(3) (Rule for II)

(a) wn is a (n , 5)-type for some n . If n > 0 we require further n > n1 .


(b) wn is a subtype of proj3 (un ). (Note this implies n < n .)

Furthermore, wn must contain the following formulae:


1 We think of new as coded by some element of V .

22 1 Basic components

(c) v3 is a dense set of conditions in the forcing col(,


),
(d) v4 is an ordinal, and
(e) v4 + 4 is the largest ordinal.
Thus, if un were realized by A, a , x, in M& + 2, then wn would be realized by
A, a , x, D, in M& + 5 for some < (in fact + 5 < + 2) and some D which
is dense in col(, ). Note the two features of IIs response to un . First, II gives some
dense set D. Secondly, II produces an ordinal , smaller than the one had before.
As with Is moves, rather than asking II to play the dense set and the ordinal directly,
we merely ask II to play their type. This gives II some extra freedom, since the type
specifies a set of D-s and -s (the set of realizations), rather than a single D and a
single .
(4) (Rule for I) If ln N, i.e., ln  = new then:
(a) un is a (n , 5)-type for some n < . We require further n > n1 .
(b) proj3 (un ) is elastic.
(c) pn M&n is a condition in col(, ), and extends pln .
(d) un exceeds wln .
Furthermore, un must contain the following formulae:
pn v3 ,
(e)
pn col(, ) $v1 , v2 % v0 ,
(f)
n col(, ) v1 (i) =
(g) For each i < ln , p ai ,
n col(, ) v2 (i) =
(h) For each i < ln , p xi , and
(i) v4 is an ordinal and v4 + 1 is the largest ordinal.
When playing ln N (rather than ln = new) I is indicating that in round n she wishes
to continue the h she left off in round ln (note ln < n always). To do this I plays a
condition pn which extends pln . To be fair we demand that this condition meets the
dense set handed by II in round ln this is the requirement (4e). To connect Is moves
with x and a we demand some agreement between these objects and the names played
by Irule (4g) demands that the first ln elements of the sequence named by v1 are
in fact a0 , . . . , aln 1 , and rule (4h) demands that the first ln elements of the sequence
named by v2 are in fact x0 , . . . , xln 1 . Rule (4f) says that $v1 , v2 % are forced to produce
an element of v0 . v0 in some sense stands for A; see rule (2d) above. So I is slowly
constructing h, relating the pair named by $v1 , v2 % to $ a , x%, and forcing this pair to
belong to A[h].
As always, we ask the player to just play the types of the objects in questions. Note
that the dense set given by II in round ln was not given with precision. II only played
the ln -type of the set. In round n we ask I to play un which exceeds the type played by
II in round ln . Thus, in a sense, I gets to pick one of the dense sets handed by II, and
meet that one.
1B Generic runs 23

Remark 1A.17. With respect to rules (4g) and (4h) we note that for i < ln both ai and
xi belong to M&n , and so
ai and
xi are valid constants for the type un . The fact that ai
belongs to M&n traces back to rules (4a) and (2a) which certainly imply that n > i .
The same rules imply that n > rank(X), so that xi too belongs to M&n .
The rules above complete the description of the game A[x]. Observe that x only
appears in rule (4h), and the only part of x relevant in round n is xln . Since ln < n,
certainly xn suffices. Our description of A[x] therefore defines a map x  A[x],
Lipschitz continuous in the sense that the rules for the first n rounds of A[x] depend
only on xn. For s X< our description gives a game A[s] of lh(s) many rounds
(the first lh(s) rounds of A[x] for any x which extends s), and it is clear that the map
A = (s  A[s]) belongs to M. We call this map the auxiliary games map associated
to A, , and X.

1B Generic runs
We work with some g which is col(, ) generic over M. Our goal is twofold. Given
x X we wish to define the concept of a generic run of A[x] and verify Property 2.
In addition we wish to demonstrate how generic runs can be constructed, with the aid
of some strategy for II in A[x].
We shall have to convert g to a generic enumeration of M&. Fix, in M, some
bijection : M&. We regard g as a surjection g : , and will talk about
the composed surjection g : M&.
Fix x X , not necessarily in M. Let a be a run of A[x], again not necessarily
in M. Say a = $a0 , a1 , . . . % where an = $ln , pn , un , wn %.
We say that e < is valid at n if g(e) is a legal move for I in A[x] following
a0 , . . . , an1 . (In particular g(e) has the form $l, p, u% for some l N {new},
some condition p, and some type u.)
Definition 1B.1. a is a generic run of A[x] (wrt , g) if for each n < , $ln , pn , un %
is exactly equal to g(e) for the least e which is valid at n.
Thus, a run a is generic if Is moves in the run are guided by a generic enumeration;
I simply plays the first legal move in each round. Note that the definition places no
restriction on IIs moves in a . The only restrictions placed are on I.
Lemma 1B.2. Suppose that a is a generic run of A[x]. Then $
a , x%  A[g]. (Note,
this is only useful if x and a belong to M[g].)
Proof. Assume for contradiction that $ a , x% A[g]. In particular a and x both belong
to M[g]. Fix names a and x in M& + 1 so that a [g] = a and x[g] = x. (We can find
such names in M& + 1 since both x and a are -sequences from M&.)
Observe that a , x, and g can meet any challenges posed by II in the game A[x]; the
pair $a , x% is forced by g to belong to A, a is forced by g to equal a , and x is forced by
24 1 Basic components

g to equal x. Our plan is to show that among her many attempts in the generic run of
A, player I also tried a , x, g. But the attempt a , x, g must actually succeed, and this is
a contradiction since II in addition to dense sets, plays rank functions to make sure
all of Is attempts fail.
To be more precise, we intend to produce sequences n0 , n1 , n2 , . . . and 0 , 1 , . . .
so that:
(1) ln0 = new and lni = ni1 for i > 0.
(2) For each i, pni is a condition in g.

(3) For i = 0, uni is realized by A, a , x, in M& +2. (Recall that > is the lower
ordinal in the least pair of local indiscernibles of M relative to ; see Section 1A.)
(4) For each i > 0, uni is realized by A, a , x, D, i in M&i + 2 for some dense
set D.
(5) 0 = and i < i1 for all i > 0.
Condition (1) simply says that n0 , n1 , . . . form a branch in Is attempts in the run
a of the game A[x]. Conditions (2)(4) relate this branch to a , x, gour guaranteed
attempt which cannot fail. Note the appearance of the ordinals i in condition (4).
These correspond to the ordinals mentioned after rule (3) in Section 1A (2), the
ordinals given by the variable v4 . These ordinals will decrease steadily, giving us the
desired contradiction.
Claim 1B.3. There exists n so that
ln = new,
pn is a condition in g, and
un is realized by A, a , x, in M& + 2.
Proof. Suppose not. The claim can be phrased inside M[g]. If it fails, this must be
forced by some condition in g. So fix q g which forces the claim to fail. Strengthening
q if needed, we may assume that q also forces a is a generic run of A[x] wrt , g.
Let p be your favorite condition in g. Strengthening q further we may assume
that q extends p. Let j < be the domain of q. Let < be large enough that
q(0), . . . , q(j 1), and p all belong to M&. Using Lemma 1A.12, fix
> max{, rank(X)}, below , so that the (, 3)-type of A, a , x in M& + 2 is elastic.
Let u be the (, 4)-type of the same parameters plus (again in M& + 2). Observe
that $new, p, u% is a legal move for I in round n of A[y], for any y X , so long as
n1 < . (This follows directly from rule (2) in Section 1A (2). Note particularly the
fact that proj3 (u) is elastic, by choice of .)
Let q be the condition q extended by j  1 $new, p, u%. We will show that

q forces the claim to hold, contradicting our initial choice of q g which forces the
claim to fail.
1B Generic runs 25

Fix some g , a generic which contains q . Let x = x[g ] and let a = a [g ].


Say a = $a0 , a1 , . . . %, and ai = $li , pi , ui , wi %. Our goal is to show that the claim
holds for these objects. Certainly it is enough to find n < so that $ln , pn , un % =
$new, p, u%.
Now g contains q, so we know that a is a generic run of A[x ] wrt , g . Thus, for
each n < , $ln , pn , un % is the least legal move, wrt to the enumeration g . To be more
precise: For each n < let en < be the first number so that $ln , pn , un % = g (en ).
We know that en is the least number valid (wrt g ) at n.
Let n be least so that en j . It is enough to show that en = j . Then $ln , pn , un % =
g (j ) = $new, p, u% as required. (Note q (j ) = $new, p, u% by our
definition of q .)
To show that en = j it is enough to check that g (j ) is a legal move following
a n. Then j is valid at n, and by the minimality condition in Definition 1B.1, en > j

is ruled out.
Remember that $new, p, u% is a legal move in round n of A[y], for any y, so long
as n1 < . So it is enough to check that u0 , u1 , . . . , un1 all belong to M&. Since
, p , u % = g (e ), certainly it is enough to check that g (e ) belongs to
$lm m m m m
M& for all m < n. Now n is least such that en j . So m < n implies that em < j,

and g (em ) then equals q(em ). We chose to begin with which is bigger than
the ranks of q(0), . . . , q(j 1). So q(em ) M&, as required. #

Claim 1B.4. Suppose m and are such that

pm is a condition in g, and

um is realized by A, a , x, in M& + 2.

Then there exists an n > m, a dense set D, and an ordinal so that

ln = m,

pn is a condition in g,

un is realized by A, a , x, D, in M& + 2, and

< .

Proof. Suppose not. The claim can be phrased inside M[g]. If it fails, this must be
forced by some condition. Fix q g which forces the claim to fail. Strengthening q as
needed we may assume that q forces a is a generic run of A[x] wrt , g.
By the rules of A[x], specifically rule (3), wm is a subtype of proj3 (um ). Since um
is realized by A, a , x, in M& + 2, it follows that there is an ordinal and a set D
so that wm is realized by A, a , x, D, in M& + 5, and + 5 < + 2. In particular
< as required for our current claim.
Find a condition p g D, extending pm , which forces the following:

(i) $a , x% A,
26 1 Basic components

(ii) for each i m, a (i) = ai , and


(iii) for each i < m, x(i) = xi .
This is possible since $a [g], x[g]% does belong to A[g] (remember our initial assumption
for contradiction at the start of the proof of Lemma 1B.2: $ a , x% A[g]), a [g](i) does
equal ai for each i < m, and x[g](i) does equal xi for each i < m.
By strengthening q again, we may assume that q extends p. Let j be the domain
of q. Let < be large enough that q(0), . . . , q(j 1), p, and a0 , . . . , am all
belong to M&. Using Lemma 1A.12, find > , below , so that the (, 3)-type of
A, a , x in M& +2 is elastic. Let u be the (, 5)-type of the same parameters plus D, .
(Precisely, u is the (, 5)-type of A, a , x, D, in M& + 2.) Observe that u exceeds
wm . This follows directly from our choice of u, the realization of wm in M& + 5
indicated above, and Definition 1A.7. (Note , and hence , is greater than m since
am belongs to M&.)
Observe further that $m, p, u% is a legal move in A[y] for any y X which extends
xm, and following any position b0 , . . . , bk in A[y] which extends a m + 1, as long as
the types played in b0 , . . . , bk have domains below . This follows directly from the
rules of A[y], specifically rule (4). (For rule (4c) note that p was chosen extending pm .
For rule (4e) remember that p was taken from g D. Rules (4f)(4h) hold because of
conditions (i)(iii) above. In the case of rule (4g) we use also the fact that bi = ai for
i < m.)
Let q be the condition q extended by j  1 $m, p, u%. An argument similar
to the one used in Claim 1B.3 now shows that q forces $m, p, u% to be played in any
generic run which extends a m + 1. By condition (ii), p and hence q force a to extend
a m + 1. Hence q forces Claim 1B.4 to hold. #
Claim 1B.5. Suppose m, C, and are such that
pm is a condition in g, and
um is realized by A, a , x, C, in M& + 2.
Then there exists an n > m, a dense set D, and an ordinal so that
ln = m,
pn is a condition in g,
un is realized by A, a , x, D, in M& + 2, and
< .
Proof. This is similar to Claim 1B.4. We start with um realized by A, a , x, C, in
M& + 2. This means that proj3 (um ) is realized by A, a , x in M& + 2. Since wm is
a subtype of proj3 (um ) it follows that there is an ordinal and a set D so that wm is
realized by A, a , x, D, in M& + 2, and + 5 < + 2. The rest of the proof is
identical to that of Claim 1B.4. #
1C Pivots 27

Equipped with Claims 1B.31B.5 we can complete the proof of Lemma 1B.2. Our
goal is to construct a sequence ni , i , i < , satisfying conditions (1)(5) on page 24.
Let n0 be some n witnessing Claim 1B.3, and let 0 = . Let n1 , D1 , and 1 witness
Claim 1B.4 applied with m = n0 and = 0 . Working inductively, let ni+1 , Di+1 ,
and i+1 for i 1 witness Claim 1B.5 applied with m = ni , C = Di , and = i .
It is easy to verify that conditions (1)(5) hold for these objects. In particular
$i | i < % forms a decreasing sequence of ordinals, giving the desired contradiction
and completing the proof of Lemma 1B.2. #
We have so far defined generic runs and obtained Property 2. Let us end by noting
that, with the crucial help of a strategy for II in A[x], one can easily construct a generic
run. The construction is simply a matter of ascribing the first legal move for player I in
each round. The following claim guarantees the existence of legal moves.
Claim 1B.6. Suppose a0 , . . . , an1 is a position in A[x]. Then there is a move $l, p, u%
which is legal for I following a0 , . . . , an1 .
Proof. Take your favorite names a and x in M& + 1. Using Lemma 1A.12 find < ,
large enough that a0 , . . . , an1 M&, larger than rank(X), and so that the -type
of A, a , x in M& + 2 is elastic. Let u be the (, 4)-type of the same objects plus ,
again in M& + 2. It is easy to see that $new, , u% is a legal move for I following
a0 , . . . , an1 . #
Lemma 1B.7 ( : M&, g col(, )-generic/M). For every x X there
exists a strategy gen [x] for I in A[x] so that every run according to gen [x] is generic.
Moreover, the association x  gen [x] is Lipschitz continuous, induced by some map
gen = (s  gen [s]). The map gen belongs to M[g].
Proof. gen [x] simply plays the first (in the enumeration given by g) legal move in
each round. Such a move exists by Lemma 1B.6. The dependence of gen [x] on x is
Lipschitz continuous because the notion of a legal move in round n of A[x] depends
only on xn. Finally, to define the map gen one needs the map A and the enumeration
g. Both exist in M[g]. #
Lemma 1B.7 states, in a precise way, that generic runs of A[x] can be constructed
with the aid of a strategy for II. Given some strategy for II in A[x] simply pit
against gen [x]. The result is a generic run.
We refer to the map gen of Lemma 1B.7 as the generic strategies map associated
to A, , and X. There is a dependence on g in the definition of gen . g will always be
clear from the context, so we suppress the dependence in our notation.

1C Pivots
A length iteration tree T is called nice if (2n) T (2n+2) for every n. 0 T 2 T 4 T
then forms a branch, called the even branch, through T . We use Mn = MnT to denote
28 1 Basic components

the models of T , and jn,m = jn,m T : M M for n T m to denote the embeddings


n m
of the tree. We use Meven = Meven T to denote the direct limit along the even branch of
T
T and use j2n,even = j2n,even to denote the direct limit embeddings. jeven stands for
j0,even . We refer to branches b other than the even branch as odd. We use Mb = MbT
T
to denote their direct limits, and jn,b = jn,b for n b to denote the direct limit maps.
jb stands for j0,b .

Definition 1C.1. A pivot for x X is a pair T , a , where

(1) T is a nice iteration tree on M;

(2) a is a run of jeven (A)[x]; and

(3) for every odd branch b of T there exists h so that

(a) h is col(, jb ())-generic/Mb , and


(b) $
a , x% jb (A)[h].

This definition is made with reference to particular M, , X, A, and A = AX, .


When there is danger of ambiguity we say A-pivot rather than just pivot.

Remark 1C.2. Condition (3) applies to all odd branches b, including ones where Mb
is illfounded.

In the previous section we saw that a generic run a of A witnesses that $ a , x% does
not belong to A[g]. A pivot is an attempt to witness the opposite, that $ a , x% does
belong, not to A[g] but at least to some shift of this set, namely jb (A)[h] for an odd
branch embedding jb and a generic h. Our goal in this section is to formulate a result
on pivots similar to Lemma 1B.7. We first phrase the construction of a pivot for x as a
game, Apiv [x]. We then describe a strategy piv [, x] which plays for II in this game.
Just as gen [x] plays for I in A[x] and always produces generic runs, piv [, x] will
play for II in Apiv [x] and produce pivots.
From the point of view of player I the new game Apiv [x] will be nothing more than
a shift of the original auxiliary game A[x]. A strategy for I in the original game
A[x] could thus be used in the new game, and pitted against piv [, x]. The resulting
run will form a pivot. The reader should contrast this with the discussion following
Lemma 1B.7, where a strategy for II was pitted against gen [x] to form a generic run.

1C (1) The game Apiv [x]. The game Apiv [x] is played according to Diagram 1.2 and
rules (1)(6) below.
Player II has the task of playing extenders defining an iteration tree T . In addition,
the players play the game A[x], but shifted along the even branch of T . I starts with
l0 , p0 , u0 , a legal move in A[x]. Then II plays E0 , E1 , which are used to determine
the first models M1 , M2 of T , and the embedding j0,2 . We then shift A[x] to M2 ,
and player II plays w0 , a legal move in j0,2 (A)[x] following j0,2 (l0 , p0 , u0 ). The
1C Pivots 29

I l0 , p0 , u0 l1 , p1 , u1 l2 , p2 , u2
II E0 , E1 , w0 E2 , E3 , w1 ...

Diagram 1.2. The game Apiv [x].

game continues in this way. Player I plays l1 , p1 , u1 which must be a legal move in
j0,2 (A)[x]. Player II replies in M4 , etc.
We list the exact rules, rules (1)(6) below, interspersed with some helpful ter-
minology. The terminology, and the rules, apply to the run of Apiv [x] displayed in
Diagram 1.2.
Let T be the unique tree order which satisfies:
the T -predecessor of 2n + 2 is 2n for each n < ,
if ln = new then the T -predecessor of 2n + 1 is 2n, and
if ln N then the T -predecessor of 2n + 1 is 2ln + 1.
Let T be the iteration tree on M determined by $En | n < % and the tree order T .
(1) (Rule for II) The extenders En must be played in a way which makes this definition
of T work. Precisely:
(a) If ln N then E2n M2n must be an extender with critical point within the
level of agreement between M2n and M2ln +1 . We set
M2n+1 = Ult(M2ln +1 , E2n ).

(b) E2n+1 M2n+1 must be an extender with critical point within the level of
agreement between M2n+1 and M2n . We set
M2n+2 = Ult(M2n , E2n+1 ).

(2) (Rule for II) If ln = new then E2n must equal pad so that M2n+1 = M2n and
j2n,2n+1 = id.
(3) (Rule for II) T must be normal, and must only use extenders taken from below .

(An iteration tree is normal if {StrengthMn (En )}n<, En =pad forms an increasing se-
quence. An iteration tree uses only extenders taken from below if for every n,
En Mn &j0,n ().)
For each n < let an = j2n,2n+2 (ln , pn , un ), wn .2 (Note that ln , pn , un are
shifted from the 2n-th model to the 2n + 2-nd model.) Let a = $a0 , a1 , . . . %. We intend
to split the following demand between the two players:
2 We treat sequences informally in this book, often dropping the enclosing brackets. (For otherwise brackets
of all kinds would quickly pile up, to the point of obscuring the objects involved.) To avoid confusion we
sometimes use the symbol to indicate a sequence, as opposed to a singleton element.
30 1 Basic components

a is a run of jeven (A)[x].

We do this with the following three rules:

(4) (Rule for I) ln , pn , un is a legal move for I in j0,2n (A)[x] following the position
$a0 , . . . , an1 %.

(5) (Rule for II) The critical point of j2n,2n+2 is larger than rank(X) and large enough
that none of a0 , . . . , an1 is moved by this embedding.

(6) (Rule for II) wn is a legal move for II in j0,2n+2 (A)[x] following the position
$a0 , . . . , an1 %, j2n,2n+2 (ln , pn , un ).

Let Pn denote $a0 , . . . , an1 %. Let Qn denote Pn , j2n,2n+2 (ln , pn , un ), which by


rule (5) is equal to j2n,2n+2 (Pn , ln , pn , un ). Using rules (4) and (6) one can verify
by induction on n that Pn is a position in the game j0,2n (A)[x]. The case of n = 0 is
clear. Suppose inductively that Pn is a position in the game j0,2n (A)[x]. By rule (4) so
is Pn , ln , pn , un . Applying j2n,2n+2 we see that Qn = j2n,2n+2 (Pn , ln , pn , un ) is
a position in j0,2n+2 (A)[x]. By rule (6) so is Qn , wn . But this is Pn+1 .
Thus we have indeed that a is a run of jeven (A)[x].
This completes the rules of Apiv [x]. Note that a run of Apiv [x] produces T and a
which satisfy conditions (1) and (2) of Definition 1C.1. To make T , a a pivot we must
further satisfy condition (3).
Observe further that the association x  Apiv [x] is Lipschitz continuous. To figure
out the rules for the first n rounds of Apiv [x] we need only know the rules for the first
n rounds of A[x] (so that we can shift those rules along the even branch), and for this
it is enough to know xn. In effect we described a map Apiv = (s  Apiv [s]) defined
on s X< . We refer to it as the pivot games map associated to A, , and X. It is clear
that the map Apiv belongs to M.

1C (2) Constructing piv [, x]. Fix x X , and fix some map  : M& + 1.
We work to define piv [, x], a strategy for II in Apiv [x]. (We use piv [, x] in our
notation to emphasize that our construction depends not only on x, but also on the
enumeration .) Assuming that  is onto we shall later on show that all runs according
to piv [, x] are pivots.
Historical Remark. Our construction here builds closely on the work of MartinSteel
[18]. In particular the iteration tree which stands at the heart of the construction is
similar to the one in MartinSteel [18]. For more on the connections between our
construction and the earlier construction of MartinSteel [18] see Neeman [32, p. 248].
We define piv [, x] by describing below the course of a run according to the strategy,
played against some imaginary opponent as I. We then check that the run described is
indeed a pivot.
In round n of the construction our imaginary opponent will play the move ln , pn , un .
It will be our task to play E2n , E2n+1 , wn . We use the notation established before:
1C Pivots 31

Our construction produces a nice iteration tree T , and a as already described in


Section 1C (1). Recall that Pn = $a0 , . . . , an1 % is a position in j0,2n (A)[x], and
Qn = $a0 , . . . , an1 %, j2n,2n+2 (ln , pn , un ) is a position in j0,2n+2 (A)[x]. Re-
call further that an = j2n,2n+2 (ln , pn , un ), wn . Of course ln is not moved by
j2n,2n+2 . We shall make sure that pn is not moved either. So in fact we shall have
an = $ln , pn , j2n,2n+2 (un ), wn %.
In addition to constructing the run of Apiv [x], we construct the following objects:

a n , xn M2n+1 &n + 1 where n = j0,2n+1 (), and

Dn M2n+1 &n + 1.

a n , xn , Dn will be constructed in round n, together with E2n , E2n+1 , wn . a n and xn


will be col(, n )-names in M2n+1 , and Dn will be a dense set.
We intend to maintain all the requirements set by the rules of Apiv [x]. Note that
this includes requirements from A[x] shifted to various models along the even branch.
In addition we intend to maintain:

(A) wn is elastic.

(B) wn is realized by An , a n , xn , Dn , and n in M2n+1 &n +5, where An = j0,2n+1 (A)


and n = j0,2n+1 ().

(C) If ln N then a n = j2ln +1,2n+1 (a ln ) and xn = j2ln +1,2n+1 (xln ).

(D) If ln N then pn belongs to j2ln +1,2n+1 (Dln ).

(E) 0 < 0 < 1 < 1 < , where n = dom(un ) and n = dom(wn ).

(F) M2n and M2n+1 agree to n + . M2n+1 and M2n+2 agree to n + .

With regard to condition (B) we point out that realizations in M2n+1 use n = j0,2n+1 ()
to interpret the constant . We remind the reader that > is the lower ordinal in the
least pair of local indiscernibles of M relative to . is the higher ordinal in that pair.
We shall use the indiscernibility properties of and , given by Definition 1A.15,
during the construction. With regard to conditions (C) and (D) let us remind the reader
that 2ln + 1 is the T -predecessor of 2n + 1 when ln N. So j2ln +1,2n+1 makes sense.
Remember that A is a name for a subset of (M&) X . The canonical name for
A therefore certainly belongs to M& + . Without loss of generality let us assume that
A itself belongs to M& + .
Let us now start round n of the construction. We assume inductively that conditions
(A)(F) hold for m < n. Our opponent opens round n by playing $ln , pn , un %, a legal
move for I in j0,2n (A)[x] following the position Pn .

Case 1. If ln = new. Set E2n = pad as required, so that M2n+1 = M2n , j2n,2n+1 =
id, and j0,2n+1 = j0,2n . We have An = j0,2n (A), n = j0,2n (), and n = j0,2n ().
The rules of j0,2n (A)[x], specifically rule (2) in Section 1A (2), tell us that there exist
32 1 Basic components

names a n and xn , both elements of M2n &n + 1 = M2n+1 &n + 1, so that un is realized
by An , a n , xn , n in M2n &n + 2 = M2n+1 &n + 2.
Let n = j0,2n+1 ( ). Using the indiscernibility of and given by Defini-
tion 1A.15 we see that un is realized by An , a n , xn , n in M2n+1 &n + 2. It follows
that

() proj3 (un ) is realized by An , a n , xn in M2n+1 &n + 2.

Pick Dn M2n+1 &n + 1, a dense set in col(, n ). The precise way we pick Dn
uses the enumeration , and will be explained later on. For the time being let us just
take Dn as given.
Using Lemma 1A.12 find n < n , greater than n = dom(un ), so that the n -type
of An , a n , xn , Dn , n in M2n+1 &n + 5 is elastic. Let wn be this type.
Observe that wn exceeds proj3 (un ) in the model M2n+1 . This follows directly from
Definition 1A.7, noting that n + 5 (the level of realization of wn ) is smaller than
n + 2 (the level of realization of proj3 (un ) given by ()). Applying the One Step
Lemma (Lemma 1A.13) inside M2n+1 find E2n+1 M2n+1 , a n + -strong extender
E 3
with critical point n , so that wn is a subtype of Stretchn2n+1 + (proj (un )). To satisfy the
normality clause in rule (3), make sure that E2n+1 is stronger than all previous extenders
on the tree. This can be done using Remark 1A.14. Note that the One Step Lemma
gives an E which belongs to M2n+1 &j0,2n+1 (). So the second clause in rule (3) is
satisfied.
Let M2n+2 = Ult(M2n , E2n+1 ) and let j2n,2n+2 be the ultrapower embedding. By
E 3 3
Lemma 1A.9, Stretchn2n+1 + (proj (un )) equals projn + (j2n,2n+2 (un )). Thus wn is a
3
subtype of projn + (j2n,2n+2 (un )). In particular this means that

wn < proj3 (j2n,2n+2 (un )).

Using this it is easy to check that wn is a legal move for II in j0,2n+2 (A)[x] following
the position Qn . This secures rule (6) of Apiv [x].
Note that the critical point of j2n,2n+2 is equal to n . For n > 0, n is greater
than j2n2,2n (n1 ). (This follows from rules (2a) and (4a) in Section 1A (2).) Thus,
for n > 0, the critical point of j2n,2n+2 is greater than j2n2,2n (n1 ). It follows that
an1 , and so certainly a0 , . . . , an2 , are not moved by j2n,2n+2 . Using the fact that
0 > rank(X) and condition (E) we see that the critical point of j2n,2n+2 is also higher
than rank(X). This secures rule (5) of Apiv [x].
The objects E2n , E2n+1 , wn defined above thus form a legal move for II in round n
of Apiv [x]. This completes the construction in case 1 of round n. Conditions (A)(F)
are easy to verify. # (Case 1)

Case 2. If ln N. We know by the rules of A[x], specifically rule (4d) in Section 1A (2),
that un exceeds wln . We know by condition (A) that wln is elastic. Applying the One
Step Lemma inside M2n we obtain E2n M2n &j0,2n (), a n + -strong extender
with critical point ln , so that un < StretchE 2n
n + (wln ). To satisfy rule (3), make sure
1C Pivots 33

that E2n is stronger than all previous extenders on the tree. This can be done using
Remark 1A.14.
Set M2n+1 = Ult(M2ln +1 , E2n ) and set j2ln +1,2n+1 to be the ultrapower embed-
dings. Note how this corresponds to the definition of T given in the previous subsection.
Note further that M2n and M2ln +1 agree to ln + , so this makes sense and complies
with rule (1a) in Section 1C(1). This agreement between M2n and M2ln +1 follows from
the inductive conditions (E) and (F) in our construction.
By Lemma 1A.9, StretchE 2n
n + (wln ) is equal to projn + (j2ln +1,2n+1 (wln )). So un
is a subtype of the latter. It follows that

un < j2ln +1,2n+1 (wln ).

Set An = j0,2n+1 (A) and n = j0,2n+1 (). Set a n = j2ln +1,2n+1 (a ln ) and xn =
j2ln +1,2n+1 (xln ) as required by condition (C). Let Cn = j2ln +1,2n+1 (Dln ). Using con-
dition (B) we see that j2ln +1,2n+1 (wln ) is realized by An , a n , xn , Cn , n in M2n+1 &n +5.
Since un is a subtype of j2ln +1,2n+1 (wln ) we conclude that un is realized by the same
objects, in M2n+1 &n + 2.
Before proceedings, lets recall that un contain the formula pn v3 ; see rule (4e)
in Section 1A (2). Our realization of un in M2n+1 has Cn standing for v3 . Thus
pn Cn = j2ln +1,2n+1 (Dln ). Our assignments so far therefore satisfy the inductive
condition (D) at n.
We now continue very much as we did in case 1. Switching between the local
indiscernibles n and n = j0,2n+1 ( ) we see that un is realized by An , a n , xn , Cn , n
in M2n+1 &n + 2. It follows that

() proj3 (un ) is realized by An , a n , xn in M2n+1 &n + 2.

Pick some Dn M2n+1 &n + 1, a dense set in col(, n ). The precise way we pick Dn
uses the enumeration  and will be explained later. Using Lemma 1A.12 find n < n ,
greater than n = dom(un ), so that the n -type of An , a n , xn , Dn , n in M2n+1 &n + 5
is elastic. Let wn be this type. Observe that wn exceeds proj3 (un ) in the model M2n+1 .
This follows as in case 1, using () and noting that n + 5 (the level of realization
of wn ) is smaller than n + 2. Applying the One Step Lemma inside M2n+1 find
E2n+1 M2n+1 &j0,2n+1 (), a n + -strong extender with critical point n , so that wn
E 3
is a subtype of Stretchn2n+1
+ (proj (un )). As usual make sure that E2n+1 is stronger than
all previous extenders on the tree.
Let M2n+2 = Ult(M2n , E2n+1 ) and let j2n,2n+2 be the ultrapower embedding. As
in case 1 we get
wn < proj3 (j2n,2n+2 (un )).
Using this it is again easy to check that wn is a legal move for II in j0,2n+2 (A)[x]
following the position Qn . It follows that E2n , E2n+1 , wn as defined above form a
legal move for II in round n of Apiv [x]. This completes the construction in round n.
Conditions (A)(C), (E), and (F) are easy to verify. Condition (D) was verified above.
# (Case 2)
34 1 Basic components

Let us take note of what we have so far. We defined a strategy for II in the game
Apiv [x], modulo some method of picking the dense sets Dn . Every run according to
our strategy satisfies conditions (A)(F) and () of cases 1 and 2. Now un is required
to include certain formulae, listed in the rules of A[x], specifically rules (4f)(4h) in
Section 1A (2). Applying these formulae to the realization given by () we get in the
case ln N:
$a n , xn % An .
M
(1) pn col(,
2n+1
n)

a n (i) = ai .
M
(2) For each i < ln , pn col(,
2n+1
n)

M
(3) For each i < ln , pn col(,
2n+1
n)
xn (i) = xi .
Moreover, using rule (4c) we get:
(4) pn extends pln .
(We also use here the fact that j2ln ,2ln +2 (pln ) = pln so that pln is not affected by
the shift of aln . This fact follows from the rules in Section 1A (2), which tell us that
pln M2ln &ln . Remember that crit(j2ln ,2ln +2 ) = ln by construction.)
For each odd branch b of T , let hb be the filter generated by the increasing conditions
{pn | 2n + 1 b}. Let Ab = jb (A) and let b = jb (). Let a b = j2n+1,b (a n ) for
some n so that 2n + 1 b. Which n we take does not matter; this follows from
condition (C). Similarly let xb = j2n+1,b (xn ). Transferring conditions (1)(3) to Mb
using the embedding j2n+1,b we get:

(1 ) hb M b
 b , xb % Ab .
col(,b ) $a

(2 ) For each i < ln , hb M b


 b (i) = ai .
col(,b ) a

(3 ) For each i < ln , hb M


col(,b ) xb (i) = xi .
b

Note that we use here the fact that pn is not moved by the embedding j2n+1,b . To see that
pn is not moved by j2n+1,b , observe that pn M2n+1 &n by rule (4c) in Section 1A (2),
n < n by condition (E), and n = crit(j2n+1,b ) by construction. A similar argument
shows that xi and ai for i < ln are not moved.
Let Db = {j2l+1,b (Dl ) | 2l + 1 b} = {j2n+1,b (Cn ) | 2n + 1 b ln N}.
Using condition (D), and again using the fact that pn is not moved by j2n+1,b we see
that:
(5) hb intersects all the dense sets in Db .
Claim 1C.3. Assume that for each odd branch b, Db contains all the dense sets in
col(, b ) which belong to Mb . Then T , a is a pivot.
Proof. This is immediate using condition (5), which under the assumption of the claim
says that hb is col(, b )-generic/Mb , and conditions (1 )(3 ) which say that $
a , x%
Ab [hb ]. #
1C Pivots 35

Remember that our goal is to make sure that all runs according to piv [, x] are
pivots. Claim 1C.3 says that the strategy we constructed achieves this, provided we
pick the sets Dn so as to satisfy the assumption of the claim.
The assumption of the claim is not hard to satisfy. Each dense set in Mb has a
pre-image in M2n+1 for some 2n + 1 b. Each dense set in M2n+1 in turn belongs to
M& + 1. (M2n+1 is a finite iterate of M, and therefore contained in M. is not moved
by a finite iteration tree using only extenders from below , so M2n+1 &n + 1 is a subset
of M& + 1.) So certainly

Db j2n+1,b  (M& + 1 M2n+1 ).
2n+1b

Using this inclusion it is easy to devise a method of picking sets Dn during the con-
struction, which uses the map  : M& + 1, and which secures the assumption
of Claim 1C.3 whenever  is onto. Devising the precise method is a simple matter of
book-keeping, which we leave to the pleasure of the reader. Let us only say that the
book-keeping can be arranged so that the choice of Dn in round n depends only on
n + 1. (If n + 1 does not give a suitable set, simply delay by taking the trivial
dense set, Dn = {all conditions}.)
Remark 1C.4. It was not necessary during the course of the construction to pick dense
sets Dn . All we needed were the types wn of these dense sets. Types, unlike the dense
sets, are elements of M&. Taking advantage of this observation, one can revise the
construction to use a generic enumeration of M& instead of the enumeration  of
M& + 1. We need a generic enumeration of M& because we cannot just play all types
(for one thing, {n }n< must be increasing). The use of genericity here is similar to its
use in Section 1B. We refer the reader to the final stages of case I in [32, Section 4] for
a detailed construction of this kind.

1C (3) Properties of piv . Given x X and  : M& + 1 our construction of


the previous subsection produces a strategy piv [, x] playing for II in Apiv [x]. We
have:
Lemma 1C.5. Suppose that  is onto M& + 1. Then every run according to piv [, x]
is a pivot for x.
Proof. This follows from our work in the previous subsection, particularly Claim 1C.3
and the book-keeping for picking the Dn -s modulo . #
The construction of piv [, x] is Lipschitz continuous. To be precise, for : n
M& + 1 and s Xn the construction produces a strategy piv [, s], playing for II in
Apiv [s]. We refer to the map piv = (, s  piv [, s]) as the pivot strategies map
associated to A, , and X. We have, for x X and  : M& + 1,

piv [, x] = piv [n, xn].
n<
36 1 Basic components

It is clear that the map piv belongs to M. Our description in Section 1C (2) in fact
defines piv inside M.

Remark 1C.6. The description in Section 1C (2) leaves the precise choice of E2n ,
n , and E2n+1 open. To literally obtain the map piv we must fix, inside M, some
wellordering of M&, and take, in each stage of the construction where the description
leaves room for choice, the first suitable object.

We say that a pivot T , a resides above an ordinal < if all the extenders used
in T have critical points higher than . Our construction in Section 1C (2) clearly yields:

Lemma 1C.7. All runs according to piv [, x] reside above rank(X). #

In applications we will often have to make sure that our pivots reside above a given
ordinal < . We can achieve this indirectly by setting Y = X M&, and working
with the maps A, Apiv , and piv associated to A, , and Y , instead of the maps associated
to A, , and X. Since A is also a name for a subset of (M&) Y , we are allowed
to do this. The change from X to Y has no effect on any of the previous arguments,
except that it forces the n -s to be greater than .
Suppose we are handed some ZFC model P and an elementary embedding
: M P . The maps A, Apiv , and piv , being elements of M, can be shifted to
P using the embedding . The relevant notions (particularly that of a pivot) can be
shifted similarly. For x (X) , and a map  : P &() + 1, let

(piv )[ , x ] = (piv )[ n, x n].
n<

Lemma 1C.8. Suppose  is onto P &() + 1. Then all plays according to


(piv )[ , x ] are pivots for x , where pivot is interpreted over P and using (A).
Furthermore, all plays according to (piv )[ , x ] reside above (rank(X)).

Proof. (piv )[ , x ] is exactly the strategy we would get if we ran the construction
of Section 1C (2) in P instead of M. Lemmas 1C.5 and 1C.7 would apply equally well
in P . #

This lemma will be used heavily throughout the book. In applications P and
will be handed to us in installments. We will be given n : M n M n+1 at a stage
n, with M 0 = M. P and : M P will be the direct limit model and embedding.
Our goal will be to build a pivot for x over P , also in installments. We will do this
using Lemma 1C.8. Without going into details, let us only say that the construction
will involve fixing n : n M n & 0,n () + 1 so that n+1 extends n ( n ) and so that

= n, ( n )
n<

is onto P &() + 1.
1D Mirror images 37

Remark 1C.9. For a fixed  : M& + 1, the map x  piv [, x] is Lipschitz
continuous in x. Why did we not suppress , and use piv to denote the map s 
piv [, s]? The reason is that this map (certainly when  is onto) need not belong to M.
This wasnt a problem before. Remember that the map gen only belonged to M[g],
not to M. But here we do need a map which belongs to M, for otherwise we could not
even phrase, let alone prove, Lemma 1C.8.

1D Mirror images
So far we defined the auxiliary game map A = (s  A[x]), defined the notions of
generic runs and pivots, and obtained the properties listed in the introduction to this
chapter, Properties 13.
Recall that Properties 2 and 3 are in some sense dual. The former places us outside
A[g], while the latter places us inside some shift jb (A)[h]. There is however a qualitative
difference between these two properties. The statement of Property 2 is only useful if
x and a belong to M[g]. Part (2) of Property 3 on the other hand tells us that x and a
belong to Mb [h].
Property 3 is thus substantially more potent, and we will always try to work with
it, rather than with Property 2, in applications. To avoid the situation of Property 2 we
consider both the auxiliary game associated to A, and the auxiliary game associated to
the complement of A with the auxiliary roles of I and II reversed. We take this section
to briefly explain the notation involved in reversing the roles of I and II. For an example
of how we use this reversal to avoid the situation of Property 2 see Section 1E.
We work as always with a ZFC model M, its Woodin cardinal , an X M&,
and some g which is col(, )-generic/M. Let B be a name in M for a subset of
(M&) X in M[g]. For each x X we define a game B[x]. Our definition
mirrors the definition of A[x] in Section 1A.
B[x] is played according to Diagram 1.3. Note that this diagram is quite simply
Diagram 1.1, with the roles of I and II reversed. The rules of B[x] are precisely the

I w0 w1 w2 ...
II l0 , p0 , u0 l1 , p1 , u1 l2 , p2 , u2

Diagram 1.3. The game B[x].

rules (1)(4) listed in Section 1A(2), but with I and II reversed, and with rule (2d)
replaced by:
(2) (d) There exist some names b and x, both members of M& + 1, so that un is
 x, in M& + 2.
realized by B, b,
(Note the change from A in the realization of un to B.) The dependence of B[x]
on x is Lipschitz continuous. We have therefore a map B = (s  B[s]). We refer
38 1 Basic components

to B as the mirrored auxiliary games map associated to B, , and X. It is clear that


this map belongs to M. We use b = $b0 , b1 , b2 , . . . % to denote runs of B[x], with
bn = $ln , pn , un , wn %.
Fix a bijection : M& in M.
Definition 1D.1. b is a generic run of B[x] (wrt , g) if for each n < , $ln , pn , un %
is equal to g(e) where e is least so that g(e) is a legal move for II in B[x]

following bn.
The next two lemmas are the mirror images of Lemmas 1B.2 and 1B.7. Note the
change from strategy for I in Lemma 1B.7 to strategy for II in Lemma 1D.3.
Lemma 1D.2. Suppose that b is a generic run of B[x]. Then $b,
 x%  B[g]. (Note,

this is only useful if x and b belong to M[g].) #
Lemma 1D.3. For every real x there exists a strategy gen [x] for II in B[x], so that
every run according to gen [x] is generic. Moreover, the association x  gen [x]
is Lipschitz continuous, induced by some map gen = (s  gen [s]). The map gen
belongs to M[g]. #
We refer to the map gen of Lemma 1D.3 as the mirrored generic strategies map
associated to B, , and X.
Let us now pass to the matter of pivots. We use the same word, pivot, in our
definitions for the mirror image. When there is danger of ambiguity we distinguish
between A-pivots and B-pivots.
Definition 1D.4. A (B)-pivot for x X is a pair T , b where
(1) T is a nice iteration tree on M,
(2) b is a run of jeven (B)[x],
(3) for every odd branch c of T there exists h so that
(a) h is col(, jc ())-generic/Mc , and
 x% jc (B)[h].
(b) $b,
The game Bpiv [x] is played according to Diagram 1.2 and the rules listed in Sec-
tion 1C (1), except that I and II are reversed, A is replaced by B, and A is replaced
by B. We now use bn to denote j2n,2n+2 (ln , pn , un ), wn . We let b = $b0 , b1 , . . . %.
A run of Bpiv [x] produces T , b which satisfy the first two conditions in Definition 1D.4.
The construction in Section 1C(2) can be mirrored to produce piv [, x], a strategy
for I (note as usual the reversal of I and II) in Bpiv [x]. piv [, x] is given by a
Lipschitz continuous map piv = (, s  piv [, s]) which belongs to M. We have
Lemma 1D.5. Suppose  is onto M& + 1. Then every run according to piv [, x] is
a B-pivot for x. #
This parallels Lemma 1C.5. Lemma 1C.8 can be mirrored similarly.
1E Sample application 39

1E Sample application
To illustrate the use of the techniques developed in the previous sections we include
here a proof of the following theorem, which concerns standard games of length . The
theorem is originally due to Woodin, proved by methods which predate the developments
in this book. Our objective in this book is to handle games of lengths greater than .
Theorem 1E.1 is only included as an illustration. More results are known on length
games within the projective hierarchy. For details on the determinacy of games of
length we refer the reader to MartinSteel [18], Neeman [29], and Neeman [32].

Theorem 1E.1 (Woodin). Suppose that there exists a model M and a M so that:

M is a class model of ZFC;

M is weakly iterable (see Appendix A for the relevant definition);

is a Woodin cardinal of M; and

M& + 1 is countable in V.

Then 12 determinacy holds.

Remark 1E.2. The assumption in Theorem 1E.1 follows from the existence of a sharp
for a Woodin cardinal in V (more precisely the existence of a Woodin cardinal in V
and a sharp for V ), see Appendix A.

The proof of Theorem 1E.1 takes the rest of this section. Fix a 12 set C R, say
the set of all reals which satisfy a given 12 statement . Our goal is to prove that the
standard length game G (C) is determined.
Fix some g which is col(, )-generic over M. Such generics exist since we are
assuming that M& + 1 is countable in V. Let A M be the canonical name for the
set of $
a , x% (M&) in M[g] so that holds of x in M[g]. Let B M be the
canonical name for the set of $b,  x% (M&) so that does not hold of x in
M[g]. Let A be the auxiliary games map associated to A, , and X = . Let B be the
mirrored auxiliary games map associated to B, , and X = .
Working in M, define a game G played as follows: I and II alternate playing
natural numbers xn , creating together the real x = $xn | n < %. In addition they play
moves anI = $ln , pn , un % and anII = wn in the auxiliary game A[x] associated to x.
The first player to violate these rules loses. If the rules are successfully followed for
moves then II wins.

I x0 a0I a1I x2 ...


II a0II x1 a1II ...

Diagram 1.4. The game G .


40 1 Basic components

We point out that our definition of G rests on the continuity of the map x  A[x].
The rules governing the moves ln , pn , un and wn depend only on xn, and are therefore
available in round n of G . We point out further that the game G exists in M. This is
true since the map A = (s  A[s]) exists in M.
Mirroring G , let H be the played as follows: Again I and II alternate playing
natural numbers xn , creating together the real x = $xn | n < %. But now they play
auxiliary moves in the game B[x]. The first player to violate these rules loses. Infinite
plays here are won by I.

I x0 b0I b1I x2 ...


II b0II x1 b1II ...

Diagram 1.5. The game H .

G M is an open game for player I, hence determined in M. Similarly H is


open for II, and hence determined in M. We will use this determinacy as part of our
proof of Theorem 1E.1.

Case 1. If I wins G in M. We intend to argue that I wins G (C) (in V).


Fix M, a winning strategy for I in G . Fix some imaginary opponent, playing
for II in the standard game on natural numbers. We describe how to play against this
imaginary opponent, and win. Our description takes the form of a construction of
(among other things) the run x = $xn | n < %. Our opponent plays xn for odd n and
it is our task to construct xn for even n. Once we complete the construction we will
verify that x C, so that indeed the run constructed is won by I.
Let piv be the pivot strategies map associated to A, , and X. Fix in V a surjection
 : M& + 1. Surjections of this kind exist since we are assuming that M& + 1
is countable in V.
We construct x = $xn | n < %, a run of the standard game on natural numbers,
and T , a , a pivot for x. The participants in the construction are:

Our imaginary opponent, producing xn for odd n.

piv [, x], producing wn for all n and the extenders which give rise to T .

and its images along the even branch of T , producing xn for even n and
ln , pn , un for all n.

The time line of the construction is presented in Diagram 1.6. At the start of round n
we have xn = $x0 , . . . , xn1 %, the iteration tree T 2n + 1 ending with M2n , and the
position Pn = $a0 , . . . , an1 % in j0,2n (A)[xn]. If n is odd, our opponent opens the
round producing xn . If n is even j0,2n ( ) opens the round producing xn . (Remember
is a strategy for I in G , the game displayed in Diagram 1.4.) Then j0,2n ( )
produces ln , pn , un which is a legal move for I in j0,2n (A)[xn + 1]. At this point we
apply piv [, xn + 1]. piv extends the iteration tree producing M2n+1 and M2n+2 .
1E Sample application 41

  
M =M0 M1 M2 M3 M4 M5 M6

_
x0

l0
p0 /o /o /o /o /o /o /o /
u0_
piv w0

Oppnt x1

l1
j0,2 ( ) p1 /o /o /o /o /o /o /o /
u1_
piv w1

j0,4 ( ) x2

l2
j0,4 ( ) p2 /o /o /o /o /o /o /o /
u2_
piv w2

Diagram 1.6. The construction in Case 1.

We let Qn = j2n,2n+2 (Pn , ln , pn , un ). Note that by rule (5) in Section 1C (1), Qn


is equal to Pn , j2n,2n+2 (ln , pn , un ). piv further produces wn , a legal move for II in
j0,2n+2 (A)[xn + 1] following Qn . This completes round n of the construction. We
set an = j2n,2n+2 (ln , pn , un ), wn , so that Pn+1 = Qn , wn = $a0 , . . . , an1 , an %,
and proceed to the next round.
42 1 Basic components

Remark 1E.3. We again point out the importance of continuity, this time the continuity
of the map x  piv [, x]. In round n of the construction we use piv [, xn + 1] for
round n of Apiv [xn + 1].
Note that x and a together form an infinite play of j0,even (G ) which is according
to j0,even ( ). This follows from our use of and its images during the construction,
and from rule (5) in Section 1C(1) which tells us that an = j2n+2,even (an ).
Now M is a winning strategy for I, the open player. It follows that there
are no infinite plays according to . On the other hand we just saw that there is an
infinite play according to j0,even ( ), the play x, a . If Meven were wellfounded, the
existence of such a play could be reflected into Meven , and then pulled back by the
elementarity of j0,even to give an infinite play according to . We thus conclude that
Meven is illfounded.
Using the iterability of M, pick a branch b of T so that Mb is wellfounded. We
know that b is an odd branch. Our use of piv [, x] during the construction guarantees
that T , a is a pivot for x; see Lemma 1C.5. In particular (see Definition 1C.1) there
exists some h which is col(, jb ())-generic/Mb and so that:

() $
a , x% jb (A)[h].

() and our definition of A tell us that holds of x in Mb [h]. is a 12 statement, hence
absolute between wellfounded class models of ZFC . We conclude that holds of x in
V, as required. # (Case 1)

Case 2. If II wins H in M. Then an argument which mirrors that of case 1 (this time
using piv ) shows that II wins G (C) in V. # (Case 2)

Case 3. Since G and H are both determined in M, the only remaining possibility is
that II wins G in M and I wins H in M. We intend to derive a contradiction in this
case. This will show that either case 1 or case 2 must hold, completing the proof of
Theorem 1E.1.
Fix M, a winning strategy for II in G . Fix similarly M, a winning
strategy for I in H .
Let gen be the generic strategies map associated to A, , and X. Let gen be the
mirrored generic strategies map associated to B, , and X. Working inside M[g] we
construct x, a , and b so that

(1) x = $xn | n < % is a real;

(2) a , x is a run of G played according to both and gen [x];

 x is a run of H played according to both and gen [x].


(3) b,

The construction is straightforward: gen produces the auxiliary moves anI ; pro-
duces the auxiliary moves anII and the numbers xn for odd n; produces the numbers
xn for even n and the auxiliary moves bnI ; and gen produces the auxiliary moves bnII .
1F Mixed pivots 43

Remark 1E.4. As usual the continuity of the maps x  gen [x] and x  gen [x] is
important; when using gen and gen in round n we only have knowledge of xn + 1.
Condition (2) tells us that $
a , x%  A[g]; see Lemma 1B.2. Similarly condition (3)
 x%  B[g]. Since a , b,
tells us that $b,  x M[g] this means that on the one hand
does not hold of x in M[g], and on the other hand does not fail for x in M[g]. This
is a contradiction. # (Case 3, Theorem 1E.1)
Remark 1E.5. Running the proof above with C = R, or more precisely with the 12
statement x = x, we see that for every real x: there exists a length iteration tree T
on M which has an illfounded even branch and such that for each odd branch b of T
there is some h with
(1) h is col(, jb ())-generic/Mb ; and
(2) x Mb [h].
Thus every real can be absorbed into a generic extension of an iterate of M. This was
noted in Neeman [29]. A different genericity iteration was previously discovered by
Woodin [45].

1F Mixed pivots
We work as usual with reference to the fixed M, , and X. But now we do not fix a
single name A. Instead we fix a map A = (  A[ ]) in M, which assigns to each
ordinal < jumpM (), a name A[ ] for a subset of (M&) X in M col(,) .
Remark 1F.1. The point of the restriction to < jumpM () is to allow us to work with
a set function A which literally belongs to M, rather than a class function definable over
M. There is nothing particularly important about jumpM (), except that it is larger than
the ordinals we shall care about in applications later on.
We work to describe games which are similar to the pivot games of Section 1C,
but allow player I some greater flexibility. This greater flexibility is the result of two
changes. We allow I to insert intervals to the iteration tree constructed; and we allow I
to pick which of the names A[ ] to play on in each round of the game. Outcomes of
these new games which satisfy a condition similar to condition (3) of Definition 1C.1
we call mixed pivots. As in Section 1C we go on to describe a strategy for II which
produces mixed pivots.

1F (1) The game. For each < jumpM () let A[ ] = (s  A[ , s]) be the auxiliary
games map associated to the name A[ ]. We use A to denote the map , s  A[ , s].
This map belongs to M.
Fix x X . We define the game Amix [x] associated to x and to the map A =
(  A[ ]).
44 1 Basic components

At the start of round n we have a natural number e(n) and an iteration tree T e(n)+1,
ending with the model Me(n) . We have some position Pn = $a0 , . . . , an1 % in Me(n) .
For n = 0 we set e(0) = 0, M0 = M.

I f (n), T f (n) + 1 n ln , pn , un
II Ef (n) , Ef (n)+1 , wn ...

Diagram 1.7. Round n of the game Amix [x].

The time line of round n is presented in Diagrams 1.7 and 1.8. At the start of round n
player I plays some natural number f (n) e(n) and extends T e(n)+1 to T f (n)+1.
We demand enough agreement between Me(n) and Mf (n) so that Pn belongs to Mf (n) .
Player I then plays n so that Pn is a legal position in j0,f (n) (A)[n , x].
The rest of the round follows the rules in Section 1C: I plays a move in
j0,f (n) (A)[n , x] following the position Pn ; II shifts this move to the model Mf (n)+2 ,
as illustrated by the squiggly arrow in Diagram 1.8, and replies there. We let Pn+1 be
the position obtained, let e(n + 1) = f (n) + 2, and proceed to the next round.

Remark 1F.2. Suppose I fixes 0 and always plays f (n) = e(n) and n = j0,f (n) (0 ).
Then the game degenerates into the pivot game associated to the name A[0 ]. Thus the
difference between our current game and the pivot games of Section 1C is in the extra
flexibility accorded to I at the start of the round. I may play f (n) > e(n) and insert
her own interval of models into the tree T . In addition, I may pick a new ordinal n to
work with.

The exact rules of Amix [x] are as follows:

(1) (Rule for I) f (n) e(n) and T f (n) + 1 extends T e(n) + 1. We make the
following requirements on the extended iteration tree:

(a) T f (n) + 1 is normal and uses only extenders taken from below ;
(b) Mf (n) and Me(n) are in sufficient agreement that Pn Mf (n) (more precisely
they agree beyond the rank of Pn );
(c) the extenders Ek for k [e(n), f (n)) have critical points above rank(X);
and
(d) for k [e(n), f (n)), the T -predecessor of k + 1 does not belong to any of
the intervals [e(m), f (m)), m < n.

Read differently, rule (1d) states that none of the nodes in the interval of numbers
[e(m), f (m)) can ever be used as an immediate predecessor in future moves by I.

(2) (Rule for I) n is an ordinal chosen so that Pn is a legal position in j0,f (n) (A)[n , x].
1F Mixed pivots 45

II

Me(n) _ _ I _ _ Mf (n) Mf (n)+1 Mf (n)+2 _ _ I _ _ Mf (n+1)

Pn
&
-
I 6 n
A
K _ _
R Y
- Pn

ln
I pn /o /o /o /o /o /o /o /o /o /o /
un_

II wn
_ Pn+1
#
&
)
-
I n+1
4
@
'
Pn+1

Diagram 1.8. Round n of Amix [x] and the beginning of round n + 1.

(3) (Rule for I) ln , pn , un is a legal move for I in j0,f (n) (A)[n , x] following Pn .

As in Section 1C(1) we intend to extend the tree order T by setting:

(i) the T -predecessor of f (n) + 2 is f (n);

(ii) if ln = new then the T -predecessor of f (n) + 1 is f (n); and

(iii) if ln N then the T -predecessor of f (n) + 1 is f (ln ) + 1.

Continuing with the rules of Amix [x]:


46 1 Basic components

(4) (Rule for II) The extenders Ef (n) , Ef (n)+1 must be played in a way which agrees
with conditions (i)(iii) above. Precisely:
(a) If ln N then Ef (n) Mf (n) must be an extender with critical point within
the level of agreement between Mf (n) and Mf (ln )+1 . We then set

Mf (n)+1 = Ult(Mf (ln )+1 , Ef (n) ).

(b) Ef (n)+1 Mf (n)+1 must be an extender with critical point within the level
of agreement between Mf (n)+1 and Mf (n) . We set

Mf (n)+2 = Ult(Mf (n) , Ef (n)+1 ).

We have now the extended tree T f (n) + 3, ending with Mf (n)+2 defined above.
In line with the rules in Section 1C(1) we make the following additional restriction on
the way II forms the models Mf (n)+1 and Mf (n)+2 :
(5) (Rule for II) If ln = new then Ef (n) must equal pad so that Mf (n)+1 = Mf (n)
and jf (n),f (n)+1 = id.
(6) (Rule for II) T f (n) + 3 must be normal and use only extenders taken from
below .
Having specified Mf (n)+2 we let Qn be the position jf (n),f (n)+2 (Pn , ln , pn , un ).
(7) (Rule for II) The critical point of jf (n),f (n)+2 must be larger than rank(X) and
large enough that Pn is not moved by this embedding.
(8) (Rule for II) wn must be a legal move for player II in the shifted game
j0,f (n)+2 (A)[jf (n),f (n)+2 (n ), x], following the position Qn .
We set Pn+1 = Qn , wn , set e(n + 1) = f (n) + 2, and let T e(n + 1) + 1 be the tree
defined above, ending with the model Me(n+1)=f (n)+2 . This completes the round. We
are now in a position to start round n + 1.
Remark 1F.3. As usual the dependence of our definition on x is Lipschitz continuous;
x only comes in through rules (3) and (8). Thus we are as usual defining a map
Amix = (s  Amix [s]).
Definition 1F.4. We say that round n of a run of Amix [x] does not contain mixing if
f (n) = e(n) and (when n > 0) n = jf (n1),e(n) (n1 ). Otherwise we say that round
n contains mixing.
A run of Amix [x] which does not contain mixing in any round is simply a run of
the pivot game associated to the name A[0 ] and x. This is a precise formulation of
Remark 1F.2.
For each n < let an = jf (n),f (n)+2 (ln , pn , un ), wn . Using rule (7) it is easy
to verify inductively that Pn equals $a0 , . . . , an1 %. Let a = $an | n < %.
1F Mixed pivots 47

Let T be the length iteration tree produced by the run of Amix [x] described above.
An infinite branch of T is even if it contains arbitrarily large nodes in {f (n) | n < }.
Otherwise the branch is odd. To illuminate this terminology we note that in the case of
a run of Amix [x] which does not contain any mixing, {f (n) | n < } is precisely equal
to {0, 2, 4, . . .}, and the only even branch is 0 T 2 T 4 T .
In the general case of a run which does contain mixing, the nodes f (n) need not all
be even. Still the division of branches indicated above is useful, and the reference to
the two kinds of branches as even and odd seems as good as any. Note that in the
general case there may well be more than one even branch. This has to do with the fact
that we do not require e(n) T f (n) in rule (1).

Claim 1F.5. f (0) T l for each l > f (0).

Proof. This follows from rule (1d) and conditions (i)(iii), which together imply that
for all k f (0), the T -predecessor of k + 1 is greater than or equal to f (0). #

Suppose that b is an infinite odd branch of T . From Claim 1F.5 it follows that
f (0) b. Let n < be largest so that f (n) b. We use root(b) to denote this n. We
refer to root(b) as the root of b.
By rule (1d) and conditions (i)(iii), b avoids the intervals [e(m), f (m)) for all
m > root(b). It follows that all nodes of b above f (root(b)) are of the form f (m) + 1.
The behavior of T on nodes of the form f (m) + 1 was specified precisely in conditions
(ii) and (iii). These conditions tell us that b corresponds to a branch in the tree order
given by {ln }n< . More precisely, there is a sequence {nk }k< so that:

n0 = root(b);

ln0 = new;

lnk = nk1 for k > 0; and

{f (n0 )} {f (nk ) + 1 | k < } forms a tail-end of b.

We draw the readers attention to the similarity between the structure of odd branches
here, and the structure of odd branches in Section 1C.

Definition 1F.6. Let P be an infinite run of Amix [x], given by T , a , f , and  say. We
use (P, b) to denote jf (n0 ),b (n0 ) where n0 stands for root(b).

Definition 1F.7. Let P be an infinite run of Amix [x], given by T , a , f , and  say. P
is a mixed pivot for x just in case that for every odd branch b of T there exists some h
so that:

(1) h is col(, jb ())-generic/Mb ; and

(2) $
a , x% jb (A)[b ][h], where b stands for (P, b).
48 1 Basic components

Note the similarity between this and Property 3 in the introduction to this chapter. Of
course here we are working not with a specific name A, but with a function A = ( 
A[ ]). We therefore have to specify which to use in condition (2) of Definition 1F.7.
The we take is the one chosen by player I in the round corresponding to the root of
the odd branch b.

Definition 1F.8. A run of Amix [x] is said to be useful if for all n > 0, n < jf (m),f (n) (m )
where m < n is largest so that f (m) T f (n).

We already saw that f (0) T l for all l > f (0). So the reference to the largest m < n
so that f (m) T f (n) in Definition 1F.8 makes sense.
If T is part of a useful run of Amix [x] then all the even branches of T lead to
illfounded direct limits. An iteration strategy faced with the tree of a useful mixed pivot
is thus forced to pick an odd branch. The branch chosen by the strategy is then subject
to the conditions of Definition 1F.7. It is this observation which makes useful mixed
pivots useful.
We shall talk also of useful position (as opposed to infinite runs). A position of
length i in Amix [x] is useful if it satisfies the condition of Definition 1F.8 for all n < i.
The following claim is clear:

Claim 1F.9. Suppose that P is an infinite run of Amix [x], and that Pi is useful for
each i < . Then P is useful. #

1F (2) The strategy mix [, x]. Fix a surjection  : M& + 1. Fix x X .
Continuing the parallel with Section 1C we wish to describe a strategy mix [, x] for II
in Amix [x] which always produces mixed pivots. Fix some imaginary opponent. We
describe how to construct a run of Amix [x], ourselves playing for II and letting the
imaginary opponent play for I. The run we construct will be a mixed pivot, as required.
The construction is similar to that of Section 1C (2), with only a few modifications.
Rather than describe the whole construction we confine ourselves to highlighting the
modifications.
In addition to constructing the objects T , a , f , and  which form a run of Amix [x],
we construct the following objects:

a n , xn Mf (n)+1 &n + 1 where n = j0,f (n)+1 ();

Dn Mf (n)+1 &n + 1; and

n , an ordinal in Mf (n)+1 .

Note already the similarity with Section 1C(2). The main difference so far is in the
addition of the ordinals n .
We intend to maintain all the requirements set by the rules of Amix [x]. In addition
we intend to maintain:

(A) wn is elastic.
1F Mixed pivots 49

(B) wn is realized by An , a n , xn , Dn , n in Mf (n)+1 &n + 5, where


An = j0,f (n)+1 (A)[n ] and n = j0,f (n)+1 (). (Recall that > is the lower
ordinal in the least pair of local indiscernibles of M relative to .)

(C) If ln N then

a n = jf (ln )+1,f (n)+1 (a ln ),


xn = jf (ln )+1,f (n)+1 (xln ), and
n = jf (ln )+1,f (n)+1 (ln ).

(D) If ln N then pn belongs to jf (ln )+1,f (n)+1 (Dln ).

(E) 0 < 0 < 1 < 1 < , where n = dom(un ) and n = dom(wn ).

(F) Mf (n) and Mf (n)+1 agree to n + . Mf (n)+1 and Mf (n)+2 agree to n + .

(G) If ln = new then Mf (n)+1 = Mf (n) , and n is precisely the ordinal n played
by I (by our imaginary opponent that is) in round n of Amix [x].

We point out several differences between our conditions here and those maintained
in Section 1C (2). First, in condition (B) we now have An = j0,f (n)+1 (A)[n ]; note
the addition of n . Secondly, in condition (C) there is an added clause that n =
jf (ln )+1,f (n)+1 (ln ). Finally, there is the new condition (G), relating the ordinals n
we construct to the ordinals n played by I as part of the run of Amix [x]. Observe that
condition (G) and the final clause in condition (C) precisely determine all the ordinals n .
The construction itself follows closely along the lines of the two cases in Sec-
tion 1C (2). Just as in Section 1C(2) we obtain the following during the construction:

() proj3 (un ) is realized by An , a n , xn in Mf (n)+1 &n +2. (n here is the indiscernible


to , shifted to Mf (n)+1 .)

This, together with the conditions in rule (4) in the Definition of A (Section 1A (2)),
can be used to verify that T , a , f , and  form a mixed pivot. The argument is similar
to the final argument in Section 1C(2), and is left to the reader. We point out only that
the argument uses the new condition (G), and the way it fits with Definition 1F.6.

Remark 1F.10. For the record we note that if ln N then crit(Ef (n) ) = ln , see case 2
in Section 1C (2). Now ln > ln , Pln belongs to Mf (ln ) &ln , and Mf (ln ) and Mf (ln )+1
agree to ln . (All this is by construction, see for example case 1 in Section 1C (2).) It
follows that Pln Mf (ln )+1 is not moved by jf (ln )+1,f (n)+1 .

1F (3) Properties of mix . Let us recall the general framework of this section. We
work with a map A = (  A[ ]) in M which assigns to each ordinal < jumpM ()
some name A[ ] for a subset of (M&) X in M col(,) . In Section 1F (1) we
worked with a specific x X and defined the game Amix [x]. As usual the definition
was Lipschitz continuous in x, giving rise to a map Amix = (s  Amix [s]). It is clear
50 1 Basic components

that this map belongs to M. We refer to Amix as the mixed pivot games map associated
to A, , and X.
Working with some  : M& + 1 we went on and modified the construction
of Section 1C (2) to construct mix [, x], a strategy for II in Amix [x]. This construction
too was Lipschitz continuous, both in x and in , giving rise to a map mix = (, s 
mix [, s]). It is clear that this map belongs to M. We refer to mix as the mixed pivot
strategies map associated to A, , and X.
We have:

Lemma 1F.11. Suppose that  is onto M& +1. Then every run according to mix [, x]
is a mixed pivot for x. #

Continuing the parallel with Section 1C, let us say that a mixed pivot given by T , a ,
f , and  resides above an ordinal < if all extenders used in T have critical points
above . We get:

Lemma 1F.12. All runs according to mix [, x] reside above rank(X).
Proof. The extenders Ef (n) and Ef (n)+1 have critical points above rank(X) by con-
struction; see Section 1C(2) and Lemma 1C.7. The extenders Ek for k [e(n), f (n))
have critical points above rank(X) by rule (1c) in Section 1F (1). #

For future use let us record the following consequence of Remark 1F.10:

Lemma 1F.13. Let P be an infinite run of Amix [x], given by T , a , f , and  say.
Suppose that P is played according to mix [, x]. Let b be an infinite odd branch of T .
Fix a node k b larger than f (root(b)), say k = f (m) + 1. Then a m belongs to Mk
and is not moved by jk,b .

Proof. This follows by repeated applications of Remark 1F.10, going over n > m so
that f (n) + 1 b. #
Chapter 2
Games of fixed countable length

The notions of the previous chapter provide a powerful tool for proving determinacy.
Here we apply this tool to games of fixed countable length. Given a countable 1 and
C R , define the game G (C) to be played as follows: In mega-round players I
and II alternate natural numbers as in Diagram 2.1 to produce y = $y (n) | n < %
R. Once mega-rounds have been played the game ends. If $y | < % belongs to
the payoff set C then I wins. Otherwise II wins.

I y0 (0) y0 (2) ...... y (0) y (2) ......


II y0 (1) ... y (1) ...

Diagram 2.1. The game G (C).

We intend to prove the determinacy of G (C) for countable > 1 and C in the
pointclass <2 11 ,1 from a sharp for 1 + Woodin cardinals. The corresponding
result for = 1 is a well-known theorem of Martin. The case of = 2 with C in
11 reduces to Theorem 1E.1, and proofs similar to the one in Section 1E can in fact
handle all finite ; see Neeman [29], [32] for details. The additional ingredient which
we develop in this chapter is a method for handling limits.
Historical Remark. For sufficiently closed limit ordinals < 1 , ordinals =
where is additively closed, the determinacy we get had been proved earlier by Woodin.
(It was a pleasant consequence of tight equiconsistency results he proved, connecting
Woodin cardinals under choice to measures on [P1 (R)] under AD.) In particular
the determinacy of all <2 11 games of fixed countable lengths from 1 Woodin
cardinals is due to Woodin.

2A General games and iteration games


Each determinacy proof in this book can be viewed as a construction which takes a
general game, decides which player to root for, and reduces this players task to the task
of winning an iteration game on a given model M. When M is iterable this yields the
determinacy of the original game.
1 <2 1 is the pointclass of sets at levels below 2 in the difference hierarchy on 1 sets. It is
1 1
somewhat larger than the simpler pointclass 11 , so its use makes our results stronger (to the point of being
optimal in fact).
52 2 Games of fixed countable length

The reduction itself can be viewed as a strategy in a long game. If rooting for
player I, the game in the case of fixed countable length consists of moves in G ,
together with moves in the weak iteration game on M (see Appendix A) with I playing
for bad and II playing for good. A strategy for I in such a game amounts to a
reduction of Is moves in G to moves for the good player in the iteration game.
Let us be more precise. Fix a countable ordinal . Let M be a ZFC model with
a Woodin cardinal . Suppose that is countable also in M. Let A be a name for
a subset of R in M col(, ) . We define a game G fixed = G fixed (M, , , A) which
begins to formalize a reduction which roots for I.
A run of G fixed serves to construct the following objects:

a sequence of reals y = $y | < %;

a sequence of models $M | 1 + 1% starting with M1 = M; and

a sequence of embeddings j, : M M for + 1.

The reals correspond to a run of G ; the models and embeddings correspond to a run
of the weak iteration game.
fixed is played in 1 + + 1 mega-rounds, indexed 1 through . By the start of
G
mega-round for a successor ordinal we have y( 1) and the model M .

(i) To start mega-round the two players alternate natural number moves in the usual
fashion, producing together the real y 1 .

(ii) Player I then plays an iteration tree T of length on the model M . T must be
normal, plus 2, and use only extenders taken from below the image of .

(iii) Player II ends mega-round by playing a cofinal branch b through T . We let


M +1 be the direct limit along b and let j, +1 : M M +1 be the direct limit
map. The maps j, +1 for < are defined by composition.

By the start of mega-round for a limit ordinal we have y , the models
$M | < %, and the embeddings $j, | < %. We let M be the direct
limit of the system $M , j, | < % and let j, be the direct limit embeddings.
Mega-round is then played according to the rules (ii) and (iii) above, but not rule (i).
Once over the limit mega-round leaves us with y and the model M +1 , precisely the
position necessary to start mega-round + 1.
A run of G fixed results in a sequence of reals y = $y | < %, a final model M +1 ,
and a final embedding j = j1, +1 : M M+1 . If M +1 , or any of the models M for
, is illfounded, then I wins. Otherwise I wins just in case that there exists some h
so that

(1) h is col(, j ( ))-generic/M+1 ; and

(2) y = $y | < % j (A)[h].


2A General games and iteration games 53

Remark 2A.1. A strategy for I in G fixed is precisely a reduction of the kind mentioned
earlier. This will be clarified further below, through the proof of Theorem 2A.3.
Given a name B for a subset of R in M col(, ) we define next a game H fixed =

Hfixed (M, , , B) which begins to formalize a reduction rooting for II. Hfixed pre-
cisely mirrors G fixed . The rules of the two games are the same except that in H fixed it
is player II who plays the iteration trees and player I who plays the branches. It is now
player II who wins if M+1 or any of the preceding models is illfounded. Otherwise II
wins just in case that there exists some h which satisfies condition (1) above, and the
mirrored condition:
(2) y = $y | < % j (B)[h].

Theorem 2A.2 (for 1 countable, both in V and in M). Suppose that there are 1+
Woodin cardinals of M below (where too is a Woodin cardinal of M). Suppose
further that M& + 1 is countable in V. Fix g V which is col(, )-generic/M.
Then at least one of the following three cases holds:
fixed ;
(1) player I has a winning strategy in G
fixed ; or
(2) player II has a winning strategy in H
(3) in M[g] there exists a sequence of reals y = $y | < % which belongs to
neither A[g] nor B[g].
Moreover M can distinguish this. To be precise, there are formulae I and II so that:
if M |= I [ , , A] then case (1) holds; if M |= II [ , , B] then case (2) holds;
and otherwise case (3) holds.
In applications to determinacy we work with A and B which name complementary
sets, so that case (3) is ruled out. Theorem 2A.2 then states that either the task of player I
in G can be reduced to playing for good in a weak iteration game on M, this is
case (1); or else the task of player II in G can be reduced to playing for good in a
weak iteration game on M, this is case (2).
We shall prove Theorem 2A.2 by induction on . The proof is delayed to Sections 2B
(limits) and 2C (successors). We finish this section with an application that shows how
the theorem can be used to prove determinacy.
Theorem 2A.3. Let 1 be a countable ordinal. Suppose that there exists a model
M and a cardinal in M so that:
M is a class model of ZFC;
M is weakly iterable;
is a Woodin cardinal of M;
there are 1 + Woodin cardinals of M below (so that including there
are 1 + + 1 Woodin cardinals in M); and
54 2 Games of fixed countable length

M& + 1 is countable in V and has a sharp.

Then the games G(+1) (C) are determined for all C R +1 in the pointclass
<2 11 .

The assumption in Theorem 2A.3 follows from the existence in V of a sharp for
1 + + 1 Woodin cardinals, by the methods described at the end of Appendix A.

Proof of Theorem 2A.3. By Remark 1E.5 every real can be made generic over an iterate
of M. From this and the fact that M& + 1 has a sharp it follows that every real has
a sharp. By Martin (see Kanamori [11, Theorem 31.4]) we therefore have recourse to
(boldface) <2 11 determinacy.
Without loss of generality we may assume that is countable in M; if not, simply
iterate the first measurable of M past , and pass to a generic extension where is
collapsed.
Fix C R +1 in <2 11 . For y R let Cy = {y R | y, y C}. Note
that Cy R is <2 11 , and therefore determined. Define D R by:

D = { y R | I has a winning strategy in G (Cy )}.

For y  D we know that II has a winning strategy in G (Cy ). It is thus enough to


prove that the game G (D) is determined; whoever wins this game can then continue
to win G(+1) (C).
D belongs to the pointclass  (<2 11 ). By Martin [24] there is a number
k < and a 1 formula so that

y D L[ y , c0 , . . . , ck1 ] whenever c0 < < ck1 are


y ] |= [ (2.1)
Silver indiscernibles for y.

Let (a, c0 , . . . , ck1 ) be the formula L[a] satisfies [a, c0 , . . . , ck1 ]. Let
u0 < < uk1 be uniform Silver indiscernibles. Let A M be the canonical
col(, ) name for the set

A[g] = { y R M[g] | M[g] |= [


y , u0 , . . . , uk1 ]}.

Let B M be the canonical name for the set

B[g] = { y R M[g] | M[g] |= [


y , u0 , . . . , uk1 ]}.

We apply Theorem 2A.2 using these names.

Case 1. If case (1) of Theorem 2A.2 holds. Fix an imaginary opponent willing to play
for II in G (D). We describe how to play against the imaginary opponent, and win.
The description as usual takes the form of a construction of, among other things, a run
of G (D). We verify at the end that the run constructed belongs to D, and is therefore
won by player I.
2A General games and iteration games 55

Using the case assumption fix , a winning strategy for I in the game G fixed . Fix

, a weak iteration strategy for M. Observe that , , and the imaginary opponent
together cover all the moves in G fixed . The imaginary opponent and  collaborate to

produce the reals y ;  plays the trees T ; and  plays the branches b . Letting ,
, and the imaginary opponent play each other we therefore obtain a complete run of
fixed , consisting of a sequence y and a weak iteration leading to a model M +1 and an
G
embedding j = j1, +1 : M M+1 .
Since  is an iteration strategy, M+1 and the models leading to it are all well-
founded. Since  is winning for I in G fixed , there exists some h so that:

(1) h is col(, j ())-generic/M+1 ; and

(2) y j (A)[h].

Condition (2), combined with the definition of A and the fact that M +1 is a well-
founded class model, tells us that

y ] |= [
L[ y , j (u0 ), . . . , j (uk1 )].

It is easy to see that j (u0 ), . . . , j (uk1 ) are Silver indiscernibles for all reals in M +1 [h].
(In fact u0 , . . . , uk1 are fixed points of j since that map belongs to L of a real
essentially a real coding M& and the sequence $T , b | 1 %.) Using
equation (2.1) it follows that y belongs to D, as required. # (Case 1)

Case 2. If case (2) of Theorem 2A.2 holds. Then an argument which mirrors that of
case 1 shows that II wins G (D). # (Case 2)

Case (3) of Theorem 2A.2 cannot occur, since A and B in the current application
name complementary sets. Under case (1) of Theorem 2A.2 we saw that I wins G (D).
Under case (2) we saw that II wins G (D). It follows that G (D) is determined.
# (Theorem 2A.3)

Remark 2A.4. Notice that the winning strategy in G( +1) (C) belongs to L(R, , ).
This is clear from the proof of Theorem 2A.3, as the proof defines the winning strategy
, and M. The strategy 
from , comes from Theorem 2A.2. Our proof of that theorem
will define  (certainly in a projective manner) from any real coding (M& + 1) .
Thus we show not only that G(+1) (C) is determined, but that the player who wins
the game has a winning strategy in L(R, ), where  is a weak iteration strategy for a
model satisfying the assumptions of Theorem 2A.3.
For a fixed C one can do even with slightly less. We did not use the fact that
iterates created using  are wellfounded, but only the (strictly weaker) fact that iteration
embeddings created using  move the type of k indiscernibles correctly for a particular
k < . This requirement on  can be phrased precisely, and any  which satisfies
the requirement is sufficient to define a winning strategy, for the appropriate player, in
G(+1) (C).
56 2 Games of fixed countable length

2B Limits
Fix a countable limit ordinal . We prove Theorem 2A.2 for , assuming that it is known
for < . Our plan is to break the first mega-rounds of the relevant gameeither
fixed or H
G fixed into blocks of mega-rounds, reaching up to ; break mega-round
into parts; and somehow distribute these parts of mega-round between the
blocks of earlier mega-rounds. Induction will allow us to work through each of the
blocks below , and the advance planning involved in distributing parts of mega-round
between these blocks will allow us to work through mega-round .
Fix a model M with a Woodin cardinal . Fix names A and B in M for subsets of
R in M col(, ) . We work under the assumptions of Theorem 2A.2. Specifically, is
countable in M, there are 1 + Woodin cardinals of M below , and M& + 1 is
countable in V.
Let { } [1,) be the first 1 + Woodin cardinals of M listed in increasing order.
We have  < for all [1, ). Since is countable in M, sup{ | [1, )} <
also. Let denote this supremum.
Fix in M a strictly increasing sequence of ordinals k , k < , starting with 0 = 0
and cofinal in . This is possible since is seen to be countable in M. Let k = k+1 k ,
or more precisely let k be the unique ordinal such that k+1 = k + k . Let k be  k+1 .
Each k is a Woodin cardinal of M, and sup{k | k < } = < . Moreover, for
each k < there are precisely 1 + k Woodin cardinals of M below k and (if k > 0)
above k1 .
We think of [k + 1, k+1 ] as a block, and later plan to handle this block by using
Theorem 2A.2 with k and k over a generic extension of M which collapses k1 .
Fix in M some continuous injection of R into R. The precise injection used is not
important, so long as knowledge of yk suffices to determine ( y )(k + 1). We think
of y R as coded by the real ( y ) and carelessly write y instead of (
y ) throughout.

Fix some g V which is col(, )-generic/M. Let A M be the canonical
name for the set (M& )  A[g ] in M[g ]. Similarly let B  M be the canonical
name for the set (M& )  B[g ]. Let A be the auxiliary games map associated
to A , , and X = M&. Let B be the mirrored auxiliary games map associated
to B  , , and X = M&. Intuitively A is a length game where I tries to witness
membership in a shift of A. B is a game where II tries to witness membership in a shift
of B.
We carelessly write A[ y ] instead of A[( y )] and similarly with B. Note that
knowledge of yk suffices to determine the rules for the first k + 1 rounds of A[ y ] and
B[ y ]. For x Rk we shall talk about A[ x ] and B[x ]. Both are games of k + 1 many
rounds.

Remark 2B.1. The use of the set X = M& instead of X = for the auxiliary games
maps will later ensure that the pivots in our constructions reside above = sup{k |
k < }, see Lemma 1C.7 and the comment following it.
2B Limits 57

2B (1) The basic definitions. Working inside M[g ] fix a sequence $gk | k < %
so that gk is col(, k )-generic/M[g k ] where for each k < , g k = [g0 gk1 ].
Below we informally talk about sets in M[g k ][gk ] where formally we should be talking
about names in M[g k ]col(,k ) , or better yet names in M col(,0 )col(,k ) . Similarly
we talk about M[g k ] where formally we should be using the forcing language to talk
about M col(,)col(,k1 ) .
Definition 2B.2 (for k < ). A k-sequence over M[g k ] is a triplet $
x , P , % which
belongs to M[g k ] and so that:
is an ordinal;
x belongs to Rk ; and
P is a position of k rounds in A[
x ].
We think of a k-sequence as the result of progressing through k blocks of mega-
rounds below , to reach an x Rk , and progressing through k parts in the advance
planning for mega-round , to reach a position P of k rounds in A[ x ].
Definition 2B.3 (for k < ). An extended k-sequence over M[g k ] is a triplet
x , P , % M[g k ] so that
$
is an ordinal;
x belongs to Rk ; and
P is a position of k + 1 rounds in A[
x ].
The difference between Definitions 2B.2 and 2B.3 is in the last condition: P is a
position of k rounds, while P is a position of k + 1 rounds. An extended k-sequence
thus involves one more part in the planning toward mega-round .
x , P , % M[g k ] we plan to define:
For each k < and each k-sequence $
x , P , ), played in M[g k ].
A finite game Gk (
The definition will result in a map $
x , P , %  Gk ( x , P , ), belonging to M[g k ].
We use Gk = Gk [g ] to denote this map. Gk M names this map in the forcing
k

col(, 0 ) col(, k1 ).
For each k < and each extended k-sequence $ x , P , % M[g k ] we plan further
to define:
x , P , ) in M[g k ][gk ], subset of Rk . We let Ak [g k ](
A set Ak ( x, P , )
M[g ] be the canonical name for this set.
k

x , P , %  Ak [g k ](
Again the definition will result in a map $ x , P , ), belonging
to M[g ]. We use Ak [g ] to denote this map, and Ak M to denote its canonical name.
k k

Let us begin to define these objects. We work by induction on and . Everything


is done inside M, though for expository simplicity we talk about M[g k ] and M[g k ][gk ]
instead of working with the appropriate forcing languages.
58 2 Games of fixed countable length

Definition 2B.4. The game Gk (


x , P , ) is played according to Diagram 2.2 and the
following rules:
(1) (Rule for I) must be an ordinal smaller than .
(2) (Rule for I and II) akI and akII must be legal moves in A[
x ] following the
position P .

I akI
II akII

Diagram 2.2. The game Gk (


x , P , ).

Once the game is concluded, let ak = $akI , akII % and P = P , ak . Player I wins
x , P , )].
just in case that (M&k+1 )[g k ] |= I [k , k , Ak [g k ](
Moves in Gk ( x , P , ) are actually elements of M. But the payoff condition is
phrased in M[g k ]. Note that the definition of Gk ( x , P , ) requires knowledge of the
x , P , ), but only for < , because of rule (1).
names Ak [g k ](
Remark 2B.5. The payoff condition in the definition of Gk ( x , P , ) involves the
formula I (. . . , k , . . . ). We assume knowledge of this formula as part of the inductive
assumption that Theorem 2A.2 holds for all < .
Remark 2B.6. Note the reference to (M&k+1 )[g k ] in the payoff condition, rather
than M[g k ]. This restriction to a rank initial segment of M[g k ] makes the definition
more local; the definition does not depend on the whole of M, only on the relevant
initial segment. The local definability will be used later on, see Claim 2B.8. There is
nothing specifically important about the particular initial segment taken. We just need
a level which satisfies ZFC , and is large enough to include k and hence the name
x , P , ).
Ak [g k ](
x , P , % is an extended k-sequence and z Rk M[g k ][gk ], then
Note that if $

x z, P , % is a k + 1-sequence.
$
x , P , ) just in case that
Definition 2B.7. For z Rk M[g k ][gk ] put z Ak (
M[g ][gk ] |=I wins Gk+1 (
k
x z, P , ).
It should be stressed that Definitions 2B.4 and 2B.7 are carried out by simultaneous
induction on and . The definition of Gk ( x , P , ) requires knowledge of the names
x , P , ), which in turn depend on the games Gk+1 (
Ak [g k ]( x z, P , ), but only

for < . The induction therefore makes sense.
The definitions produce maps Gk and Ak [g k ], or more precisely names Gk and Ak
for these maps. Since the definitions, when formalized to talk about names, are phrased
inside M,2 $Gk | k < % and $Ak | k < % belong to M. This is important; we shall
2 Remember that the Lipschitz continuous map A, which is relevant to the definition of G (
k x , P , ),
belongs to M.
2B Limits 59

later have to shift the definitions via elementary embedding, and re-interpret them using
various generics. Thus we shall talk about j (Gk )[hk ] for example, where j : M Mk
is elementary and hk is generic for col(, j (0 )) col(, j (k1 )). j (Gk )[hk ] is
the association of Definition 2B.4, but carried out over Mk [hk ].
For the sake of a brief discussion let Sk be the set of k-sequences $ x , P , % so that I
has a winning strategy in Gk ( x , P , ). Combining Definitions 2B.4 and 2B.7 we may
intuitively think of Sk as the set of k-sequences from which I can win to produce: (1)
an ordinal < , (2) a one round extension P of P , and (3) an extension of x by k
mega-rounds to some x Rk+1 , so that $ x , P , % belongs to a shift of Sk+1 . The
moves of (1) and (2) are produced directly through Is strategy in Gk . Then through
the use of the formula I [. . . k . . . ] in the payoff for Gk an inductive application of
Theorem 2A.2 with k shows that I has a strategy to create a shift of Sk+1 and an
extension x of x which enters this shift. Thus from Sk player I can win to enter a shift
of Sk+1 . Then from there I can win to enter a shift of Sk+2 , etc.
This intuition will be made precise in Section 2B (2). Here let us end with a claim
on the local definability of the set Sk alluded to above. The claim will be needed later
on, for example in the proof of Claim 2B.16 in Section 2B (4).

Claim 2B.8. There is a formula , with parameters , {n }n< , {n }n< , {gn }n< ,
and A, so that for all x, P , M[g ], and for any limit ordinal > max{ , }:

$
x , P , % is a k-sequence and I wins Gk [g k ](
x , P , )
x, P (M&)[g ] and (M&)[g ] |= [k, x, P , ].

Proof. Definitions 2B.4 and 2B.7 essentially provide the formula . Note that the only
parameters used in the definitions are the ones listed in the claim. Note further the local
nature of the definitions. Positions P in the games A[ x ] are all elements of M& ,
and the map A is an element of M& + . Each game Gk [g k ]( x , P , ) is an element
of (M& + )[g ], and the same is certainly true of each set Ak [g k ][gk ]( x , P , ).
Furthermore, for any limit ordinal > , (M&)[g ] can figure out the assignments
x , P , %  Gk [g k ](
$ x , P , %  Ak [g k ](
x , P , ) and $ x , P , ), restricted to
< and < . We point out only that this uses the restriction to (M&k+1 ) in the
payoff in Definition 2B.4, see Remark 2B.6. #

A 0-sequence is simply a triplet $, , % for some ordinal M. Thus the end


result of our definitions is finite games G0 (, , ), defined in M for ordinals .

2B (2) I wins. Suppose that there exists some M so that I wins the game
G0 (, , ). We intend to show that I wins G fixed = G
fixed (M, , , A).
Fix an imaginary opponent willing to play for II in G fixed . We describe how to

construct a run of Gfixed , playing against the imaginary opponent. We shall verify at
the end that the run constructed is won by I.
Let Apiv be the pivot games map associated to A , , and X = M&. Let piv be
the pivot strategies map associated to these objects. (Recall that piv is a map which
60 2 Games of fixed countable length

provides strategies for II in the pivot games, see Section 1C (3) and Lemma 1C.5.) These
maps both belong to M and we are free to move them with elementary embeddings.
We divide the construction into stages, followed by a finishing stage. Stage k
will concern mega-rounds [k + 1, k+1 ] of G fixed , and the finishing stage will concern
mega-round . At the start of stage k [0, ) we shall have the objects of items (A)(E)
below.
fixed up to and including mega-round k .
(A) Moves in G
The moves indicated in item (A) produce, among other things: reals $y | < k %;
fixed starts
a model Mk +1 ; and an embedding j1,k +1 : M1 Mk +1 . (Recall that G
with mega-round 1 and M1 = M.) We refer to Mk +1 as Nk , to j1,k +1 as ik , and to
$y | < k % as xk .
(B) Some hk = h0 hk1 which is col(, i1 (0 )) col(, ik (k1 ))-
generic/Nk .
(C) A finite map k : k Nk &ik ( ) + 1.
xk ], played according to
(D) A position Pk of k rounds in the pivot game ik (Apiv )[
the strategy ik (piv )[k , xk ].
The position Pk indicated in item (D) includes an iteration tree on Nk of length 2k + 1.
We use Uk to denote this tree, use W0k , . . . , W2k
k to denote its models, and use k to
,
denote its embeddings. Pk further includes a position Pk of k rounds in the auxiliary
k
game (0,2k ik )(A)[
xk ].
k .
(E) An ordinal k in W2k
We point out that Uk only uses extenders with critical points above ik (), see
Remark 2B.1. In particular i0 (0 ), . . . , ik1 (k1 ) are well below the critical points
used in Uk . We can thus think of Uk as an iteration tree on Nk [hk ], giving rise to the
k [h ].
final model W2k k
We shall make sure that:
k [hk ].
(i) xk belongs to W2k
xk , Pk , k % is a k-sequence over the model W2k
Note how this means that $ k [hk ]. We shall

further make sure that:


k [hk ] |=I wins ( k
0,2k ik )(Gk )[h ](
(ii) W2k k x , P , ).
k k k

This last condition serves to keep the construction going. It corresponds to the statement
xk , Pk , k % belongs to a shift of the set Sk mentioned in the brief discussion in
that $
Section 2B (1). Remember that intuitively player I can win from Sk to enter a shift of
Sk+1 . To perpetuate condition (ii) we simply have to make this intuition precise.
For the start of stage 0 set W00 = N0 = M. Condition (i) holds trivially, as x0 = .
The existence of some 0 which satisfies condition (ii) is given by the case assumption,
that there is some M so that I wins G0 (, , ).
2B Limits 61

Let us start stage k of the construction, assuming inductively that the previous stages
were taken care of. Fix some k+1 : k + 1 Nk &ik ( ) + 1 which extends k . The
precise way we extend k has to do with book-keeping demands and will be explained
later on.
Appealing to condition (ii) fix k W2k k [hk ], a winning strategy for I in ( k
0,2k
ik )(Gk )[h ](
k
xk , Pk , k ). Let Gk denote this game, which follows Diagram 2.2. The
strategy k opens the game by playing:

(a) k+1 < k ; and

(b) a kI , a legal move for I in (0,2k


k ik )(A)[
xk ] following the position Pk .

We continue the construction through a use of ik (piv )[k+1 , xk ]. This strategy creates
an extension of Uk , takes the move a kI of condition (b) which was played in the final
even model of Uk , shifts it to the final even model of the extension, and replies to it
there.
More precisely, ik (piv )[k+1 , xk ] produces a two model extension of Uk , including
models W2k+1 k k
, W2k+2 k
and an embedding 2k,2k+2 . Let Uk+1 denote the extended tree.
ik (piv )[k+1 , xk ] continues by shifting (0,2k ik )(A)[
k xk ] to M2k+2
k and playing:
k
(c) A move akII for II in (0,2k+2 ik )(A)[
xk ] following the shifted position
2k,2k+2 (Pk , a kI ).
k

Let k+1 = 2k,2k+2


k (k+1 ). Let akI = 2k,2k+2
k (a kI ), and let ak = $akI , akII %.
Let Pk+1 = 2k,2k+2 (Pk ), ak . The triplet k+1 , akI , akII represents a run of
k

k ), played according to = k
(G ( ). Since is a winning strat-
2k,2k+2 k k 2k,2k+2 k k
egy for I we get:
k
(iii) (W2k+2 &k+1
s )[hk ] |= [ s , , C ], where s = ( k
I k k k k 0,2k+2 ik )(k ), k+1 =
s
k
(0,2k+2 ik )(k+1 ), and Ck = (0,2k+2
k ik )(Ak )[hk ](
xk , Pk+1 , k+1 ).

Remember that all the extenders used in Uk+1 have critical points above the image
of = sup{n | n < }. The forcing which gives rise to hk resides well below
k
these critical points. The embedding 0,2k+2 : Nk W2k+2k therefore extends to an
embedding 0,2k+2 : Nk [h ] W2k+2 [h ]. Ck is a canonical name for a subset of
k k k k

Rk in W2k+2
k [hk ]. As such it resides well below the critical point of 0,2k+2
k . Thus Ck
belongs to Nk [h ], and 0,2k+2 (Ck ) = Ck . Using condition (iii) and pulling back by
k k
k
0,2k+2 we conclude that:

(iv) (Nk &ik (k+1 ))[hk ] |= I [ik (k ), k , Ck ].

We now appeal to the main inductive assumption in this section, that Theorem 2A.2
holds for < . Combined with condition (iv) it tells us that player I wins the game
62 2 Games of fixed countable length

fixed (N [hk ], ik (k ), k , Ck ), where N = Nk &ik (k+1 ). Fix 


G k , a winning strategy for
I in this game.
Note that G fixed (Nk [hk ], ik (k ), k , Ck ) are really
fixed (N [hk ], ik (k ), k , Ck ) and G
the same game. The reason is rule (ii) in Section 2A, which forces the iteration trees
played in G fixed (Nk [hk ], ik (k ), k , Ck ) to only use extenders taken from below the
image of k . In particular these trees are then iteration trees on the shorter model
N [hk ]. Thus  k is a winning strategy for I in the game G fixed (Nk [hk ], ik (k ), k , Ck ).

Remark 2B.9. We could have saved ourselves the roundabout argument in the preced-
ing paragraph, if we used M[g k ] instead of (M&k+1 )[g k ] in Definition 2B.4. But that
would have ruined the local nature of Definition 2B.4.
Nk is equal to Mk +1 , the starting model for mega-round k +1 of the game G fixed =

Gfixed (M, , , A). Mega-rounds [1, k ] in Gfixed (Nk [h ], ik (k ), k , Ck ) therefore
k

correspond precisely to mega-rounds [k + 1, k + k = k+1 ] in the main game G fixed .


(Note that iteration trees on Nk [h ] can be viewed as iteration trees on Nk , see Claim 11
k

in Appendix A.)  k can thus be viewed as a strategy for I handling mega-rounds


[k + 1, k+1 ] in the main game. Let  k play these mega-rounds against the imaginary
opponent. Together they produce:
(d) the reals y for [k , k+1 ); and
(e) iterates of Mk +1 , leading ultimately to the iterate Mk+1 +1 and the iteration
embedding jk +1,k+1 +1 .

If any of the iterates indicated in item (e) is illfounded then I wins the game G fixed
by default, regardless of how the game proceeds from here. So let us assume that the
iterates are wellfounded.
Let zk = $y | k < k+1 % Rk . Let Nk+1 = Mk+1 +1 and let ik,k+1 : Nk
fixed (Nk [hk ], ik (k ), k , Ck )
k is a winning strategy for I in G
Nk+1 be jk +1,k+1 +1 . As 
we know that there exists some hk so that:
(1) hk is col(, ik+1 (k ))-generic/Nk+1 [hk ]; and
(2) zk ik,k+1 (Ck )[hk ].
Let Uk+1 = ik,k+1 (Uk+1 ). Let W0k+1 , . . . , W2k+2
k+1
denote the models of Uk+1 ,
and let , denote the embeddings. Note that for each l 2k + 2, Wlk+1 is simply
k+1

ik,k+1 (Wlk ). The same is true of the embeddings. Using the elementarity of ik,k+1 we
k+1
get that W2k+2 agrees with W0k+1 = Nk+1 up to ik+1 () and crit(0,2k+2
k+1
) > ik+1 ().
Let h k+1 = h hk . Let xk+1 = $y | < k+1 % = xk zk . Using condition (2)
k

we see that certainly xk+1 Nk+1 [hk+1 ]. Using the agreement between Nk+1 and
k+1
W2k+2 k+1
it follows that xk+1 W2k+2 [hk+1 ]. This secures the inductive condition (i) for
k + 1.
Let Pk+1 = ik,k+1 (Pk+1 ) and let k+1 = ik,k+1 (k+1 ). Taking condition (2) and
folding in the definition of Ck from condition (iii) we get:
2B Limits 63

k+1
zk (0,2k+2 ik+1 )(Ak )[hk ][hk ](
xk , Pk+1 , k+1 ).
By Definition 2B.7 this simply says that:
k+1
I wins the game (0,2k+2 ik+1 )(Gk+1 )[hk ][hk ](
xk+1 , Pk+1 , k+1 ).
Thus the inductive condition (ii) is secured for k + 1.
Let k+1 = ik,k+1 (k ). Let Pk+1 be the position given by Uk+1 and Pk+1 in the
xk+1 ]. For the record we note that:
pivot game ik+1 (Apiv )[
(v) k+1 extends ik,k+1 (k ) and Pk+1 extends ik,k+1 (Pk ).
k+1
(vi) k+1 is strictly smaller than (2k,2k+2 ik,k+1 )(k ).
Condition (vi) follows from condition (a) above, which in turn traces back to rule (1)
in Definition 2B.4.
These final assignments and observations complete the construction for stage k,
putting us in a position to start stage k + 1. Running through stages [0, ) of the
construction we obtain moves corresponding to mega-rounds [1, ) of G fixed . In par-
ticular we obtain the reals $y | < % which form y, and the models and embeddings
$M , j, | < %. To complete the proof we must play mega-round , and verify
that I wins the resulting run.
Let M be the direct limit of the system $M , j, | < %. This is also the
direct limit of the system $Nk , il,k | l k < %. Let ik, : Nk N = M denote
the direct limit embeddings of this system. Let i = i0, : N0 N . Note that this
is the same embedding as j1, : M1 M .
If N = M is illfounded then I wins by default, regardless of how this run of
fixed continues. So suppose that N is wellfounded.
G  
Let P = k< ik, (Pk ). Let = k< ik, (k ). Both assignments make
sense by condition (v). P is an infinite run of i (Apiv )[ y ], played according to
i (piv )[ , y], see item (D) above.
Remember that one of the initial assumptions in this section says that M& + 1 is
countable in V. Using this assumption we can arrange that:
(vii) : N &i ( ) + 1 is onto.
Arranging (vii) is a simple matter of book-keeping, involving the choice of k+1 ex-
tending k in stage k of the construction. (The construction itself places no restrictions
on this extension.) Let us only emphasize the use of: (a) the initial assumption that
M& + 1 is countable; and (b) the fact that all the iteration trees created during the
construction are countablein fact they have length by rule (ii) in Section 2Aand
use extenders taken from below the image of . It follows from these facts that all the
relevant embeddings preserve countability (see Appendix A), and N &i ( ) + 1 is
therefore countable in V.
Applying Lemma 1C.5, or more precisely Lemma 1C.8, we see that P is a pivot
for y over N . This uses (vii), together with the fact that P is by construction played
according to i (piv )[ , x].
64 2 Games of fixed countable length


Let U = k< ik, (Uk ). U is the length iteration tree played as part of the
pivot P . Let Wl , l < , denote the models of U . Let ,
denote its embeddings.

Note that Wl is equal to ik, (Wl ) for any k < so that l 2k. The same is true for
k

the embeddings.
For each k < let k = ik, (k ). Using condition (vi) we see that:
<
(viii) For each k < , k+1
2k,2k+2 (k ).
It follows that the direct limit along the even branch of U is illfounded.
We are now ready to play the final mega-round of G fixed , mega-round . We play
U , a length iteration tree on M , as Is move. The imaginary opponent, playing for
II, picks b , a cofinal branch through U . Let Nb = M +1 be the direct limit along
b , and let i,b = j, +1 be the direct limit embedding.
This completes our construction, joint with the imaginary opponent, of a run of
fixed . It remains to verify that the run constructed is won by player I.
G
If by chance b is the even branch through U then I wins by default; remember
that the even branch of U leads to an illfounded direct limit by condition (viii). We
may therefore assume that b is odd. Since U is part of an i (A)-pivot for y over
N it follows that there exists some h so that:
(1) h is col(, (i,b i )( ))-generic/Nb ; and
a , y% (i,b i )(A )[h].
(2) $
Here al = ik, (a  = $al | l < %. In
l ) for any sufficiently large k < , and a
other words a = k< ik, (Pk ) is the sequence of auxiliary moves played as part of
the pivot P .
Now Nb is the same as M+1 , and i,b i is the same as j, +1 j1, = j : M
M +1 . Remember also that A simply names the product of (M& ) with A. Condi-
tions (1) and (2) can thus be re-written as:
(1 ) h is col(, j ( ))-generic/M+1 ; and
(2 ) y = $y | < % j (A)[h].
fixed . The run we constructed is therefore
These precisely are the winning conditions in G
won by player I, as required.
Through the construction above we proved the following claim:
Claim 2B.10. Suppose that there exists an ordinal M so that I wins G0 (, , ).
Then I wins the game G fixed (M, , , A).
fixed = G #
As a closing remark let us note that the construction of P above is essentially a
matter of taking the construction in case 1 of the sample application Theorem 1E.1 of
Section 1E, breaking it up into separate stages, and spreading these stages over the
models Nk , k < . More precisely the passage above from Pk to Pk+1 is similar to
round k in the construction in Section 1E, done over Nk and then shifted to Nk+1 . The
end use of P is similar to the use of the pivot in Section 1E, done over N .
2B Limits 65

2B (3) II wins. Recall that B is a name for a subset of R in M col(, ) , that B  names
the product of (M& ) with B, and that B is the mirrored auxiliary game associated
to B  . In Section 2B(1) we developed a condition which, in Section 2B (2), was shown
to imply that I wins G fixed (M, , , A). Here we mirror this development to come up
with a condition which implies that II wins H fixed (M, , , B).

Definition 2B.11 (for k < ). A mirrored k-sequence over M[g k ] is a triplet $


x , Q, %
M[g k ] so that:

is an ordinal;

x belongs to Rk ; and

Q is a position of k rounds in B[


x ].

Definition 2B.12 (for k < ). A mirrored extended k-sequence over M[g k ] is a triplet
x , Q , % M[g k ] so that:
$

is an ordinal;

x belongs to Rk ; and

Q is a position of k + 1 rounds in B[


x ].

These definitions should be compared with Definitions 2B.2 and 2B.3. The only
difference is the change from the auxiliary games map A associated to the name A , to
the mirrored auxiliary games map B associated to the name B  .
As in Section 2B(1) we define games Hk ( x, P , )
x , Q, ) and names Bk [g k ](
by induction on and .

x , Q, ) is played according to Diagram 2.3 and the


Definition 2B.13. The game Hk (
following rules:

(1) (Rule for II) must be an ordinal smaller than .

(2) (Rule for II and I) bkII and bkI must be legal moves in B[
x ] following the
position Q.

I bkI
II bkII

Diagram 2.3. The game Hk (


x , Q, ).

Once the game is concluded, let bk = $bkII , bkI % and Q = Q, bk . Player II wins
x , Q , )].
just in case that (M&k+1 )[g k ] |= II [k , k , Bk [g k ](
66 2 Games of fixed countable length

Definition 2B.13 should be compared with Definition 2B.4. There are the standard
differences: instead of working with a k-sequence we work with a mirrored k-sequence;
x , Q , ),
the payoff now is for player II; the name appearing in the payoff is Bk [g k ](
to be defined below, instead of a name given by A; and the formula used in the payoff
condition is II (which is assumed known for < ) instead of I . We point out one
more difference. The roles of players I and II are reversed here. This applies of course
to the mirrored auxiliary moves bkII and bkI , see Section 1D. But it also applies to
the move . This move is now played by II.

x , Q , ) just in case that


Definition 2B.14. For z Rk M[g k ][gk ] put z Bk (
M[g ][gk ] |=II wins Hk+1 (
k
x z, Q , ).

Definition 2B.14 is the standard mirror image of Definition 2B.7. It determines


x , Q , ), the canonical name for Bk (
Bk [g k ]( x , Q , ), modulo knowledge of the
games Hk+1 ( x z, Q , ). As in Section 2B (1), Definitions 2B.14 and 2B.13 are
carried out simultaneously, by induction on and , inside M.
A mirrored 0-sequence is simply a triplet $, , % for some ordinal M. An
argument which mirrors that in Section 2B(2) gives the following claim, which mirrors
Claim 2B.10:

Claim 2B.15. Suppose that there exists an ordinal M so that II wins H0 (, , ).


fixed (M, , , B).
Then II wins H #

2B (4) The third case. Suppose that there does not exist a M so that I wins
G0 (, , ), and that there does not exist a M so that II wins H0 (, , ). We
shall show that case (3) of Theorem 2A.2 holds. In other words we shall produce some
$y | < % R M[g ] which belongs to neither A[g ] nor B[g ].
Let $L , H % be a pair of local indiscernibles of M relative to . Recall that this
means that L < H and:

(M&L + ) |= [L , c0 , . . . , ck1 ] (M&H + ) |= [H , c0 , . . . , ck1 ]

for any k < , any formula with k + 1 free variables, and any c0 , . . . , ck1
(M& + ).

Claim 2B.16. Let x be an element of Rk M[g k ], and let P be a position of k rounds


in A[ x ]. Then I wins Gk (
x , P , L ) iff I wins Gk (
x , P , H ). (Note the switch from L
to H .)

Proof. This is immediate from the indiscernibility property of L and H , using


Claim 2B.8. #

The following claim is the mirror image of Claim 2B.16:

Claim 2B.17. Let x be an element of Rk M[g k ], and let Q be a position of k rounds


in B[
x ]. Then II wins Hk (
x , Q, L ) iff II wins Hk (
x , Q, H ). #
2B Limits 67

Let gen M[g ] be the generic strategies map associated to A , , and X =


M&. Let gen M[g ] be the mirrored generic strategies map associated to B  , ,
and X = M&.
We work in stages k < to construct:

(A) y = $y | < % R M[g ];

y ], played according to gen [


(B) a , an infinite run of A[ y ]; and
 an infinite run of B[
(C) b, y ], played according to gen [
y ].

The construction takes place inside M[g ]. Note in this respect the fact that gen and
gen belong to M[g ].
 We use xk to denote yk ,
At the start of stage k we shall have yk , a k, and bk.
 We shall maintain:
Pk to denote a k, and Qk to denote bk.

xk , Pk , L ); and
(i) I does not win the game Gk (

(ii) II does not win the game Hk (


xk , Qk , L ).

For k = 0 both conditions (i) and (ii) hold because of the case assumption, that I
does not win G0 (, , ) for any in M, and similarly with II and H0 .
Let us begin stage k of the construction, assuming inductively that the previous
stages were taken care of.
From condition (i) and Claim 2B.16 if follows that I does not win the game
Gk (xk , Pk , H ). This game is finite, hence determined. So II wins. Fix a winning
strategy k for II. We proceed by using this strategy to construct a run of Gk ( xk , Pk , H ).

Play = L as a first move, for I, in Gk ( xk , Pk , H ). The move is legal since
L < H . Gk ( xk , Pk , H ) continues with moves in round k of A[ xk ] following the
position Pk . gen [ xk ] and k together cover these moves, playing akI and akII re-
spectively. This produces ak , and with it Pk+1 . Note that ak is consistent with gen [ xk ],
as required by condition (B).
k is winning for II in Gk ( xk , Pk , H ), so the run constructed above is won by
player II. It follows that:

(iii) (M&k+1 )[g k ]  |= I [k , k , Ak [g k ](


xk , Pk+1 , L )].

Working similarly, but using condition (ii) and Claim 2B.17, we get moves bkII
and bkI so that bk = $bkII , bkI % is consistent with gen [
xk ] and so that:

(iv) (M&k+1 )[g k ]  |= II [k , k , Bk [g k ](


xk , Qk+1 , L )].

By induction Theorem 2A.2 holds for < . In particular it holds for k . Applying
the theorem and using conditions (iii) and (iv) it follows that there exists some zk Rk
xk , Pk+1 , L ) and Bk [g k ][gk ](
M[g k ][gk ] which avoids both Ak [g k ][gk ](  k+1 , L ).
xk , Q
Fix such a z. Let yk + = z for < k . This defines xk+1 = yk+1 . The
construction for stage k is now complete. Condition (i) holds for k + 1 because zk
68 2 Games of fixed countable length

avoids Ak (xk , Pk+1 , L ). Similarly condition (ii) holds for k + 1 because zk avoids
Bk (
xk , Qk+1 , L ).
Going through stages [0, ) we obtain y, a , and b according to conditions (A)(C).
Since everything is done inside M[g ], certainly y R M[g ]. Using condi-
tion (B) and Lemma 1B.2 we see that $ a , y%  A [g ], and hence y  A[g ]. Using
condition (C) and Lemma 1D.2 we see that y  B[g ]. So case (3) of Theorem 2A.2
holds, as required.

2B (5) Summary. Our work so far establishes that at least one of the cases in Theo-
rem 2A.2 holds: if there is a M so that I wins G0 (, , ) then I wins
fixed (M, , , A), see Section 2B(2); if there is a M so that II wins H0 (, , )
G
then II wins H fixed (M, , , B), see Section 2B (3); and otherwise there exists y
R M[g ] which belongs to neither A[g ] nor B[g ], see Section 2B (4).

Our work also shows that M can distinguish between the three cases. The statements
there exists a so that I wins G0 (, , ) and there exists a so that II wins
H0 (, , ) can be phrased inside M, see Claim 2B.8. Moreover the parameters needed
to phrase these statements are definable in M from , A and B respectively, and the
sequence {k }k< fixed at the outset in Section 2B. (Note that the auxiliary games map
A is definable in M from the name A . The definition is given by Section 1A (2).) Let
I ( , , A) be the formula saying that there exists some strictly increasing sequence
{k }k< cofinal in so that for some player I wins the game G0 (, , ) defined using
{k }k< . If M |= I ( , , A) then I wins G fixed (M, , , A). Define II ( , , B)
similarly.
In conclusion: working with a limit ordinal we phrased the two formulae I (, , )
and II (, , ) and proved Theorem 2A.2 for , assuming that the theorem holds for
< and assuming knowledge of the formulae I (, , ) and II (, , ) for < .

2C Successors
Fix a countable 1. We prove Theorem 2A.2 for + 1 assuming that it holds for .
The proof involves breaking G fixed (. . . + 1 . . . ) into two parts, the first dealing with
mega-rounds [1, ] and handled using an inductive application to Theorem 2A.2 for ,
and the second dealing with mega-round + 1 and handled using an adaptation of the
construction in Section 1E. A similar proof shows that Theorem 2A.2 holds for = 1,
and we shall comment on this later.
Fix a ZFC model M with a Woodin cardinal  . Suppose that 1 + + 1 is seen
to be countable in M. Suppose that there are 1 + + 1 Woodin cardinal of M below
 . Let A M and B  M be col(,  ) names for subsets of R +1 . Assume that
M&  + 1 is countable in V. Fix g  which is col(,  )-generic/M. We will prove
Theorem 2A.2 for these objects.
Fix <  , a Woodin cardinal of M, least so that there are 1 + Woodin cardinals
of M below . This is possible since there are 1 + + 1 Woodin cardinals of M
2C Successors 69

below  . Fix g M[g  ] which is col(, )-generic/M. We shall use the fact that
Theorem 2A.2 holds for , applying this theorem with , g, and names to be defined
shortly.
As a matter of notation, we use y for elements of R , and z for elements of R. y, z
is then an element of R+1 . Working in M, fix a continuous injection : R +1 R.
The precise nature of this injection is not important, so long as knowledge of y and zk
suffices to determine ( y , z)k. We think of y, z as coded by the real ( y , z),
and carelessly write y, z instead of ( y , z) throughout.
Let A M be the canonical name for the set (M&  )  A [g  ] in M[g  ].
Similarly let B  M be the canonical name for the set (M&  )  B  [g  ]. Let A
be the auxiliary games map associated to A ,  , and X = M&. Let B be the mirrored
auxiliary games map associated to B  ,  , and X = M&.
We carelessly write A[ y , z] instead of A[(y , z)] and similarly with B. Note
that knowledge of y and zk suffices to determine the rules for the first k rounds of
A[ y , z] and B[ y , z].
Remark 2C.1. The use of the set X = M& for the auxiliary games maps will later
ensure that the pivots produced in the construction reside above (an image of) , see
Lemma 1C.7.
For each y R M[g] let G (
y ) be the game played according to Diagram 2.4
and the rules below:

I z(0) a0I a1I z(2) ...


II a0II z(1) a1II ...

Diagram 2.4. The game G (


y ).

I and II alternate playing natural numbers forming together the real z. In addition
they play moves anI and anII in the auxiliary game A[ y , z]. The first player to
violate these rules loses. If the rules are successfully followed for moves then II wins.
Mirroring the above definition, define for each y R M[g] a game H ( y ) played
according to Diagram 2.5 and the rules below:

I z(0) b0I b1I z(2) ...


II b0II z(1) b1II ...

Diagram 2.5. The game H (


y ).

Again I and II collaborate to create the real z. In addition they play auxiliary moves,
this time in the mirrored auxiliary game B[ y , z]. It is now player I who wins infinite
runs.
The games G ( y ) and H (
y ) are defined modulo knowledge of the maps A and
B, and the sequence y. A and B exist in M. y exists in M[g]. So the games exist in
70 2 Games of fixed countable length

M[g]. Notice that the games are open and closed respectively. So they are determined,
in M[g].
Let A M be the canonical col(, ) name for the set of y R M[g] so that I
wins G (
y ). Let B M be the canonical name for the set of y R M[g] so that II

wins H (y ).
We intend to apply Theorem 2A.2 using , , and the names A and B.
fixed (M, , , A). Then player I wins
Claim 2C.2. Suppose that player I wins G

Gfixed (M, , + 1, A ).
 


Proof. Fix an imaginary opponent willing to play for II in the game G fixed =

Gfixed (M, , + 1, A ). We describe how to play against this imaginary opponent,
 

and win.
Using the assumption of the claim, fix , a winning strategy for player I in
fixed (M, , , A). Pit 
G against the imaginary opponent. This takes care of mega-

rounds [1, ] of Gfixed , creating the reals y for < and the models and embeddings
$M , j, | + 1%. If M+1 , or any of the preceding models, is illfounded
then player I wins G  , regardless of how the run continues. So let us assume that all
fixed
models encountered are wellfounded. Since  fixed (M, , , A)
is winning for I in G
we know that there exists some h so that:

(1) h is col(, j1, +1 ())-generic/M+1 ; and

(2) y j1, +1 (A)[h].

It remains to play mega-round + 1 of G  against the imaginary opponent, and


fixed
win. Mega-round + 1 consists of moves giving rise to a real z = y , an iteration tree
T +1 , and a cofinal branch b+1 through that tree.
Condition (2) and the definition of A together imply that player I, the open player,
wins j1, +1 (G )(y ). Fix witnessing this. Since winning an open game is absolute
we may fix M+1 [h].
Let Apiv be the pivot games map associated to A ,  , and X = M& in M. Let piv
be the pivot strategies map associated to these objects.
Moves given by j1, +1 (piv ), and shifts of that strategy, and the imaginary
opponent combine to produce a real z and a j1,+1 (A)-pivot a , U for x = ( y , z)
over the model M+1 , with the even branch of U illfounded. The construction, which
we omit, is similar to that in case 1 in the proof of Theorem 1E.1. j1,+1 (piv ) takes care
of the production of U and of IIs auxiliary moves in shifts of j1,+1 (G )( y ) along the
even branch of U; the imaginary opponent takes care of IIs part of the real z; and
and its shifts along the even branch of U take care of Is moves in j1,+1 (G ), including
Is part of the real z.
Remark 2C.3. The construction of z, U, and a involves shifting the strategy along
the even branch of U. In Section 1E, was an element of M and could thus be shifted
without concern. Here we do not have M +1 , but only know that is an element
2C Successors 71

of M +1 [h] where h is generic for col(, j1, +1 ()). Fortunately the critical points of
the extenders used in U are higher than j1, +1 (). The embeddings of U can thus be
extended to act on M+1 [h], and can be shifted using the extended embeddings.
The fact that the extenders of U have critical points above j1,+1 () traces back to
Remark 2C.1, and uses the initial settings of  > .
Remark 2C.4. The assumption that M&  + 1 is countable in V is used implicitly in
the construction of z, U, and a . It implies that M +1 &j1,+1 (  ) + 1 is countable in V,
and this is needed for the use of j1, +1 (piv ).
The real z constructed above corresponds to rule (i) of Section 2A, in mega-round
+ 1 of the game G  . Proceeding to rule (ii) we play the tree U constructed above as
fixed
the move T +1 . The imaginary opponent responds with a cofinal branch b +1 through
U. N +2 is set to be the direct limit along b+2 , and j +1, +2 is set to be the direct limit
embedding. This ends the game G  . It remains to verify that the run we constructed
fixed
is won by player I.
If b = b +1 is the even branch of U, then M +2 is illfounded3 and I wins G  by
fixed
default. So assume that b+1 is an odd branch.
Since a , U is a j1, +1 (A)-pivot for x = (y , z) we know that there exists an h
so that:

(1) h is col(, (jb j1, +1 )(  ))-generic/Mb where Mb is the direct limit along b
and jb the direct limit embedding; and

a , x% (jb j1, +1 )(A )[h ].


(2) $

Using the definition of A , and the fact that M +2 = Mb and j +1, +2 = jb , we get:

(1 ) h is col(, j1, +2 (  ))-generic/M+2 ; and

(2 ) y, z j1, +2 (A )[h ].

 .
This precisely is the winning condition for I in G #
fixed

fixed (M, , , B). Then player II wins


Claim 2C.5. Suppose that player II wins H
fixed (M, , + 1, B ).
H  

Proof. Mirror the proof of Claim 2C.2. #

Let I (  , + 1, A ) be the formula expressing I (, , A) holds true in M&  ,


where and A are defined as above. (Note that and A are definable in M from + 1,
 , and A .) Define II (  , + 1, B  ) similarly.
If M |= I [  , + 1, A ] then M&  |= I [, , A]. Using Theorem 2A.2 for it
follows that I wins G fixed (M, , , A),
fixed (M&  , , , A). This game is the same as G
3 The fact that the even branch of U is illfounded has to do with the fact that is a winning strategy for
the open player. See the construction in case 1 of Theorem 1E.1.
72 2 Games of fixed countable length

since both games involve only objects in M at the level of . So Claim 2C.2 applies,
and it follows that I wins G fixed (M,  , + 1, A ).
fixed (M,  , + 1, B  ). The next
Similarly, if M |= II [  , + 1, B  ] then II wins H
claim therefore completes the proof of Theorem 2A.2 for + 1:
Claim 2C.6. Suppose that M  |= I [  , + 1, A ] and M  |= II [  , + 1, B  ]. Then
there is a y R M[g] and a z R M[g  ] so that y, z M[g  ] belongs to
neither A [g  ] nor B  [g  ].
Proof. By assumption both I [, , A] and II [, , B] fail in M&  . Theorem 2A.2
applied in M&  with and the names A and B produces y R M[g] which belongs
to neither A[g] nor B[g]. Using the definition of A and B it follows that II (the closed
player) wins G ( y ) and that I (the closed player again) wins H (
y ). Fix M[g] and

M[g] witnessing this. Let gen M[g ] be the generic strategies map associated


to A ,  , and X = M&. Let gen M[g  ] be the mirrored generic strategies map
associated to B  ,  , and X = M&. Working inside M[g  ] combine , , gen , and
gen to form:
(1) z R;
(2) a , a generic infinite run of A[
y , z]; and
 a generic infinite run of B[
(3) b, y , z].
The construction, which we omit, is similar to that of case 3 in the proof of Theo-
rem 1E.1.
From condition (2) it follows that y, z  A [g  ]. From condition (3) it follows
that y, z  B  [g  ]. #
Claims 2C.2, 2C.5, and 2C.6 complete the proof of Theorem 2A.2 for +1, assuming
that the theorem holds for .

Remark 2C.7. Adapting the above arguments to the case that = 0 one can prove
Theorem 2A.2 for + 1 = 1. R = {} in the case = 0. G ( y ) and H (
y ) thus

degenerate into games G = G () and H = H (). Claim 2C.2 should be revised
for the case = 0 to say that if I wins G then I wins G fixed (M,  , 1, A). Claim 2C.5

should be revised similarly. I ( , 1, A) is the formula which expresses the statement
I wins G . II (  , 1, B) is defined similarly. The proofs of the revised claims for
= 0 are degenerates of the proofs given above. We leave the exact details to the reader.
Remark 2C.7, the work in this section on the successor case, and the work in
Section 2B on the limit case complete the proof of Theorem 2A.2.
Some words are due on our methodology. Let I (, A) denote the formula
(, , A). Define II similarly. The proof of Theorem 2A.2 for used inductive
knowledge of I and II , for < . (This was true for both limit and successor .)
The proof then provided I and II based on that knowledge.
The passage from $I | < %or more precisely from the sequence of Gdel
numbers of these formulaeto the formula I was uniform and definable over any
2D Limits again 73

model of ZFC . This can be seen easily by following the constructions in Sections 2B
and 2C. We can thus obtain a single formula I (, , A) which expresses the statement
I (, A). Similar reasoning allows us to come up with II .

2D Limits again
We have so far seen how to prove Theorem 2A.2 (Sections 2B and 2C), and how to
use Theorem 2A.2 to prove the determinacy of games of countable length ( + 1)
(Theorem 2A.3). Here we handle games of length for limit < 1 . We prove:

Theorem 2D.1. Let be a countable limit ordinal. Suppose that there exists a model
M and a cardinal in M so that:

M is a class model of ZFC;

M is weakly iterable;

M has Woodin cardinals below ; and

M& is countable in V.

Then the games G (C) are determined for all C R in the pointclass <2 11 .

The assumption in Theorem 2D.1 follows from the existence in V of a sharp for
Woodin cardinals, by the methods at the end of Appendix A.
The proof of Theorem 2D.1 takes the rest of this section. We plan to break into
blocks, use Theorem 2A.2 to progress through each of the blocks, and intersperse
between the blocks moves in the game of Martin [24] for witnessing membership in a
<2 11 set.
Let { } < enumerate the Woodin cardinals of M below , in increasing order.
Without loss of generality we may assume that is seen to be countable in M. Fix
in M a strictly increasing sequence of ordinals k , k < , starting with 0 = 0 and
cofinal in . Let k = k+1 k , or more precisely let k be the unique ordinal such
that k+1 = k + k . Let k be  k+1 . The following list encapsulates the properties of
{k }k< , in M, which we intend to use:

each k is a Woodin cardinal of M;

there are 1 + k Woodin cardinals of M below k and (if k > 0) above k1 ;


and

for each k, M&k + 1 is countable in V.

These properties allow us to use Theorem 2A.2 with k , k , and models M[g k ] where
g k is generic for the collapse of 0 , . . . , k1 . (Note that there are still 1 + k Woodin
74 2 Games of fixed countable length

cardinals below k , left after the collapse.) We shall make such applications of Theo-
rem 2A.2 without further comment.
Using Remark 1E.5 every real can be absorbed into a generic extension of an iterate
of M, via a forcing which only collapses the first Woodin cardinal of the iterate, and
therefore leaves plenty of measurable cardinals in the extension. It follows from this
that every real has a sharp. Since M& is countable in V, M& has a sharp. Replacing
M by L(M&) if needed let us assume that:
M = L(M&).
The indiscernibles given by (M&) are then Silver indiscernibles for M.
Fix C R in the pointclass <2 11 . There is then some < 2 and some
sequence $C | < % of 11 sets so that

y C the least such that = or y  C is of the same parity as .

When we say of the same parity as we mean odd if is odd, and even if is even.
Theorem 2D.1 in the case of = 0 is trivial. We may therefore assume that > 0.
Increasing if needed, let us also assume that is odd. The reader who prefers not to
make this assumption should change odd to of parity the same as and even to
of parity opposite to in rules (1)(3) below.
Let m < be such that < 2 belongs to ( m, m + ]. Let u0 < < um
be the first m + 1 uniform Silver indiscernibles.
For each < there is some Lipschitz continuous map y  R [ y ] so that R [
y]
is a linear order of and y C iff R [y ] is wellfounded. Without loss of generality
we may assume that 0 is the largest element in R [ y ], for all and all y. Lipschitz
continuity above is meant with reference to some coding of elements of R as reals.
The precise manner in which this is done is irrelevant, so long as its done in M and so
long as knowledge of yk suffices to determine R [ y ]k + 1.
Let C[ y ] be the following game, taken from Martin [24]: I and II alternate playing
ordinals, embedding the relations R [ y ] into the ordinals.
(1) I is responsible for R [
y ] for all even < , and II is responsible for R [
y ] for
all odd < .
(2) R +1 [
y ] must be embedded into ordinals larger than those used for R [
y ]. (This
is a rule on I if + 1 is even, and on II if + 1 is odd.)
(3) R [
y ] for [ n, n + ) must be embedded into ordinals in [un1 , un ).
(This is a rule on I if is even, and on II if is odd. In the case n = 0 we mean
ordinals below u0 .)
We think of a run of C[ y ] as a sequence of ordinals $i | i < %. I plays 2i and II
plays 2i+1 . is divided into disjoint sets r , and {i }i< r is the embedding of
R [
y ] into the ordinals. The division is arranged in accordance with rule (1).
The division is also arranged so that min(r +1 ) > min(r ) for all . In other words
ordinals corresponding to R +1 [ y ] are only played after the first ordinal corresponding
2D Limits again 75

to R [y ]. This particular arrangement is needed to make sense of rule (2). Note that
the first ordinal corresponding to R [ y ] is the largest among all ordinals corresponding
to R [
y ], since 0 is the largest element in R [ y ].
Observe how rules (2) and (3) imply that R [ y ] is embedded into ordinals below
those used for R [ y ] whenever < .
We refer the reader to Martin [24] for an exact description of the rules. An auxiliary
game of this kind was used by Martin to prove the determinacy of length games with
<2 11 payoff.
Observe that y  C[ y ] is Lipschitz continuous. We refer to this map as C, and
note that knowledge of yk suffices to determine at least the first k + 1 rounds of C[ y ].
The parameters used in the definition of C are the sequences $R | < % and
$un | n m%. Both sequences belong to M; the first because C is lightface <2 11 ,
and the second because it is finite. The Lipschitz continuous map C therefore belongs
to M. We shall use it as we used A and B in Section 2B.
A position Q = $0 , . . . , l1 % in C[
y ] is good (over M) if the ordinals played by
I, namely j for j < l even, are Silver indiscernibles for M; and the ordinals played
by II, namely i for i < l odd, are definable in M from the ordinals played by I and
additional parameters in M& {u0 , . . . , um }.
A good position Q = $0 , . . . , l1
 % is called an M-shift of the good position Q if

for each odd i < l, i is definable from M& {u0 , . . . , um } and {j | j < l even} in


the same way that i is definable from M& {u0 , . . . , um } and {j | j < l even}.
The following facts, and their mirror images listed in Section 2D (2), distill the
properties of C[
y ] which we shall need:

Fact 2D.2. Suppose that Q is a good position of even length in C[ y ]. (It is Is turn to
play following Q.) Then there is a shift Q of Q, and a Silver indiscernible for M, so
y ] following Q .
that is a legal move for I in C[

Fact 2D.3. Suppose that $Qk | k < % is a sequence of good positions in C[ y ]. Suppose
that for each k < , Qk+1 extends a shift of Qk . Suppose further that there is no (single)
infinite run Q which extends a shift of each Qk , k < . Then y  C.

The first fact says that it is always possible to ascribe indiscernibles as moves for I;
by shifting previous moves its possible to make room for the next one. The second fact
says that if II can keep up with these indiscernible moves for I, then y does not belong
to C.
The reader who does not know how to prove the second fact should consult Martin
[24]. We comment only that the proof has to do with the specific order imposed by
rules (2) and (3). The assumption in Fact 2D.3 that there is no Q which extends a
shift of each Qk is needed to make sure that the least so that = or y  C is not
equal to .

2D (1) II wins. For expository simplicity fix a sequence $gk | k < % so that each gk
is col(, k )-generic/M[g0 gk1 ]. Let g k denote g0 gk1 . We work in
76 2 Games of fixed countable length

the forcing extensions M[g k ] and M[g k ][gk ] though strictly speaking we should work
in M and use the forcing languages.
We shall follow roughly the outline of definitions in Section 2B, only using C instead
of A and B. It is convenient here to start with the case that II wins. So we begin by
imitating the definitions in Section 2B(3).
Definitions 2B.11 and 2B.12 can be copied verbatim, simply changing B to C. We
make only a notational comment: a position Q of k rounds in C[ x ] now consists of
ordinals $0 , 1 , . . . , 2k1 %.
Having defined mirrored k-sequences and mirrored extended k-sequences, let us
proceed to the definition of the names Bk [g k ]( x , Q , ) and the games Hk ( x , Q, ).
The definition, as usual, is by induction on and . Definition 2B.14, the definition
x , Q , ), can be copied verbatim. The definition of Hk (
of Bk ( x , Q, ) is modified
as follows, to use C[ x ] rather than the auxiliary games map of Section 2B.
x , Q, ) is played according to Diagram 2.6 and the
Definition 2D.4. The game Hk (
following rules:
(1) (Rule for II) must be an ordinal smaller than .
(2) (Rule for I and II) 2k and 2k+1 must be legal moves in C[
x ] following the
position Q.
I 2k
II 2k+1

Diagram 2.6. The game Hk (


x , Q, ).

Once the game is concluded, let Q = Q, 2k , 2k+1 . Player II wins just in case
x , Q , )].
that (M&k+1 )[g k ] |= II [k , k , Bk [g k ](
Recall that the map C is Lipschitz continuous, and knowledge of x Rk suffices
y ] (for any y R extending x). Rule (2) in
to determine the first k + 1 rounds of C[
Definition 2D.4 therefore makes sense.
Claim 2D.5. Suppose that there exists an ordinal M so that II wins H0 (, , ).
Then II wins G (C).
Proof. Fix an imaginary opponent playing for I in G (C). Fix a weak iteration
strategy  for M. We describe how to play against the imaginary opponent, using ,
strategies in various games H fixed (. . . k . . .), and indiscernible moves in shifts of C.
As usual the description takes the form of a construction of (among other things) a run
y = $y | < % of G (C).
We construct by induction on k < . At the start of stage k [0, ) we have:
(A) The reals y for < k . We use xk to denote $y | < k %.
(B) A -iterate Nk of M = N0 , and an iteration embedding ik : M Nk .
2D Limits again 77

(C) Some hk = h0 hk1 which is col(, i1 (0 )) col(, ik (k1 ))-


generic/Nk .
xk ].
(D) A position Qk of k rounds in ik (C)[
(E) An ordinal k in Nk .
The iteration map ik will be the result of a weak iteration of M, of countable length,
using length iteration trees, with extenders taken from images of M&k which is
countable in V. It is easy to see that iteration maps of this kind (when their final model
is wellfounded) do not move the uniform Silver indiscernibles. Our notation henceforth
will use this fact without further comment: we write u0 rather than ik (u0 ), etc.
During the construction we shall make sure that:
(i) xk belongs to Nk [hk ].
(ii) Nk [hk ] |=II wins ik (Hk )[hk ](
xk , Qk , k ).
xk ], where good is meant over Nk .
(iii) Qk is a good position in ik (C)[
(iv) k is definable in Nk from Qk and additional parameters in Nk &ik ()
{u0 , . . . , um }.
We informally use tk (Qk ) to refer to the definition of k given by condition (iv).
If Q is a shift of Qk we write tk (Q ) for the result of interpreting tk using Q instead
of Qk .
The objects needed for the start of stage 0 are for the most part given trivially. The
assumption of Claim 2D.5 states that there is some M so that II wins H0 (, , ).
Let 0 be the least such. H0 is definable in M from {u0 , . . . , um } and a real parameter.
(Note that these objects suffice to define C.) The least so that II wins H0 (, , ) is
thus definable in N0 = M using parameters allowed in condition (iv).
We note before we begin that the sequences $Hk | k < % and $Bk | k < % are
definable in M from {u0 , . . . , um } and a real parameter. (The real parameter codes,
among other things, the sequence {k }k< .)
Let us start stage k of the construction. Let Qk be the shift of Qk given by Fact 2D.2,
and let be the Silver indiscernible given by that fact. Let k = tk (Qk ). Since Qk is
a shift of Qk , there is an elementary embedding : Nk [hk ] Nk [hk ] which fixes the
uniform indiscernibles and sends Qk to Qk . By necessity sends k to k . Hk , being
definable from uniform indiscernibles and a real, is not moved by . Applying to
xk , Qk , k ). Let k Nk [hk ]
condition (ii) it follows that II wins the game ik (Hk )[hk ](
witness this. We may assume that k is definable in Nk [h ] from {u0 , . . . , um }, Qk , and
k

parameters from (Nk &ik ())[hk ]. (Note that xk belongs to (Nk &ik ())[hk ], and that k
is definable from the parameters listed. It follows that the game ik (Hk )[hk ]( xk , Qk , k )
is definable from the parameters listed.)
Let k+1 be the first move played by k . Play the Silver indiscernible as the next
move, 2k , for I. Let 2k+1 be the response played by k .
Let Qk+1 = Qk , 2k , 2k+1 . Since k is a winning strategy for player II we get:
78 2 Games of fixed countable length

(v) (Nk &ik (k+1 ))[hk ] |= II [ks , k , Bks [hk ](


xk , Qk+1 , k+1 )], where ks = ik (k )
and Bks = ik (Bk ).

By our choice of a definable k , both k+1 and 2k+1 are definable in Nk from
parameters in Nk &ik () {u0 , . . . , um }, Qk , and . (Initially we get definability in
Nk [hk ] using xk as a parameter. But we can pass to Nk and use forcing conditions,
which belong to Nk &ik (), as parameters.) Thus:

xk ].
(vi) Qk+1 is a good position in ik (C)[

(vii) k+1 is definable in Nk using parameters in Nk &ik () {u0 , . . . , um } and Qk+1 .

For the record let us also note that:

(viii) k+1 < tk (Qk+1 2k).

This follows from rule (1) in Definition 2D.4. (Note that Qk+1 2k = Qk .)
Applying Theorem 2A.2 to condition (v) we see that player II wins the game
H fixed (Nk [hk ], s , k , B s [hk ](
xk , Qk+1 , k+1 )), where Nk = (Nk &ik (k+1 )). Notice
k k
that this game is precisely the same as H fixed (Nk [hk ], . . . . . . ), since Nk is a rank
initial segment of Nk taken above all relevant objects. So player II wins the game
H fixed (Nk [hk ], s , k , B s [hk ](xk , Qk+1 , k+1 )). Fix  k witnessing this.
k k

k , the imaginary opponent, and  combine to produce a complete run of
H fixed (Nk [hk ], s , k , B s [hk ](xk , Qk+1 , k+1 )), including reals $y | [k , k+1 )%
k k
and a weak iteration of Nk of length 1+k +1. Let Nk+1 be the final model of this weak
iteration and let ik,k+1 : Nk Nk+1 be the iteration embedding. Let ik+1 = ik,k+1 ik .
Remark 2D.6. Strictly speaking moves in H fixed (Nk [hk ], . . . . . . ) involve iteration trees
on Nk [hk ]. But we can view them as trees on Nk (see Appendix A).
The use of the iteration strategy  guarantees that Nk+1 is wellfounded. Let zk =
k is winning for II, there exists some hk so that:
$y | [k , k+1 )%. Since 

(1) hk is col(, ik,k+1 (ks ))-generic/Nk+1 [hk ]; and

(2) zk ik,k+1 (Bks )[hk ][hk ](


xk , ik,k+1 (Qk+1 ), ik,k+1 (k+1 )).

Let xk+1 = xk zk , let Qk+1 = ik,k+1 (Qk+1 ), and let k+1 = ik,k+1 (k+1 ).
Conditions (i)(iv) for k + 1 are now easy to verify. We only point out that condition
(ii) for k + 1 follows from condition (2) and that condition (iv) for k + 1 follows from
condition (vii) above. This completes the construction in stage k, putting us in a position
to start stage k + 1.
Inductively the construction is now complete. For future record let us note that:

(ix) Qk+1 extends a shift of ik,k+1 (Qk ).

(x) tk+1 (Qk+1 ) < ik,k+1 (tk )(Qk+1 2k), where both tk+1 and ik,k+1 (tk ) here are
interpreted inside Nk+1 .
2D Limits again 79

I 2k
II 2k+1

Diagram 2.7. The game Gk (


x , P , ).

The last condition follows from condition (viii), and ultimately traces back to rule (1)
in Definition 2D.4.
It remains to verify that $y | < % is won by player II in G (C).
Let N be the direct limit of the models Nk . Let ik, : Nk N be the direct
limit embeddings. Let i denote i0, . The use of the iteration strategy  during the
construction guarantees that N is wellfounded. Let Qk = ik, (Qk ). Since i (C)[ xk ]
y ], Qk is a position in i (C)[
consists precisely of the first k +1 rounds of i (C)[ y ]. By
condition (iii) each Qk is good, where this is meant over N . Moreover by condition
(ix), Qk+1 extends a shift of Qk for each k. The following claim will allow us to apply
Fact 2D.3:
Claim 2D.7. There is no Q which extends a shift of each Qk .

Proof. Suppose otherwise. Let k = ik, (tk )(Q 2k), where the definition ik, (tk )

is interpreted in N . Using condition (x) we get k+1 < k for each k. But this
contradicts the fact that N is wellfounded. #

Applying Fact 2D.3 we finally conclude that y  C. Playing for II against some
imaginary opponent we produced a run y = $y | < % of G (C). We then went on
to verify that y  C so that this run is won by II. This completes the proof of Claim 2D.5.
#

2D (2) I wins. We work here to mirror the argument of Section 2D (1). The process
of mirroring is similar to that in Section 2B, except that we keep using the same map C
for the mirrored argument.
Let us start with the basic definitions. The definitions of k-sequences and extended
k-sequences are verbatim copies of Definitions 2B.2 and 2B.3 in Section 2B (1), only
changing A to C. Definition 2B.7 can also be copied verbatim. Definition 2B.4 should
be modified to the following:

x , P , ) is played according to Diagram 2.7 and the


Definition 2D.8. The game Gk (
following rules:

(1) (Rule for II) must be an ordinal smaller than .

(2) (Rule for I and II) 2k and 2k+1 must be legal moves in C[
x ] following the
position P .

Once the game is concluded let P = P , 2k , 2k+1 . Player I wins just in case that
x , P , )].
(M&k+1 )[g k ] |= I [k , k , Ak [g k ](
80 2 Games of fixed countable length

Note that is played by II, not by I. This is the one essential difference between
the definition here and Definition 2B.4.
Let $L , H % be the least pair of local indiscernibles of M relative to um . Note that
L is definable in M from um .

Claim 2D.9. Suppose that I wins G0 (, , L ). Then I wins G (C).

Proof sketch. For the purpose of this argument, a position P in C[ y ] is called good
if the ordinals played by II are indiscernibles for M, and the ordinals played by I are
definable from the ordinals played by II. Note how this mirrors the earlier definition of
a good position Q. Mirror the definition of a shift similarly. The following facts are
adapted from Martin [24]:
Fact 2D.10. Suppose that P is a good position of odd length in C[ y ]. (It is IIs turn to
play following P .) Then there is a shift P  of P , and a Silver indiscernible for M, so
y ] following P  .
that is a legal move for II in C[
Fact 2D.11. Suppose that $Pk | k < % is a sequence of good positions in C[ y ].
Suppose that for each k < , Pk+1 extends a shift of Pk . Then y belongs to C.
Fact 2D.10 states that it is always possible to ascribe indiscernibles as moves for II.
Fact 2D.11 states that if I can keep up with these indiscernible moves, then y C.
These facts mirror the earlier facts, 2D.2 and 2D.3. But notice that the condition of
Fact 2D.3 that $Qk | k < % cannot be combined to form an infinite run Q is not
mirrored into the current Fact 2D.11. That condition was needed earlier to make sure
that the least so that = y  C is not equal to . Here there is no need for such
an assumption; if the least so that = y  C is equal to , then y belongs to C.
The proof of the current Claim 2D.9 mirrors the earlier proof of Claim 2D.5, using
Facts 2D.10 and 2D.11 instead of Facts 2D.2 and 2D.3. We omit the details and only
make the following observations:
The strategy k (obtained during the proof) is now a strategy for I. Since is still
played by II it is our responsibility during the construction to come up with k+1 ; it is
not played for us by k .
We play k+1 ourselves using the usual trick of switching between local indis-
cernibles, starting with L . Conditions (viii) and (x) must therefore be relinquished;
as we switch between local indiscernibles we certainly do not obtain an infinite de-
creasing sequence. As a result we cannot mirror Claim 2D.7. But we dont need to.
Fact 2D.11unlike the earlier Fact 2D.3can be applied knowing only that for each
k < , Pk+1 extends a shift of P . #
k

2D (3) Determinacy. Remember that our goal is to prove the determinacy of G (C).
Given Claims 2D.5 and 2D.9 it is enough to prove:

Claim 2D.12. Either there exists some so that II wins H0 (, , ), or else I wins
G0 (, , L ).
2D Limits again 81

Proof. Assume for contradiction that I does not win G0 (, , L ), and that for all , II
does not win H0 (, , ). In particular II does not win H0 (, , L ).
Recall that for each k < , gk is col(, k )-generic/M[g k ] where g k = g0
gk1 . We work by induction on k < to construct:
(A) xk Rk M[g k ];
(B) Pk , a position of k rounds in C[
xk ]; and
(C) an ordinal k .
We shall make sure that xk+1 extends xk ; that Pk+1 extends Pk ; and most importantly
that k+1 is smaller than k . This last clause will yield the desired contradiction. As
we construct we shall maintain:
xk , Pk , k ); and
(i) I does not win Gk (
(ii) II does not win Hk (
xk , Pk , k ).
We start with x0 = , P0 = , and 0 = L . Conditions (i) and (ii) hold by the
initial assumption for contradiction.
Suppose xk , Pk , and k have been constructed. Suppose that conditions (i) and (ii)
hold at k. Since Gk ( xk , Qk , k ) is a finite game, condition (i) implies that it is won by
II. Fix a strategy k witnessing this. Using condition (ii) similarly, fix a strategy k ,
winning for I in Hk ( xk , Pk , k ).
Let k and k play against each other. (Note that the games Hk ( xk , Pk , k ) and
Gk (xk , Pk , k ) have identical moves, see Diagrams 2.6 and 2.7.) This produces k+1 ,
2k , and 2k+1 . The rules of Gk ( xk , Pk , k ) are such that:
(iii) k+1 < k .
Let Pk+1 = Pk , 2k , 2k+1 . Since k is winning for II in Gk (
xk , Pk , k ), we get:
(M&k+1 )[g k ]  |= I [k , k , Ak [g k ](
xk , Pk+1 , k+1 )].
Since k is winning for I in Hk (
xk , Pk , k ), we get:
(M&k+1 )[g k ]  |= II [k , k , Bk [g k ](
xk , Pk+1 , k+1 )].
We may therefore apply Theorem 2A.2 and obtain zk Rk M[g k ][gk ] which belongs
to neither Ak [g k ][gk ](
xk , Pk+1 , k+1 ) nor Bk [g k ][gk ](
xk , Pk+1 , k+1 ). Using the def-
initions of these two sets (see Definitions 2B.7 and 2B.14) it follows that I does not
win Gk+1 ( xk zk , Pk+1 , k+1 ) and that II does not win Hk+1 ( xk zk , Pk+1 , k+1 ).
Setting xk+1 = xk zk completes the construction at stage k, securing conditions (i)
and (ii) for k + 1.
Inductively the construction is now complete. Condition (iii) implies that {k }k<
is an infinite descending sequence of ordinals, a contradiction. #
Claims 2D.5, 2D.9, and 2D.12 together imply that G (C) is determined. This
completes the proof of Theorem 2D.1.
82 2 Games of fixed countable length

2E Universally Baire sets


By a tree on a set Y we mean a subset of Y < , closed under initial segments. An infinite
branch through T is a sequence y Y so that yn T for all n < . [T ] denotes the
set of infinite branches through T .
We make the natural identification between U W and (U W ) : Given u U
and w W we use $u, w% to denote both the actual pair $u, w% U W , and the
sequence $$u(i), w(i)% | i < % (U W ) . We make a similar identification for
products of more than two spaces.
Let T be a tree on X U . Define p[T ] X , the projection of T to X , by setting
x p[T ] iff there exists some u U so that $x, u% [T ]. Notice that p[T ] is not
absolute between models of set theory. The following exercises touch on this issue of
absoluteness.
Exercise 2E.1. Let M and M[g] be (wellfounded) models of ZFC . Suppose that
M M[g]. Let S and S be trees in M, on U and U respectively. Suppose
that M |=p[S] p[S ] = . Prove that p[S] p[S ] = also in M[g].
Hint. Look at the tree of attempts to construct $x, u, u % so that $x, u% [S] and
$x, u % [S ], and use the absoluteness of being wellfounded. #
Exercise 2E.2. Give an example of a generic extension V[g] of V, and trees T and T
in V, on 1 and respectively, so that p[T ] p[T ] = holds in V but
not in V[g].
Let X be countable, so that X is a Polish space. A set C X is -universally
Baire if all its continuous preimages, to topological spaces with regular open bases
of cardinality , have the property of Baire. C is -universally Baire if it is -
universally Baire for all cardinals . FengMagidorWoodin [5] provides the following
convenient characterization of universally Baire sets:
Definition 2E.3. A pair of trees T and T on X U and X U respectively is
exhaustive for a poset P if the statement p[T ] p[T ] = X is forced to hold in all
generic extensions of V by P.
Theorem 2E.4 (FengMagidorWoodin [5]). Let X be countable, let C X , and let
be an infinite cardinal. C is -universally Baire iff there are trees T and T so that:
(1) p[T ] = C and p[T ] = X C; and
(2) the pair $T , T % is exhaustive for all posets of size . #
We work with this characterization, rather than the definition. Our aim is first to
prove determinacy for games with -universally Baire payoff for sufficiently large ,
and second to propagate the property of being universally Baire along the hierarchy of
definability.
We work below with substructures of initial segments of the universe. By rank
initial segment of V we always mean a rank initial segment which satisfies enough of
ZFC for the relevant application.
2E Universally Baire sets 83

Exercise 2E.5. Let C be -universally Baire, and let T , T witness this. Let V
be a rank initial segment of V, large enough that T , T , and belong to V . Let N be
a countable model which embeds elementarily into V , say by . Suppose T , T , and
belong to the range of . Let T , T , and be such that (T ) = T , (T ) = T , and
() = .
Let h V be col(, )-generic over N . Let x belong to N[h]. Prove that
N[h] |=x p[T ] if and only if x C.

The next exercise obtains determinacy for universally Baire sets. It reduces the large
cardinal assumption in FengMagidorWoodin [5, Theorem 5.4], where determinacy
is obtained using two Woodin cardinals. The exercise involves a modification of the
proof of Theorem 1E.1. Combined with Fact 2E.7 below it basically leads back to the
determinacy proved in Section 1E.

Exercise 2E.6. Let be a Woodin cardinal (in V). Let C be -universally Baire.
Prove that G (C) is determined.

Hint. Let T and T witness that C is -universally Baire. Let V be a rank initial
segment of V, large enough that T , T , and belong to V . Let H be a countable
elementary substructure of V , with T , T , H . Let M be the transitive collapse of
H and let : M H be the anti-collapse embedding. Let S, S , and be such that
(S) = T , (S ) = T , and () = .
Let g be col(, )-generic over M. Let A M be the canonical name for the
set (M&) p[S] as computed in M[g]. Let B M be the canonical name for
(M&) p[S ].
Imitate the proof of Theorem 1E.1 using these names. In case 3 use the elementarity
of to see that $S, S % is exhaustive for col(, ) over M. In case 1 pick the branch
b using Theorem 12 in Appendix A, so that you get an embedding : Mb V with
jb = . (A branch b for which such an embedding exists is called realizable.)
Then replace the use of 12 absoluteness with a use of Exercise 2E.5. #

We pass now to the matter of propagating the property of being universally Baire.
The starting point is the following fact:

Fact 2E.7 (FengMagidorWoodin [5, Theorem 3.4]). Let X be a countable set, let
be an infinite cardinal, and let C X be 12 . Suppose that (V+1 ) exists. Then C
is -universally Baire.

Given C X and x X define Cx to be the set {y | $x, y% C}.


Define the length game quantifier  , acting on subsets of X , by setting
x  C iff player I has a winning strategy in the game G (Cx ).
Suppose now that C X is -universally Baire, and let T and T witness
this. For x X let Tx and Tx be the trees {$t, u% | $x lh(t), t, u% T } and {$t, u % |
$x lh(t), t, u % T } respectively. It is easy to check that Tx and Tx witness that Cx
is -universally Baire. In particular the pair $Tx , Tx % is exhaustive for col(, ). This
last fact is true not only for x V, but also for x in small generic extensions of V.
84 2 Games of fixed countable length

The following exercise makes this statement (or rather the equivalent statement over a
countable substructure) precise.
Exercise 2E.8. Let V be a rank initial segment of V, large enough that T , T , and
belong to V . Let H be a countable elementary substructure of V , with T , T , H .
Let P be the transitive collapse of H and let : P H be the anti-collapse embedding.
Let S, S , and be such that (S) = T , (S ) = T , and () = . Suppose for
simplicity that X H and (X) = X. (This is certainly true if X = .)
Let M = P [h] be a generic extension of P by a poset of size strictly less than . Let
x belong to X M. Prove that the pair $Sx , Sx % is exhaustive for col(, ) over M.
We now have everything set up for propagating the property of being universally
Baire. The next exercise phrases the propagation precisely.
Exercise 2E.9. Let be a Woodin cardinal (in V). Let X be a countable set. Let
C X be -universally Baire. Prove that  C is -universally Baire for all
< .
Hint. Since X is countable you may assume that X = . Let T and T witness that C
is -universally Baire. Let V be a rank initial segment of V, large enough that T , T ,
and belong to V .
Let Z be the set of tuples $P , , S, S , , h% so that P is a countable transitive model,
is an elementary embedding of P into V , (S) = T , (S ) = T , () = , and h
is generic over P for a poset of size less than .
Let $P , , S, S , , h% belong to Z. Let M = P [h]. Notice that is a Woodin
cardinal in M, since M is a generic extension of P by a poset of size less than . Given
x M let Ax be the canonical name, in the forcing col(, ) over M, for the set
(M&) p[Sx ]. Let G be defined as in the proof of Theorem 1E.1, using the name
Ax over M. Let (x, S, ) be the statement that player I wins this game G . (This is a
statement made over M, using x, S, and as parameters.)
Show that if there is a tuple $P , , S, S , , h% Z so that P [h] |= [x, S, ], then
x belongs to  C. You will essentially imitate your work on case 1 in Exercise 2E.6.
Show that if there is a tuple $P , , S, S , , h% Z so that P [h] |= [x, S, ],
then x belongs to  C. You will essentially imitate your work on cases 2 and 3
in Exercise 2E.6. You can use Exercise 2E.8 to exclude case 3.
The tuples in Z can be coded by countable sequences. Using this fact, define trees
R and R so that:
(1) x p[R] iff there is $P , , S, S , , h% Z so that P [h] |= [x, S, ]; and
(2) x p[R ] iff there is $P , , S, S , , h% Z so that P [h] |= [x, S, ].
Your work above shows that p[R]  C and p[R ]  C. It remains to
show that the pair $R, R % is exhaustive over V for all posets of size strictly less than
. Here you will at last use the way Z allows not just countable substructures of V ,
but also their small generic extensions. Your precise definition of the trees R and R is
also important here. The argument works with the natural definition of these trees, but
there are plenty of unnatural definitions with which the argument would fail. #
2E Universally Baire sets 85

It is easy to check that { C | C 1n } is precisely the pointclass  1n+1 . The


following is therefore immediate from Fact 2E.7 and repeated applications of Exer-
cise 2E.9:
Corollary 2E.10. Let n > 0. Let be an infinite cardinal. Suppose that there are n
Woodin cardinals 1 < 2 < n above , and suppose that (Vn +1 ) exists. Then
all 1n+2 sets are -universally Baire. #
Let us now pass to games of arbitrary fixed countable length.
Exercise 2E.11. Let 1 be a countable ordinal. Let be a Woodin cardinal and
suppose that there are 1+ Woodin cardinals below . (Altogether then V is assumed
to have 1 + + 1 Woodin cardinals.) Let C R be -universally Baire. Prove that
G (C) is determined.
Hint. Let T and T witness that C is -universally Baire. Let V be a rank initial
segment of V large enough that T , T , V . Let H be a countable elementary
substructure of V , with T , T , H . Let M be the transitive collapse of H and let
: M H be the anti-collapse embedding. Let S, S , and be such that (S) = T ,
(S ) = T , and ( ) = .
Let g be col(, )-generic over M. Let A M be the canonical name for p[S]
as computed in M[g]. Let B M be the canonical name for p[S ].
Imitate the proof of Theorem 2A.3 using these names. As in Exercise 2E.6 you
should use Theorem 12 in Appendix A to choose branches through the iteration trees
that come up during the construction, so that all your branches are realizable. In case 1
you will then reach an iteration embedding j1,+1 : M M+1 , and an embedding
: M+1 V so that j1,+1 = . As in Exercise 2E.6, use Exercise 2E.5 to
argue that the run you constructed belongs to C. #
Let 1 be countable. Given C X R and x X define Cx to be the
set {
y R | $x, y% C}. Define the length game quantifier  , acting on
subsets of X R , by setting x  C iff player I has a winning strategy in the
game G (Cx ).
Exercise 2E.12. Let 1 be a countable ordinal. Let be a Woodin cardinal. Let
< , and suppose that there are 1+ Woodin cardinals between and . (Altogether
then V is assumed to have 1 + + 1 Woodin cardinals above .) Let X be countable.
Let C X R be -universally Baire. Prove that  C is -universally Baire.
Hint. Imitate your solution to Exercise 2E.9, using the formula I of Theorem 2A.2
and appealing to your determinacy proof in the last exercise. #
Remark 2E.13. From the last exercise it follows that if there is a class of Woodin
cardinals then the pointclass of -universally Baire sets is closed under applications
of  for each countable . This was proved previously by Woodin, using stationary
tower forcing. The stationary tower proofs do not require iterability, and this becomes an
advantage when one deals with stronger large cardinal axioms, beyond the level where
86 2 Games of fixed countable length

iterability is known. For example Woodin showed, using stationary tower forcing, that
if there is a class of measurable Woodin cardinals then the pointclass of -universally
Baire sets is closed under applications of the quantifier associated to open games of
length 1 . It should be noted that determinacy for these games is not yet known
from large cardinals in V. Assuming full iterability, through realizable branches, for
countable substructures of rank initial segments of V, this determinacy would follow
from measurable Woodin cardinals by the results in Neeman [33]. But the necessary
iterability is open.
Our determinacy proofs can be applied not only to games where the payoff sets are
universally Baire, but also to games where the payoff sets are homogeneously Suslin
(see MartinSteel [18, 2] for the definition). We shall give a concrete example of this
in the context of games of continuously coded length, in Exercise 3E.6. Let us here
only say that the argument in Section 2D can be adapted to deal with homogeneously
Suslin payoff instead of <2 11 payoff, producing the following result:
Exercise 2E.14. Let be a countable limit ordinal. Let be an infinite cardinal.
Suppose there are Woodin cardinals below . Let C R be -homogeneously
Suslin. Then G (C) is determined. #

Hint. Adapt the proof of Theorem 2D.1 to replace Martins games C[
y ] with games on
the tree witnessing that C is homogeneously Suslin. For more details see the hint to
Exercise 3E.6. #

For limit ordinals = with additively closed, the determinacy in Exer-


cise 2E.14 is due to Woodin, obtained as a consequence of his work on equiconsistency
results connecting Woodin cardinals under choice with measures on [P1 (R)]
under AD.
Chapter 3
Games of continuously coded length

Given a set C R<1 and a partial function : R , the game Gcont (, C) is played
as follows: In mega-round the two players collaborate as usual to produce a real y .
If (y ) is undefined, the game ends. Player I wins iff $y | % C. If (y ) is
defined, set n to be its value. If n {n | < } then again the game ends. Again I
wins iff $y | % C. Otherwise the game continues.

I ......... y (0) y (2) .........


II y (1) y (3) ...

Diagram 3.1. The game Gcont (, C).

Let $y | % be a position in Gcont (, C). By necessity  n , < ,


is injective; else the game would have ended before reaching mega-round . is
therefore countable. For limit the injection  n is produced continuously as the
game approaches . Gcont (, C) is said to have continuously coded length in light of
this fact. The notion traces back to Steel [41].
We develop here methods for proving the determinacy of games of continuously
coded length. For example we show that:
Theorem (3D.1). Suppose that there exists an iterable class model M with cardinals
< so that (a) is a Woodin cardinal of M; (b) is strong to + 1 in M (see
Section 3B); and (c) M& + 1 is countable in V. Then the games Gcont (, C) are
determined for all in the class  02 and all C in the pointclass <2 11 .
Other results of similar flavor are listed later in the chapter. The methods developed
here, apart from yielding the determinacy of games of continuously coded length, also
serve as a foundation for some of the work in the next chapters.

3A Codes
We begin with a way to code positions in Gcont by reals. The coding is continuous at
limits, in a manner made precise in Claim 3A.5.
Fix a partial map : R . Let $y | < % be a countable sequence of reals.
Let n = (y ) for < . $y | < % is said ro be a -position (a position for short)
if the n -s are distinct.
Claim 3A.1. Suppose that $y | < % is a position of limit length . Then n
as .
88 3 Games of continuously coded length

Proof. Immediate using the fact that the n -s are distinct. #

Given a position y = $y | < % let o( y ) be the partial linear order on defined
by o(y ) = {$n , n % | < < }. Let d( y ) = {n | < } . d( y ) is the domain
of the order o(y ). Its clear that the order type of o( y ) is . Moreover for any <
the order type of n in o( y ) is precisely .
Let R be the Polish space consisting of R and an extra isolated point which we
think of as standing for undefined. Define z( y ) (R ) by:

y if n = n ;
z(y )n =
if n  d( y ).

Fix some recursive map , injecting (R ) {partial linear orders on } into R. We
use y | < , or  y , to denote (z( y )). We refer to y | <  as the
y ), o(
real coding $y | < %. We say that x is a code if it is a real coding some position
y = $y | < %. Note that y can be recovered from $z( y )%. The position y can
y ), o(
thus be recovered from the real coding it.
The precise specifications of are not important, so long as satisfies the following
property:

Property 3A.2. For any position y and any natural number l, 


y l depends only on
o( y )0 l, . . . , z(
y )(l l) and on $z( y )l1 l%.

Using Property 3A.2 one can easily prove the following claim:

Claim 3A.3. Let $y | < % be a position of length . Let $y | < + 1% be


a position of length + 1 which extends $y | < %. Then y | <  and
y | < + 1 agree to n . #

In fact one can prove:

Claim 3A.4. For any position $y | < % and any < , if n n for all [, ),
then y | <  and y | <  agree to n. #

Combining this with Claim 3A.1 immediately yields:

Claim 3A.5. Let $y | < % be a position of limit length . For each let
x = y | < . Then x x as . #

A -run (a run for short) is a sequence y = $y | < + 1% so that y is a


position; y is a real; and either (y ) is not defined, or else (y ) {(y ) | < }.
A -run is thus a terminal stage in the game of continuously coded length associated
to .
Fix C R<1 . Fix a pointclass . We say that C is  in the codes if there exists
some C R R in  so that for any run y = $y | < + 1%, y C
y , y % C . We say that C belongs to the pointclass  if it is  in the codes.
$
3B First determinacy result, part I 89

3B First determinacy result, part I


Work with a fixed : R , and assume that is continuous. Fix some C R<1
which is <2 11 in the codes. We aim to prove that Gcont (, C) is determined. We
shall use the following large cardinal assumption: there exists a model M with cardinals
< so that:
M is a class model of ZFC ;
M is mildly iterable (see Appendix A);
is a Woodin cardinal of M;
is strong to + 1 in M. In other words, for every Z M& + 1, there exists
some extender E in M so that crit(E) = and Z Ult(M, E);
M& + 1 is countable in V.
This assumption is well above a measurable limit of Woodin cardinals ( is easily seen
to be a measurable limit of Woodin cardinals in M), but well below a Woodin limit of
Woodin cardinals.
By passing to L(M& + 1) we may add the following assumption:
M = L(M& + 1).
We shall use this extra condition in the proof.

3B (1) Names. Using the assumption that is continuous, fix some partial map
: < so that for any y , (y) = n iff (yi) = n for all sufficiently
large i < . is coded by a real. We may assume that this real belongs to M; if not,
use Remark 1E.5 or Woodin [45] with the first Woodin cardinal of M to absorb this real
into a small (relative to the image of ) generic extension of an iterate of M and replace
M by this generic extension.
Fix C R R in the pointclass <2 11 so that for any position y = $y | <
+ 1% of successor length + 1, y C $x, y % C where x =  y . We

may assume that C is lightface in a real which belongs to M; if not then again replace
M by a small generic extension of an iterate which absorbs the necessary real.
For $x, y % R R let C[x, y ] be the Martin auxiliary game associated to $x, y %,
and C ; see Section 2D and Martin [24]. The map C = (x, y  C[x, y ]) is Lipschitz
continuous and belongs to M. We shall refer to this map in our definition; it will be our
aid in deciding membership in C.
Using the assumption that is strong to + 1 in M, fix a sequence of extenders

E = $E | < lh(E)%  in M, all with critical point , all with M& Ult(M, E) and
iE () > , and such that for every Z M& + 1 there exists some < lh(E)
M  so that
Z Ult(M, E ).
For expository simplicity fix g which is col(, )-generic/M. Observe that g is also
generic over Ult(M, E) for any extender E in M.
90 3 Games of continuously coded length

Claim 3B.1. Suppose x is a real which belongs to M[g]. Then there exists some
 so that x belongs to Ult(M, E )[g].
< lh(E)
Proof. This follows immediately from the properties of E using some Z M& + 1
which codes a name for the real x. #
Definition 3B.2 (for a real x M[g]). ord(x) (the order of x) is the least so that x
belongs to Ult(M, E )[g].
For a code x = y | <  M[g], a sequence a (M&) M[g], and an
ordinal , we intend to define a game G ( a , x, ) played inside M[g]. G ( a , x, )
will consist of moves to produce a real y in the usual way, and auxiliary moves through
which player I tries to witness that either: (1) $y | < +1% is terminal in Gcont (, C)
and won by I; or (2) $y | < + 1% is non-terminal and I can win (G )( a, x , )
where x = y | < + 1, (G ) is a shift of G by some embedding , a is a
shift by of some initial segment of the auxiliary moves made in G ( a , x, ), and
is smaller than the shift of . Condition (2), when made precise, is self perpetuating
and will enable a construction that continues until securing condition (1).
It will be a while before we get to the actual definition of G . We start by developing
some notation around it.
Definition 3B.3. A[ ] is the canonical name for the set of $ a , x% M[g] so that:
a (M&) ; x is a code; and player I wins G (
a , x, ) in M[g].
Let A[ ] denote the auxiliary games map associated to A[ ], , and X = M&.
Remark 3B.4. The use of the set X = M& instead of X = for the auxiliary games
map will later ensure that the pivots we construct reside above .
The definition of G (
a , x, ) will make reference to the auxiliary games maps A[ ]
for < , and in fact to the association A = (  A[ ]). The definition is

thus an induction on .
Fix an ordinal . Work under the assumption that A is known. We shall end
with a definition of G (. . . , ), giving rise to A[ ] and by extension to A[ ]. We start
with the definition of an intermediary game.
Let M[g] be a small generic extension of M, where small means of size less than .
Let x = y | <  be a code belonging to M[g]. Let P (M&)< . Let s < .
For simplicity assume that lh(s) lh(P ). Suppose that (s) is defined. (Later on we
shall work in situations where lh(P ) = (s), but for the time being let us not make this
assumption.) We define the game Gmain (x, s, P , ) associated to these objects.
Set n = lh(P ). Gmain (x, s, P , ) is played inside M[g] according to Diagram 3.2
and rules (1)(3) below.
We use m to index rounds in Gmain (x, s, P , ). The game starts with m = lh(P ).
The objects played in round m are m , amI , a
mII , and y (m). We use Pn to denote P ,

and inductively define Pm+1 = Pm , amI , amII . We let y n = sn, and inductively
define y m + 1 = y m, y (m). We let x denote $y | < %, y . Note
that x depends continuously on y .
3B First determinacy result, part I 91


I n anI n+1 an+1I
y (n) y (n + 1)
II anII an+1II

Diagram 3.2. The game Gmain (x, s, P , ).

(1) (Rule for I) m is an ordinal chosen so that:

(a) Pm is a legal position in A[m , x ];


(b) m < if m = n; and

(c) m < m1 if m > n.

(2) (Rule for I and II) y (m) is a natural number, played by I if m is even and by II
if m is odd. For m < lh(s) we require y (m) = s(m).

(3) (Rule for I and II) amI
and amII are legal moves in the game A[m , x ], fol-

lowing the position Pm .

The first player to violate any of the rules of Gmain (x, s, P , ) loses. Note that,
because of rule (1c), there are no infinite runs of Gmain (x, s, P , ); one of the players
must violate a rule at some point.
Observe how rule (2) and the assignment y n = sn together force y to extend s.
(The assumption for simplicity that lh(s) n is used in the indexing of moves, allowing
us to start with y (n) in round n.) The fact that y extends s implies that (y ) = (s).
o(y ) is therefore determined from the start, and does not depend on the moves y (m).
Using Property 3A.2 it follows that x m depends only on (y for < and) y m.
So the rules for the first m + 1 rounds in A[m , x ] are known once y (m) is played.
Rule (3) therefore makes sense. The rules for the first m rounds of A[m , x ] are
known before y (m) is played, and so rule (1a) too makes sense. Note finally that
m < by rules (1b) and (1c). Knowledge of A is thus sufficient for the definition
of Gmain (x, s, P , ).
Let main (A)[x, s, P , ] express the statement player I wins Gmain (x, s, P , ).
(The only part of A relevant to the truth value of main (A)[. . . , ] is A .)
Since winning a clopen game is absolute, player I wins Gmain (x, s, P , ) iff
M[g] |= main (A)[x, s, P , ].

Remark 3B.5. A position of length m in Gmain (x, s, P , ) consists of the tuples y m


. The game G
and Pm , and the ordinals n , . . . , m1 main (x, s, P , ) from that position

onward is simply the game Gmain (x, max{s, y m}, Pm , m1 ). (s and y m are com-

patible sequences. By their maximum we mean the longer of the two.) In particular it
% is a winning position for I in G
follows that $y m, Pm , m , . . . , m1 main (x, s, P , )
iff I wins Gmain (x, max{s, y m}, Pm , m1 ), and similarly for II.
92 3 Games of continuously coded length

Let us now define the game G (a , x, ). We work with some code x = y | < 
belonging to M[g], and some a (M&) , again belonging to M[g].
G (
a , x, ) is played inside M[g] according to Diagram 3.3. Players I and II
collaborate as usual to produce a real y = $y (i) | i < %. In addition they play
auxiliary moves in the game C[x, y ]. They continue this way until, if ever, reaching
an i < so that:

(a) (y i) is defined;

(b) (y i)  {n | < }; and

(c) i (y i).

The first player to violate any of the rules of C[x, y ] loses. Infinite runs where an
i < satisfying conditions (a)(c) was not reached are won by player I.
If an i < satisfying conditions (a)(c) is reached, the game ends. We set
s = y i and P = a (s). We let N = Ult(M, Eord(x) ) and let be the ultra-
power map. (See Definition 3B.2 for the definition of ord(x).) I wins iff N[g] |=
main ((A))[x, s, (P ), ( )].

I y (0), 0 y (2), 2 ...


II y (1), 1 y (3), 3 ...

Diagram 3.3. G (
a , x, ), so long as (y i) is not defined or belongs to {n | < }.

An i satisfying conditions (a)(c) corresponds to a stage where its clear that $y |


< +1% will not be terminal in Gcont (, C). Intuitively G (
a , x, ) consists of moves
to produce y , and auxiliary moves verifying membership in C for as long as it seems
that $y | < + 1% may be terminal. If ever it becomes clear that $y | < + 1% is
not going to be terminal then the game ends, and instead the players pass to a shift of
Gmain through the reference to main in the payoff.
Strictly speaking condition (c) in the definition of G is not necessary. It has to do
with the assumption for simplicity that lh(s) lh(P ) in the definition of Gmain (. . . ),
and amounts to nothing more than some delay in the ending of G ( a , x, ). We could
drop this condition at the price of complicating Diagram 3.2.
Note that N[g] is a small generic extension of N, where small is meant with respect
to (). This follows from the fact that the extenders in E send above and has to
do with the large cardinal assumption of this section. Moreover x belongs to this small
generic extension of N by Definition 3B.2 (see also Claim 3B.1). The use of main in
the payoff of G therefore makes sense.
Note finally that the definitions here are made by induction on . The definition
of A[ ] requires knowledge of G (. . . , ), and the definition of G (. . . , ) in turn
require knowledge of the map  A[ ] for < .
3B First determinacy result, part I 93

3B (2) The basic step. The work in Section 3B (1) defines, in M, a name A[ ] for
each ordinal . Let A be the map which assigns to each < jumpM () the name A[ ].
The restriction to < jumpM () is in line with Section 1F. Its point is to make A a set
in M, rather than a class over M. There is nothing of importance about the specific use
of jumpM (), except that it is larger than the -s we shall care about later on.
Let Amix be the mixed pivot games map associated to A, , and X = M&. Let
mix be the mixed pivot strategies map associated to the same objects. See Section 1F
for the definition of these maps.
We work below with a code x = y | <  and a mixed pivot P for x, and show
how I can win to construct y and a mixed pivot P for x = y | < + 1, so that
P extends P(y ). The construction will be used later for the successor mega-round
in a construction of a run of Gcont (, C).
Let us be more precise. Fix some small (with respect to ) generic extension M[g]
of M. Fix some code x = y | <  in M[g]. Fix some  : M& + 1,
not necessarily a surjection. Fix an infinite run P of Amix [x], played according to
mix [, x]. P consists of the objects T , a , f , and  . We constantly use the terminology
of Section 1F when working with these objects.
Fix some s < . Suppose that (s) is defined. Suppose that lh(s) (s). Let
P = a (s). Let n denote (s) = lh(P ).
Let b be an infinite odd branch of T . Let Mb be the direct limit along b and let
jb : M Mb be the direct limit map.
Recall that b = (P, b) is the shift to Mb of the ordinal of  which corresponds
to the root of b, see Definition 1F.6.
Suppose that:

(1) Mb [g] |= main (jb (A))[x, s, P , b ].

Remark 3B.6. The extenders used in T have critical points above , by Lemma 1F.12
and through the use of X = M& noted in Remark 3B.4. The embeddings of T
therefore extend to act on M[g], and x M[g] also belongs to Mb [g].

Fix an imaginary opponent, willing to play for II in the standard length game on
natural numbers, starting from s. Fix some  : M& + 1, which agrees with 
to n. We shall construct:

(A) A real y extending s. y (i) for odd i lh(s) will be played by the imaginary
opponent.

(B) An infinite run P of Amix [x ], played according to mix [ , x ], which agrees


with P on the first n rounds. x here equals $y | < %, y .

If  is chosen onto M& + 1, then P is mixed pivot for x by Lemma 1F.11.


Observe that (y ) = n, since y extends s. If follows using Claim 3A.3 that
x n = xn. This implies that Amix [x] and Amix [x ] have the same rules for the first n
rounds. Similarly mix [, x] and mix [ , x ] agree on the first n rounds. Condition (B)
therefore makes sense.
94 3 Games of continuously coded length

Let n < be largest so that f (n) belongs to b. n is equal to the root of b in the
terminology of Section 1F. Let k be the first element of the odd branch b which is larger
than f (n) and larger than e(n). Let equal jf (n),k (n ). is the shift to Mk of the
ordinal of  corresponding to the root of b. Since b is the shift of that ordinal to Mb ,
b = jk,b ( ). By Lemma 1F.13, P = a n belongs to Mk , and is not moved by jk,b .
Pulling assumption (1) down using jk,b therefore gives:
(2) Mk [g] |= main (j0,k (A))[x, s, P , ].
(Note that jk,b extends to act on Mk [g], and therefore on x, see Remark 3B.6.) Using
this condition fix Mk [g], a winning strategy for player I in j0,k (Gmain )(x, s, P , ).
Let us begin round n of P . (P n is set equal to Pn.) Playing for I we
play f (n) = k. This forces us to play an extension T k + 1 of the iteration tree
T e(n) + 1 = T e(n) + 1. We play T k + 1 = T k + 1. Thus Mf (n) = Mk . This
move is presented in Diagram 3.4. The top line in the diagram presents the old config-
uration of P. The bottom line shows how P departs from P in round n. We leave it to
the reader to check that the conditions of rule (1) in Section 1F (1) are satisfied by our
move for P . For rule (1d) one uses the fact that T satisfies rule (1d) and conditions
(i)(iii) in Section 1F(1).

II
$
P Me(n) _ _I _ Mf (n) Mf (n)+1 Mf (n)+2 _ _I _ Mf (n+1) Mk

P Me(n) _ _ _ _ _ _ _ _ _ _ I_ _ _ _ _ _ _ _ _ _ _ Mk
(Me (n) ) (Mf (n) )

Diagram 3.4. First move in round n of P .

The rest of the construction follows the lines of Section 1E. We proceed in rounds
indexed by m, starting with m = n.
The moves m , lm , p , u , E
m m f (m) , Ef (m)+1 , and wm are played by a combined
effort involving mix [ , x ] and appropriate shifts of the strategy fixed using con-

dition (2) above. We refer the reader to Diagram 3.5. The natural number y (m) is
determined by s if m < lh(s), played by the imaginary opponent if m lh(s) is odd,
and played by the shift of if m is even. (Diagram 3.5 is drawn on the assumption
that n is even.) Note that once y (m) is played, enough of x is known to determine
the behavior of mix [ , x ] in round m.
We set f (m) equal to e (m) in all rounds m > n. This means that I does not play
an interval of T in those rounds; we simply set T f (m) + 1 equal to T e (m) + 1.
Remark 3B.7. The construction requires shifting along the branch k, k+2, k+4, . . .
of T . This shifting is possible since belongs to Mk [g], a small generic extension
of Mk relative to , and since all the extenders used in T have critical points above
3B First determinacy result, part I 95

  
_ _ _ _ _ _ Mk
Mk+1
Mk+2
Mk+3
Mk+4
Mk+5
Mk+6
(Mf (n) ) (Mf (n+1) ) (Mf (n+2) ) (Mf (n+3) )

_
n

y (n)

ln
/o /o /o /o /o /o /
pn
un
mix [ ,x ]
_ wn


jk,k+2 ( )
n+1

Oppnt or s(n+1) y (n+1)


ln+1

jk,k+2 ( )
/o /o /o /o /o /o /
pn+1
un+1
mix [ ,x ]
_
wn+1


jk,k+4 ( )
n+2


jk,k+4 ( ) y (n+2)


ln+2

jk,k+4 ( )
/o /o /o /o /o /
pn+2
un+2
mix [ ,x ]
_
wn+2

Diagram 3.5. The construction of P .

by Lemma 1F.12 and Remark 3B.4. The fact that belongs to Mk [g] follows form
the assumption that x belongs to M[g]. Another use of this assumption was made in
Remark 3B.6 and in passing from assumption (1) to condition (2).
96 3 Games of continuously coded length

We leave it to the reader to check that mix , shifts of , and the imaginary opponent
together cover all the necessary moves for the construction of P , and only note the
following properties of the construction:

(i) n < . (Recall that = jf (n),k (n ).)



(ii) m < jk+2(mn)2,k+2(mn) ) for m > n.
(m1

Both properties follow from the rules of j0,k (Gmain )(x, s, P , ) and its shifts to the

models Mk+2(mn) . Property (i) follows from rule (1b) in Section 3B (1), and property
(ii) follows from rule (1c).

Lemma 3B.8. If P is useful (see Definition 1F.8), then so is P .

Proof. Fix m > 0. Let m < m be largest so that f (m) T f (m). We wish to show
that:

() m < jf (m),f (m) (m ).

If m < n then condition () follows from the assumption that P is useful, since P
agrees with P for the first n rounds.
If m > n then condition () follows immediately from property (ii) above. f (m)
is equal to k + 2(m n) in this case, and m is simply equal to m 1.
So suppose that m = n. f (m) is equal to k in this case. m < m is largest so that
f (m) T k. Since T k + 1 was set equal to T k + 1, and since f n was set equal

to f n, it follows that f (m) T k. Since k belongs to the odd branch b it follows that
f (m) b.
Recall that n < is largest so that f (n) b. Combining this with the conclusion
of the previous paragraph we see that f (m) [0, f (n)]T . Using the fact that P is
useful it follows from this that:

(iii) n jf (m),f (n) (m ).

Equality holds if m = n. If m < n then the inequality of condition (iii) follows directly
from the conditions placed on  by Definition 1F.8.
Shifting condition (iii) to Mk via jf (n),k , and composing the result with the inequality
of condition (i), we obtain:

(iv) n < jf (m),k (m ).

jf (m),k and jf (m),k are the same since T k + 1 = T k + 1. f (m) and f (m) are the
same since f n = f n. m and m are the same since  n =  n. Folding these
facts into condition (iv) we immediately obtain condition () in the case m = n. #

3B (3) I wins. Work in M. Define the game Gini to be played according to Diagram 3.6
and rules (1) and (2) below.
for $a , a
We write am mI mII % and Pm for $a0 , . . . , am1 %.
3B First determinacy result, part I 97

I 0
a0I 1
a1I ...
II
a0II
a1II ...

Diagram 3.6. The game Gini .

(1) (Rule for I) m is an ordinal chosen so that:

(a) Pm is a legal position in A[m , ];


(b) m < jumpM () if m = 0; and

(c) m < m1 if m > 0.

(2) (Rule for I and II) amI
and amII are legal moves in the game A[m , ],

following the position Pm .

The first player to violate any of the rules loses. Because of rule (1c) there are no
infinite runs of Gini .
Gini is similar to the games Gmain (. . . , ) defined in Section 3B (1). There are a
couple of differences. There is no y played here and we work with the code x = 
for the empty position. There is no ordinal given at the start; instead 0 must be
smaller than jumpM (). There is also no Pn = P given at the start; the game begins
with round 0.
Observe that Gini , being a clopen game, is determined. The winning player has a
winning strategy in M.

Lemma 3B.9. Suppose that I wins Gini . Then I wins the long game Gcont (, C).

Proof. Fix an imaginary opponent, willing to play for II in Gcont (, C). Fix a mild
iteration strategy  for M. Working with  and the imaginary opponent we construct
(among other things) a run of Gcont (, C). We shall verify at the end that the run
constructed is won by player I.
We work in stages, starting with stage 0. At the start of stage we have the objects
indicated in items (A)(E).

(A) Reals y for so that + 1 < .

Note the restriction to + 1 < in item (A). In stage , for a successor, we construct
the real y1 . We do not construct any reals in stage 0, and we do not construct any
reals in limit stages.
Let n denote (y ). We shall end the construction as soon as we construct some
y so that (y ) is not defined, or defined and an element of {n | < }. At the start
of each stage we therefore have:

(1) n , + 1 < , are all distinct.


98 3 Games of continuously coded length

We use x to denote y | < . When ending the constructionwhen constructing


some y so that (y ) is not defined or defined and an element of {n | < }we
shall make sure that $x , y % C . We shall thus see that our run of Gcont (, C) is
won by player I, as required.

(B) A -iteration of M = M0 giving rise to models $M | < % and embeddings


$j, | < % between them.

(C) Maps  : M &j0, () + 1 for < .

(D) Infinite runs P of j0, (Amix )[x ], played according to j0, (mix )[ , x ], for
< .

Using Claim 3A.4 we know that:

(2) (For n < and < .) Suppose n n for all [, ). Then x agrees with
x to n. Precisely, x n = x n.

By means to be described shortly we shall obtain the following additional agreement


condition:

(3) (For n < and < .) Suppose n n for all [, ). Then  agrees with
the image of  to n. Precisely,  n = j, ( n).

As we construct in stage we shall make sure further that:

(i) P is useful.

(ii) (For n < and < .) Suppose n n for all [, ). Then P agrees
with the image of P on the first n rounds. Precisely, P n = j, (P n).

Note that conditions (2) and (3) are needed to make sense of condition (ii). More
precisely, condition (2) is needed to ensure that the first n rounds of j0, (Amix )[x ]
and the first n rounds of the image to of j0, (Amix )[x ] have the same rules, and
both conditions are needed to make sure that j0, (mix )[ , x ] and the image to of
j0, (mix )[ , x ] are the same for the first n rounds.
Finally we would like to know that:

(4)  is onto M &j0, () + 1.

This is a matter of bookkeeping, but the bookkeeping is not entirely straightforward.


For limit condition (3) completely determines  , leaving us no freedom in trying to
obtain condition (4). We must thus plan ahead, before reaching , to make sure that
condition (4) holds at .
To best describe the bookkeeping involved in the construction of the maps  :
M &j0, ()+1 it is convenient to change from maps with domain to maps with domain
( ). Fix some bijection e : ( ). The precise nature of e is not
important, so long as:
3B First determinacy result, part I 99

(a) For each n < , e n (n ).


Fix some surjection 0 : M& + 1. This is possible using the initial large cardinal
assumption, specifically the condition saying that M& + 1 is countable in V.
At the start of stage we have, in addition to the objects described in items (A)(D)
above:
(E) Maps +1 : {n } M +1 for + 1 < .
We do not construct any for limit . for successor will be constructed
during stage . We shall pick so that:
(iii) (For successor .) is onto M &j0, () + 1.
By induction then we shall know that +1 : {n } M +1 &j0, +1 () + 1 is onto
for each + 1 .
Let us now describe how to construct  in stage . Using the surjections 0 ,
{ } +1< , and (if is a successor) , define a partial map : ( )
M &j0, () + 1 by:

= (j0, 0 ) (j +1, +1 ).
+1

The domain of is equal to +1 ({n } ). Moreover:
(b) The range or contains j0,  (M0 & + 1).
(c) For each + 1 , the range of contains j +1,  (M +1 &j0, +1 () + 1).
Condition (b) follows from the fact that 0 is onto M& + 1. Condition (c) follows from
the fact that +1 is onto M +1 &j0, +1 () + 1.
Define  : M &j0, () + 1 by:

(e(i)) if e(i) dom( ); and
 (i) = (3.1)
otherwise.

This definition takes care of both conditions (3) and (4) above. Condition (3) fol-
lows immediately from the definition of  , the definition of , and condition (a). For
successor condition (4) follows from condition (c) with + 1 = . For = 0 condi-
tion (4) follows from condition (b). For limit condition (4) follows from condition (c)
so long as we make sure that:
(iv) (For limit .) Every element of M is of the form j +1, (w) for some + 1 <
and some w M +1 .

It is time now to start the construction. We divide into the cases = 0; is a limit;
and is a successor. For = 0 we must simply construct P0 . For limit we must
construct M and P . For successor we must construct M , y1 , , and P . In
each case we must verify conditions (i)(iv).
100 3 Games of continuously coded length

Initial stage. We start with = 0. Using the assumption of Lemma 3B.9 fix a winning
strategy M for player I in Gini .
Recall that 0 , defined above using 0 , is a map from onto M0 & + 1.
Let P0 be the run of Amix [] generated as follows:
mix [0 , ] supplies moves for player II.
f0 (m) is set equal to e0 (m) in each round m. This means that the iteration tree
created has a single even branch consisting of {f0 (m) = 2m | m < }.
and its shifts along the even branch supply moves for player I.
These three conditions cover all moves in Amix [], giving rise to an infinite run P0 .
The ordinals m supplied by and its shifts must satisfy rule (1c) of Gini and its shifts
(else would lose). It follows that P0 is useful. Condition (i) therefore holds for
= 0. Conditions (ii)(iv) are vacuous. # (Initial stage)
Successor stage. Let = + 1 be a successor ordinal. By conditions (D), (4), and (i),
P is a useful mixed pivot for x where mixed pivot is interpreted over M and using
j0, (A). Let b be the cofinal branch through T (the iteration tree given by P ) chosen
by the iteration strategy . Let Q be the direct limit along b . Let k : M Q
be the direct limit embedding. Since P is useful (see Definition 1F.8), b must be an
odd branch. Since P is a mixed pivot for x there must exist some h so that:
(S1) h is col(, (k j0, )())-generic/Q ; and
a , x % (k j0, )(A)[ ], where = (P , b ).
(S2) $
Let G denote the game (k j0, )(G )( a , x , ). Combining condition (S2)
with Definition 3B.3 we see that I wins G in Q [h ]. Fix a strategy Q [h ]
which witnesses this.
G is played according to Diagram 3.3. Let us begin to form a run of this game.
We use to obtain moves for I. The imaginary opponent supplies natural number
moves for II, playing y (i) for odd i. We ascribe indiscernibles for Q as ordinal
moves for II.
More precisely: Let = (k j0, )() + 1. M &j0, () + 1 is countable in Vthis
is witnessed by  and from this it follows that Q & is countable in V. The large
cardinal assumption at the start of the section is more than enough to guarantee that every
real has a sharp. (This is seen as usual using the genericity iterations, Remark 1E.5
or Woodin [45], with the first Woodin cardinal of M.) Q & therefore has a sharp.
The initial assumption on M includes the condition M = L(M& + 1). Thus Q =
L(Q &). The Silver indiscernibles for Q & are therefore indiscernibles for Q .
Working by induction on i < we define y i and sequences Pi of ordinals, of
length i. We construct so that y i and Pi together form a position in G according to
; so that IIs move in Pi are Silver indiscernibles; and so that IIs moves in y i are
the ones given by the imaginary opponent. The argument is standard, see for example
Martin [24].
3B First determinacy result, part I 101

We continue this way until, if ever, reaching an i which satisfies conditions (a)(c)
of Section 3B (1).
If no such i is reached we end up producing a real y so that (y ) is not defined,
or defined and an element of {n | < }. We also produce a sequence of positions Pi
of increasing lengths in (k j0, )(C)[x , y ] consisting of indiscernibles for II and
moves given by for I. The argument of Martin [24] shows under such circumstances
that $x , y % belongs to C . This is the end of the construction. We have reached a
position $y | < + 1% where Gcont (, C) ends, and ends with a win for player I.
So suppose an i satisfying conditions (a)(c) of Section 3B (1) is reached. G ends
at the first such i. Let s denote the position y i obtained in G . Since was picked
a winning strategy for I, G ends with a win for I. Looking at the winning condition
indicated in Section 3B(1) we see that:
(S3) N [h ] |= main (( k j0, )(A))[x , s , (P ), ( )].
Recall that = (P , b ). N here is equal to Ult(Q , F ), where F is the
 is the ultrapower embedding.
(k j0, )(ord)(x )-th extender of (k j0, )(E).
P is equal to a (s ). Note how all this corresponds to the winning condition of
G (. . . ) defined in Section 3B(1).
The current configuration of models (M , Q , N ) and the configuration of em-
beddings between them are presented in the lower line of Diagram 3.7. To proceed we
shall have to look at these models in a different light.

j,+1 j

j j
jjjjjjj F $
jjjjjjj 
jTT
M T / M+1 TjTTTT / Q
TTTT k Q TTTTk O
TT
T T 
D 

copy with j,+1




jjj 
j
Tjjjjj / Q

/ N
M TTTT
TTTTk F
T

Diagram 3.7. F applied to Q (lower line); and F applied to M (upper line) followed by
copying.

Let = j0, (). By Lemma 1F.12 and Remark 3B.4, P resides above . It
follows first that M and Q agree beyond , and secondly that crit(k ) > . Now
F has critical point k ( ), equal to since k does not move this ordinal. So
crit(F ) < crit(k ). F can therefore be applied to M .
102 3 Games of continuously coded length

Let M+1 = Ult(M , F ). Let j,+1 be the ultrapower embedding. The situation
is presented in the upper left part of Diagram 3.7.
We have so far defined M = M+1 . Let us already here note that regardless of
how we continue to construct the real y = y1 , we shall have:
(S4) (y ) = (s ).
This is because (s ) is defined and y , which we shall finish constructing shortly, is
going to extend s = y i. We thus know already here the value of n ; it is equal to
(s ).
From the fact that M &j0, () + 1 is countable in V and the fact that the iteration
leading form M to M+1 is countable it follows that M+1 &j0,+1 () + 1 is countable
in V. Fix a surjection +1 : {(s )} M+1 &j0,+1 () + 1. Let +1 , the
surjection of item (C), be defined using +1 via the bookkeeping algorithm leading to
equation (3.1) on page 99.
Note that +1 and j,+1 agree on ((s ) ); the only difference
between these two maps is the addition of +1 , whose domain is {(s )} . It
follows using condition (a) on page 99 that:
(S5) +1 agrees with j,+1  to (s ).

Let  denote j,+1  . Observe that  is then equal to m< j,+1 ( m).

Let P equal m< j,+1 (P m). P is thus a run of j0,+1 (Amix )[x ], played
according to j0,+1 (mix )[ , x ].
T , the iteration tree given by P , is simply the copy of T from M to M+1
via j,+1 . The copying is indicated in dotted lines in Diagram 3.7. Let Q be the
direct limit of the models of T along b , and let k be the direct limit embedding. Let
j : Q Q be the copy embedding.

It is easy to see that (P , b ) = j ( (P , b )). It is also true, though not as
easy, that a  m = j (
a m) for each m < . To prove this remember that a m is not
moved by a tail-end of the embedding of T along b , and similarly with a and T .
Lemma 3B.11, to be proved in Section 3B(5), states under the current circumstances
that there exists an elementary embedding : N Q , with crit( ) > ( ) and
such that = j . This embedding is presented in dashed line in Diagram 3.7.
Since crit( ) > (), extends to act on N [h ], a small generic extension of
N relative to ( ). Applying the extended embedding : N [h ] Q [h ] to
condition (S3) and using the equality = j we get:

Q [h ] |= main ((j k j0, )(A))[x , s , j (P ), j ( )].

By standard copying arguments, j k is equal to k j,+1 . We can thus re-write


the last equation as:

Q [h ] |= main ((k j0,+1 )(A))[x , s , j (P ), j ( )].


3B First determinacy result, part I 103

Now P is equal to a (s ). j (P ) is therefore equal to a (s ). is equal to


(P , b ). j ( ) is therefore equal to (P , b ). Making these substitutions in the
equation we get:

(S6) Q [h ] |= main ((k j0,+1 )(A))[x , s , P ,  ], where P = a (s ) and


 = (P , b ).

This precisely is the assumption (1) of Section 3B (2), only shifted to M+1 . (Remember
that Q is the direct limit of models of T along b , and kb : M+1
 Q is the direct
limit embedding.) We can now run the construction of Section 3B (2) with P standing
for P,  standing for , x standing for x, s standing for s, b standing for b, and
using +1 for  (note that +1 agrees with  to (s ) by condition (S5)).
Remark 3B.10. The construction in Section 3B (2) is made under the assumption that
x belongs to a smallwith respect to generic extension of M. (The need for
this assumption is explained in Remark 3B.7.) In appealing to the construction of
Section 3B (2) we are therefore making use of the fact that x belongs to a small
with respect to j0,+1 ()generic extension of M+1 . This fact traces back to the
large cardinal assumption on M, which allowed picking the extender F in such a
way that x was absorbed into Ult(Q , F )[h ]. From x Ult(Q , F )[h ] a quick
agreement calculation shows that x belongs to Ult(M , F )[h ]. h is generic for
col(, (k j0, )()). This is a small forcing relative to j0,+1 () since j,+1 , the
ultrapower embedding by F , sends its critical point above (k j0, )().
Running the construction of Section 3B(2) we obtain the real y , extending s , and
a run P of j0,+1 (Amix )[x+1 ], played according to j0,+1 (mix )[+1 , x+1 ]. Let
P+1 be this run. By construction, see Section 3B (2), P+1 agrees with P on the
first (s ) rounds. In other words:

(S7) P+1 n = j,+1 (P n ).

By Lemma 3B.8:

(S8) P+1 is useful.

The construction for stage = + 1 is now complete. Condition (i) holds for
by condition (S8) above. Condition (ii) for follows from the same condition for ,
and condition (S7). Condition (iii) was secured by the choice of +1 . Condition (iv)
is vacuous. # (Successor stage)

Limit stage. Work now with a limit ordinal . We have the picture of models presented
in Diagram 3.8. Let M be the direct limit of the models $M | < % under the
embeddings $j, | < %. Every element of M has pre-images in M for all
sufficiently large < . Condition (iv) therefore holds trivially. Condition (iii) is
vacuous here. So we need simply construct P and verify conditions (i) and (ii).
Using Claim 3A.1 fix an increasing sequence {n }n< so that:
104 3 Games of continuously coded length

j, +1 j +1, +2

kkk #
SkSkSkS
kk $
Sk
Sk / Q +1 F +1 M +2
F
M k SSSS k / Q M +1 k
SSk +1

T T +1

Diagram 3.8. The situation at limit stages.

Each n is smaller than and sup{n | n < } = .


n n for each [n , ).
For n m < it follows from the last item and from condition (ii) at m that:
(L1) jn ,m (Pn n) is equal to Pm n.
For each n < let Pn = jn , (Pn n). By condition (L1) then Pn is an initial segment
of Pm for n < m. Define:
P = Pn .
n<
Pn is a position of length n in j0, (Amix )[xn ]. By condition (2), xn n = x n. Pn is
therefore a position in j0, (Amix )[x ]. A similar argument using condition (3) shows
that Pn is according to j0, (mix )[ , x ]. So P is a run of j0, (Amix )[x ] played
according to j0, (mix )[ , x ], as required.
For each n < , Pn n is useful. This follows from condition (i) at n . Using the
elementarity of jn , we see that Pn is useful. By Claim 1F.9, P is useful. This takes
care of condition (i) at . Condition (ii) at follows directly from the definition of P ,
and the same condition at the n -s. # (Limit stage)
The initial, successor, and limit stage constructions together complete the proof of
Lemma 3B.9. # (Lemma 3B.9)

3B (4) Discussion. Morally the limit case is always the most important part in a de-
terminacy proof (in the context of long games). Each determinacy proof involves
perpetuating a carefully engineered condition through a transfinite construction. It is
easy enough to engineer the condition so that it propels itself through successor stages.
But propelling through limits requires special care. The condition must involve advance
planning in anticipation of future limit stages.
Consider for example the situation of Section 2B. We had there a model with
1 + + 1 Woodin cardinals, for [1, ]. We had a name A for a subset of R
in M col(, ) , and an associated auxiliary games map A. was a limit ordinal, and we
had a fixed sequence of ordinals $k | k < % converging to .
For illustrative purposes let G ( y ), where y R , denote the game in which
players I and II alternate moves in A[ y ] and player I in addition plays a descending
sequence of ordinals. The game is presented in Diagram 3.9.
3B First determinacy result, part I 105

I 0 a0I 1 a1II ......


II a0II a1II

Diagram 3.9. The illustrative game G (


y ).

Note how round k of G ( y ), following the position consisting of $a0 , . . . , ak1 % and
$0 , . . . , k1 % say, is exactly the same as the game Gk ( x , P , ) of Diagram 2.2 with
x = yk , P = $a0 , . . . , ak1 % and = k1 . The determinacy proof in Section 2B
was thus a matter of breaking G ( y ) into its constituent rounds, and plugging round k
of the game in stage k following the production of yk . As the construction in
Section 2B approached , moves in these rounds accumulated to produce the kind of
condition needed in stage . The planning for the limit stage was thus achieved
through spreading G ( y ) over stages cofinal in .
Consider now the case of continuously coded length. Let y be a position of limit
length in Gcont (, C), and let be an ordinal. For illustrative purposes consider the
game G ( y , ), played along the format of Diagram 3.9, but with the following rules
(where A[. . . ] is the map defined in Section 3B (1)):
(1) (Rule for I) m is an ordinal chosen so that:
(a) $a0 , . . . , am1 % is a legal position in A[m , 
y ];
(b) m < if m = 0; and
(c) m < m1 if m > 0.
(2) (Rule for I and II) amI and amII are legal moves for I and II respectively in the
game A[m ,  y ], following the position $a0 , . . . , am1 %.
Note the similarities between this game and the game Gmain (. . . ) displayed in
Diagram 3.2. Here too the determinacy proof can be viewed as breaking G ( y, )
into its constituent rounds, and distributing these rounds over earlier stages. A careful
look at the limit stage construction in Section 3B (3) shows precisely how this happens.
Round m of G ( y , ) is plugged into stage = + 1 for < largest so that n m.
Let ,m denote this stage.
A couple of points which distinguish the breaking and spreading done here from
that done in Section 2B are worth highlighting.
First, note that the manner of spreading here is not fixed at the outset. The sequence
$,m | m < % leading to depends not only on but also on y. In other words the
sequence depends on the actual moves played in the game Gcont (, C) up to .
Next, note that a given may serve more than one limit . In fact it may have to
serve infinitely many; there may be some m and infinitely many distinct limit ordinals
i so that i ,m = for all i.
This is the main point which distinguishes the method of approaching limits here
from the method used in Chapter 2. It creates a substantial difficulty. A tail-end of
the moves in the auxiliary game at stage must be reserved to propel the inductive
106 3 Games of continuously coded length

condition to stage + 1. Thus only a finite part of this auxiliary game can be devoted to
advance planning for future limits. Yet this finite part must somehow plan for infinitely
many limits.
Our solution is to engineer the proof in such a way that for  > , the auxiliary
game at stage  is the image under some elementary embedding of the auxiliary game
at stage . The moves planned in stage and then used in stage can thus be shifted
via this elementary embedding and used again in stage  . In fact the moves in round
m of stage are shifted and used in all limit stages with ,m = . Planning for one
limit now becomes the same as planning for all.
The key to this solution is lifting moves in the current auxiliary game so that they
can be used again in the next one. The game G (. . . ) designed in Section 3B (1) does
precisely this lifting, through its use of an ultrapower embedding by Eord(x) for the
reference to main .

3B (5) Internal ultrapowers vs. copying. During the proof of Lemma 3B.9 we ap-
pealed to a lemma which allowed us to switch between an internal ultrapower embedding
and a copy map. Here we phrase and prove this lemma.
We work with a model M, and a cardinal in M. Let T be a length iteration
tree on M, using only extenders with critical points above . Let b be a cofinal branch
through T , let Q denote the direct limit model along b, and let k : M Q be the direct
limit embedding. Fix an extender F Q with critical point equal to k() = . Note
then that F can be applied to M. Let M  = Ult(M, F ), and let j : M M  be the
ultrapower embedding. Let T  be the result of copying T to a tree on M  . Let Q be
the direct limit of the models of T  along b, and let k  : M  Q be the direct limit
embedding. Let j : Q Q be the copy map. The situation is similar to that presented
in the upper line of Diagram 3.7.
Let N = Ult(Q, F ), and let : Q N be the ultrapower embedding. The
situation is similar to that presented in the lower line of Diagram 3.7. We wish to prove:

Lemma 3B.11. There is an elementary embedding : N Q with crit( ) > ()


and such that = j.

Proof. We use the following simple claim throughout the proof:


Claim 3B.12. Let X < be given. Then j(X) is equal to (X).

Proof. First calculate that:

j(X) = (j k)(X)
= (k  j )(X)
= j (X).

The first equality uses the fact that crit(k) > (all extenders in T have critical points
above by assumption). The second equality uses the commutativity of Diagram 3.10.
The third equality uses the fact that crit(k  ) > j ().
3B First determinacy result, part I 107
j
jjjjjjj
jTT
MO  T / 
TTTT k  QO
TT
F Q T

j j

jj
jjjjjjj
MT
jTTT /
TTTT k Q
TT
T

Diagram 3.10. F Q applied to M, and copying.

Now note that j (X) is equal to (X); both j and are ultrapower embeddings
by F , and so agree on subsets of crit(F )< = < . #

Every element of N is of the form (f )(a) for some a ( < ) in the support of
F and some function f : lh(a) Q in Q. ( lh(a) ) = j( lh(a) ) by Claim 3B.12, so
a belongs to the domain of j(f ). We can therefore consider the object j(f )(a).

Definition 3B.13. For w = (f )(a) N , where a belongs to the support of F and


f : lh(a) Q is a function in Q, define (w) = j(f )(a).

Definition 3B.13 is somewhat premature, but the next claim shows that it makes
sense.

Claim 3B.14. Suppose a ()< belongs to the support of F . Suppose f, g Q are


both functions from lh(a) into Q. Suppose that (f )(a) = (g)(a). Then j(f )(a) =
j(g)(a).

Proof. Let X = {t lh(a) | f (t) = g(t)}. By assumption a (X). By Claim


3B.12, j(X) = (X). It follows that a j(X), so j(f )(a) = j(g)(a) as required.
#

is now well defined. It remains to show that it satisfies the requirements of


Lemma 3B.11.

Claim 3B.15. : N Q is elementary.

Proof. Fix w N , say w = (f )(a). Fix a formula . Suppose N |= [w]. We wish


to show that Q |= [ (w)].
Let X = {t lh(a) | Q |= [f (t)]}. Since N |= [(f )(a)], a (X). Using
Claim 3B.12 it follows that a j(X), so Q |= [j(f )(a)] as required. #

Claim 3B.16. = j.
108 3 Games of continuously coded length

Proof. Let w Q be given. Let f : 0 Q be the function whose value at is w.


(w) is equal to (f )(). ((f )()) by definition is j(f )(), which is equal to j(w).
Thus ( )(w) = j(w) as required. #

Note that (()) = j() = () by commutativity and Claim 3B.12. To complete


the proof we must verify that () is the identity. We shall then know that crit( ) >
().
Fix some w (). Then w = (f )(a) for some a in the support of F and
some f : lh(a) Q with range(f ) . Now (w) = j(f )(a) by definition, and
j(f ) = (f ) by Claim 3B.12. (f can be regarded as a subset of < ; this uses
the fact that range(f ) .) It follows that (w) = (f )(a) = w, as required.
# (Lemma 3B.11)

We end by noting that, with a slight extra assumption involving the extenders used
in T , Lemma 3B.11 actually improves to the statement that Q is equal to N and j is
equal to .

Lemma 3B.17. Suppose that the extenders used in T are all closed. Then Q equals
N and j equals .

Proof. It is enough to show that the embedding given by Lemma 3B.11 is onto. We
would then have an -preserving map from one transitive structure onto another. The
structures would have to be equal and the map, , would have to be the identity.
So fix some u Q . We wish to show that u belongs to the range of .
Let $Mi | i < % denote the models of T  . For i b let ki,b
 : M  Q denote the
i b
embedding given by T  . Let Mi and ki,b denote the corresponding objects for T . Fix a
node l b large enough that u has a pre-image in Ml . Fix w Ml so that kl,b
 (w) = u.
  
Now Ml , being a finite iterate of M , is actually contained in M which is equal
to Ult(M, F ). So there is some a j ()< in the support of F and some function
g : lh(a) M in M so that w = j (g)(a). Since w belongs to Ml , g can be picked
with range(g) Ml .
Ml is closed under sequences of length in M; this uses the assumption that the
extenders of T are closed. It follows that g belongs to Ml . Let f = kl,b (g). This is
a function which belongs to Q.
By standard copying arguments, specifically the commutativity of Diagram 3.11,
j(f ) = kl,b
 (j (g)). Using this identity calculate that:

j(f )(a) = kl,b



(j (g))(a)

= kl,b (j (g)(a))

= kl,b (w)
= u.
 ) > j () and a
The equality leading to the second line uses the fact that crit(kl,b
j ()< .
3C First determinacy result, part II 109

MO  / M / Q

k0,l Ol 
kl,b O

j j Ml j

M / Ml /Q
k0,l kl,b

Diagram 3.11.

j(f )(a) belongs to the range of by Definition 3B.13. The calculation above shows
that u = j(f )(a). So u range( ) as required. #

The closure assumed in Lemma 3B.17 is easy to arrange. But the only advantage
of the conclusion of Lemma 3B.17 over the conclusion of Lemma 3B.11 is aesthetical;
it simplifies slightly the presentation of Diagram 3.7, eliminating the need for . It is
hard to decide which is better: putting in the extra work needed to secure the assumption
of Lemma 3B.17; or settling for the weaker conclusion of Lemma 3B.11. We chose the
latter.

3C First determinacy result, part II


Let us continue with the proof that Gcont (, C) is determined. Our plan is to mirror the
definitions of the previous section to end with a game Hini ; prove that if II wins H
ini
then II wins Gcont (, C); and most importantly rule out the possibility that II wins Gini
.
and I wins Hini

3C (1) II wins. Mirroring the definitions in Section 3B(1) we work to define games
 x, ) associated to a b (M&) M[g], a code x = y | <  M[g],
H (b,
and an ordinal . The games are clopen and played inside M[g]. Simultaneously we
set:
 x% M[g] so that:
Definition 3C.1. B[ ] is the canonical name for the set of $b,
 
b (M&) ; x is a code; and player II wins H (b, x, ).

We use B[ ] to denote the mirrored auxiliary games map associated to B[ ], ,


and X = M&.
Fix for a moment a small, relative to , generic extension M[g] of M. Fix an
ordinal . Let x = y | <  be a code belonging to M[g]. Let Q (M&)< .
Let s < . Assume that lh(s) lh(Q). Suppose that (s) is defined. We define
110 3 Games of continuously coded length

under these assumptions the game Hmain (x, s, Q, ). The definition mirrors that of

Gmain (x, s, P , ) in Section 3B(1).


(x, s, Q, ) is played inside M[g] according to Diagram 3.12
Set n = lh(Q). Hmain
and rules (1)(3) below.


bnI
bn+1I
I
y (n) y (n + 1)
II n bnII n+1 bn+1II

(x, s, P , ).
Diagram 3.12. The game Hmain

(x, s, Q, ). The game starts with m = lh(Q).


We use m to index rounds in Hmain

The objects played in round m are m , bmII , and y (m). We use Q to denote
, bmI n

Q, and inductively define Qm+1 = Qm , bmII , bmI . We let y n = sn, and in-

ductively define y m + 1 = y m, y (m). We let x denote $y | < %, y .

(1) (Rule for II) m is on ordinal chosen so that:

(a) Qm is a legal position in B[m , x ];


(b) m < if m = n; and

(c) m < m1 if m > n.

(2) (Rule for I and II) y (m) is a natural number, played by I if m is even and by II
if m is odd. For m < lh(s) we require y (m) = s(m).

(3) (Rule for II and I) bmII
and bmI are legal moves in the game B[m , x ],

following the position Qm .
(x, s, P , ) loses. Note that
The first player to violate any of the rules of Hmain
sooner or later one of the players must violate a rule; there are no infinite runs of the
game, because of rule (1c).
Let main (B)[x, s, Q, ] express the statement II wins Hmain (x, s, Q, ). Only

B is relevant to the truth value of main (B)[. . . , ].


We can now define H (b,  x, ). Work with a code x = y | <  M[g], some
b (M&) M[g], and an ordinal . H (b,
 x, ) is played according to Diagram 3.3.
Players I and II collaborate as usual to produce a real y . In addition they play auxiliary
moves in the game C[x, y ]. They continue this way until, if ever, reaching an i <
which satisfies condition (a)(c) in Section 3B(1). The first player to violate any of the
rules of C[x, y ] loses. Infinite runs, where an i < which satisfies conditions (a)(c)
has not been reached, are won by player I, just as they were in Section 3B (1). If an
i < which satisfies conditions (a)(c) is reached, the game ends. We set s = y i

and Q = b(s). As in Section 3B(1) we let N = Ult(M, Eord(x) ) and let be the
ultrapower embedding. II wins iff N |= main ((B))[x, s, (Q), ( )].
3C First determinacy result, part II 111

The only difference between H (. . . ) and G (. . . ) is in the payoff condition when


an i < which satisfies conditions (a)(c) is reached. In the case of G (. . . ) we used
main to determine the payoff for I. Here we use main to determine the payoff for II.
H (. . . , ) is defined modulo knowledge of the map  B[ ] for < .
The definition of B[ ] requires knowledge of H (. . . , ). As usual both definition are
made by simultaneous induction on .
, mirroring the game G of Section 3B (3). H is played
Define finally a game Hini ini ini
according to Diagram 3.13 and rules (1) and (2) below.

I
b0I
b1I ...
II 0
b0II 1
b1II ...

.
Diagram 3.13. The game Hini

for $b
We write bm mII , bmI % and Qm for $b0 , . . . , bm1 %.

(1) (Rule for II) m is an ordinal chosen so that:

(a) Qm is a legal position in B[m , ];


(b) m < jumpM () if m = 0; and

(c) m < m1 if m > 0.

(2) (Rule for II and I) bmII
and bmI are legal moves in the game B[m , ],

following the position Qm .

The first player to violate any of the rules loses. The game is clopen; because of
rule (1c) there are no infinite runs.
The following lemma is a direct mirror image of Lemma 3B.9:
. Then II wins
Lemma 3C.2. Suppose that there exists an ordinal so that II wins Hini
the long game Gcont (, C). #

Proof. Simply mirror the construction of Section 3B (3). For the most part the adapta-
tion is trivial. Let us only comment on the situation at the end of the construction.
The construction ends in the successor stage + 1 if, while playing H , an i <
which satisfies conditions (a)(c) of Section 3B (1) is not reached.
The situation, adapted from the successor stage in Section 3B (3), is as follows:
H denotes the game (k j0, )(H )(b , x , ), played according to Diagram 3.3.

Infinite runs where an i < which satisfies conditions (a)(c) has not been reached are
won by player I, see the definition of H (. . . ) above. We have a strategy Q [h ]
which is winning for II in H , see the successor stage in Section 3B (3), particularly
conditions (S1), (S2) and the paragraph following them.
Working by induction on i < we define y i and sequences Qi of ordinals, of
length i. We construct so that y i and Qi together form a position in H according
112 3 Games of continuously coded length

to ; so that Is ordinal moves in Qi are Silver indiscernibles; and so that Is natural


number moves in y i are the ones given by the imaginary opponent.
Suppose we never reach an i < which satisfies conditions (a)(c) of Section 3B (1).
Gcont (, C) then ends with mega-round . We must check that the end position is won
by player II, in other words we must check that $x , y %  C .
The sequences Qi are positions in Martins game (k j0, )(C)[x , y ]. Note that
the positions Qi cannot be shifted and combined to form an infinite run of
(k j0, )(C)[x , y ]. This is because infinite runs of H are won by player I, yet
is a winning strategy for player II. It follows from this that $x , y % does not belong to
C , as required. The argument is standard; let us only refer the reader to Fact 2D.3.
#

3C (2) Determinacy. To complete the proof that Gcont (, C) is determined, it is now


sufficient to verify that the hypotheses of Lemmas 3B.9 and 3C.2 cannot both fail.
Suppose they do. In other words suppose that I does not win Gini and II does not win
. We shall derive a contradiction. We work inside M[g] throughout. In particular,
Hini
when we say a small (relative to ) generic extension of M we mean a small extension
M[g] with g M[g].
Let L equal cardM (M& + + 1). Let H be the successor in M of L . This is
precisely jumpM ().
Claim 3C.3. Let x be a code belonging to a small generic extension of M. Let P
(M&)< and s < be given, with (s) defined and lh(s) lh(P ).
Then I wins Gmain (x, s, P , L ) iff I wins Gmain (x, s, P , H ). A similar claim holds
.
for Hmain
Proof. We prove the claim for Gmain . Notice that the only dependence of Gmain on
is through rule (1b) on page 91. Increasing in this rule simply adds more possible
moves for I. Its clear therefore that Gmain (. . . , ) becomes easier for I as increases.
Working in M[g] let U be the set of triples $x, s, P % so that x is a code belonging
to a small generic extension of M, P (M&)< , s < with (s) defined and
lh(s) lh(P ), and player I wins Gmain (x, s, P , ). From the discussion in the previous
paragraph it follows that U U  wherever <  . In other words $U | ON%
is -increasing. Using a Skolem hull and absoluteness argument one can check that
$U | ON% stabilizes before the successor of cardM (M& + ) in M. The claim
follows. #
Let gen [L ] be the generic strategies map associated to A[L ], , and X = M&.
Let gen [L ] be the mirrored generic strategies map associated to B[L ], , and M&.
Definition 3C.4. A foothold is a quadruplet x, s, P , Q so that:
(1) x is a code, for a position y | <  say, which belongs to a small, relative to
, generic extension of M;
(2) s < , (s) is defined, (s)  {(y ) | < }, and lh(s) (s);
3C First determinacy result, part II 113

(3) P is a position of (s) rounds in A[L , x], played according gen [L , x];
(4) player II wins Gmain (x, s, P , L );
(5) Q is a position of (s) rounds in B[L , x], played according gen [L , x];
(x, s, Q, ).
(6) player I wins Hmain L

We think of P and Q as sequences of length (s). Note that lh(P ) = (s) lh(s),
so the reference to Gmain (x, s, P , L ) in condition (4) makes sense. The reference to
(x, s, Q, ) in condition (6) makes sense similarly.
Hmain L
Condition (4) in Definition 3C.4 is an indication that $y | < % $s% should, at
least in some weak sense, be a winning position for II in Gcont (, C). Condition (6)
on the other hand is some indication that $y | < % $s% should be a winning
position for I. Our expectation is that these two conditions combined should lead to a
contradiction. We should be able to reach this contradiction by progressing from one
foothold to the next. A progression of this kind could never reach an end position in
Gcont (, C), since that would be a position which is won by both players. On the other
hand we shall see that the progression cannot go on forever.
Definition 3C.5. Let x  , s  , P  , Q and x, s, P , Q be two footholds, with x  equal to
y | <  , and x equal to y | <  say. x  , s  , P  , Q extends x, s, P , Q if:

(1)  > ;
(2) y = y for all < ;

(3) y extends s (hence (y ) = (s));


(4) (y ) (s) for all > ;

(5) P  and P agree. Precisely: if (s) (s  ) then P is an initial segment of P  ,


and if (s  ) (s) then P  is an initial segment of P ; and
(6) Q and Q agree.
Conditions (2), (3), and (4) together imply that x  (s) is equal to x(s). It follows
that A[L , x] and gen [L , x] agree with A[L , x  ] and gen [L , x  ] on the first (s)
rounds. Condition (5) therefore makes sense, and similarly with condition (6).
Claim 3C.6. Suppose that x, s, P , Q is a foothold. Say x = y | <  . Then there
are y , a , and b in M[g] so that:
(1) y is a real which extends s;
(2) a is an infinite run of A[L , x ], played according to gen [L , x ], where x =
$y | < %, y ;
(3) a extends P ;
114 3 Games of continuously coded length

(4) b is an infinite run of B[L , x ], played according to gen [L , x ]; and

(5) b extends Q.

Proof. We work in M[g] to construct y , a , and b maintaining the following condi-


tions for each m (s):

(i) y m is compatible with s;

(ii) a m is a position in A[L , x ] played according to gen [L , x ];

(iii) player II wins Gmain (x, max{s, y m}, a m, L );

(iv) b m is a position in B[L , x ] played according to gen [L , x ]; and

(v) player I wins Hmain  m, L ).


(x, max{s, y m}, b

We start with m = (s), setting y m = sm; a m = P ; and b m = Q.


Conditions (ii)(v) follow directly from conditions (3)(6) in Definition 3C.4, since x
and x agree to (s).
Suppose now that y m, a m, and b m are known, and that conditions (i)(v)
hold for m. Using condition (iii) and Claim 3C.3 fix a winning strategy m for II
in Gmain (x, max{s, y m}, a m, H ). Note the switch from L to H here. Working
(x, max{s, y m}, b
similarly with (v) fix a winning strategy m for I in Hmain  m, H ).
We construct moves in round m of Gmain (x, max{s, y m}, a m, H ), and in round
(x, max{s, y m}, b
m of Hmain  m, H ):

We play m = L in both games. This is a legal move since L < H .

m (if m is even) or m (if m is odd) plays y (m).



gen [L , x ] and gen [L , x ] respectively play amI
and bmII .

m and m respectively play amII .
and bmI

This completes the construction. The position given by a m + 1, y m + 1, and L


must be a winning position for II in Gmain (x, max{s, y m}, a m, H ), since m is
a winning strategy for II in that game. Condition (iii) for m + 1 follows from this
using Remark 3B.5. A similar argument gives condition (v). The other conditions are
immediate. #

Claim 3C.7. Suppose x , a , b M[g] are such that:

(1) x is a code, x = y | <  say;

(2) a is an infinite run of A[L , x ], played according to gen [L , x ];

(3) b is an infinite run of B[L , x ], played according to gen [L , x ].


3C First determinacy result, part II 115

Then there exists some s < so that:

(1) (s ) is defined, (s )  {(y ) | < }, and lh(s ) (s );

(2) player II wins (Gmain )(x , s , (P ), (L )) in N[g], where P denotes


a (s );
)(x , s , (Q ), ( )) in N[g], where Q denotes
(3) player I wins (Hmain L
b (s ).

In conditions (2) and (3), N denotes Ult(M, Eord(x ) ) and : M N denotes the
ultrapower embedding. Note then that x belongs to N[g].

Proof. Using Lemma 1B.2 and condition (2) in the hypothesis of the claim we see
that $a , x %  A[L ][g]. By Definition 3B.3, and since both a and x belong to
M[g], this means that I does not win G ( a , x , L ) in M[g]. Since G ( a , x , L ) is

a closed game, it must be that II wins. Fix then M[g], a winning strategy for II
in G (a , x , L ). Working similarly with condition (3) of the hypothesis of the claim
fix M[g], a winning strategy for I in H (b , x , L ).

G ( a , x , L ) and H (b , x , L ) are played according to the same rules exactly.


The only difference between the games is in the payoff when reaching an i < which
satisfies conditions (a)(c) in Section 3B(1). We can thus pit (playing for II in
Diagram 3.3) against (playing for I). They must at some point reach an i < which
satisfies conditions (a)(c) in Section 3B(1); for otherwise they play the exact same
game, with the exact same payoff, yet they both win.
Now both G ( a , x , L ) and H (b , x , L ) end at the first i which satisfies con-
ditions (a)(c). Let s = y i be the position reached when the games end. Let
P = a (s ) and let Q = b (s ).
Since is a winning strategy for II in G ( a , x , L ), we know that N[g]  |=

main ((A))[x , s , (P ), (L )]. In other words, player I does not win the game
(Gmain )(x , s , (P ), (L )) in N [g]. This game is clopen, hence determined in
N [g]. So it must be won by II.
A similar argument using the fact that is winning for I in H (b , x , L ) shows
that player I wins (Hmain )(x , s , (Q ), ( )) in N[g]. #
L

Claim 3C.7 can be thought of as producing s so that x , s , (P ), (Q ) is a


foothold in the sense of N. Let us now reflect this to obtain some x  so that x  , s , P , Q
is a foothold in the sense of M. We have to replace x , which only belongs to M[g],
by some approximation x  which belongs to a small generic extension of M. The next
claim does this. Conditions (2)(6) specify the ways in which x  approximates x .

Claim 3C.8. Continue with the terminology of Claim 3C.7. Let s be given by the
conclusion of that claim. Suppose that < (strictly) is such that $y | < %
belongs (not only to M[g] but also) to a small generic extension of M. Let j < be
given. Then there is code x  , for a sequence $y | <  % say, which satisfies:
116 3 Games of continuously coded length

(1) x  , s , P , Q is a foothold (in particular x  belongs to a small generic extension


of M);
(2)  > strictly;
(3) y = y for all < ;

(4) y j is equal to y j ;
(5) x  j = xj ; and
(6) {(y ) | <  } j is equal to {(y ) | < } j .
Proof. Increasing j if necessary we may assume that j (s ). Let r denote xj . Let
I = {(y ) | < } j . Let t = y j . Note that I , r, and t are finite sequences, and
so belong to M. Let [x  , P , Q, L , $y | < %] be the formula stating that:
(i) x  is a code for a position, $y | <  % say;

(ii) the sequence coded by x  extends $y | < % strictly;


(iii) y j = t;
(iv) x  j = r;
(v) {y | <  } j = I ;

(vi) {(y ) | <  } does not have (s ) as an element;

(vii) player II has a winning strategy in Gmain (x  , s , P , L ); and


(x  , s , Q , ).
(viii) player I has a winning strategy in Hmain L

By assumption $y | < % belongs to a small generic extension of M. Fix g


witnessing this. Note that : M N , the ultrapower embedding by Eord(x ) , extends
to an embedding of M[g] into N[g]. $y | < % is not moved by this embedding.
Now N [g] is a model of the statement
() There exists a cardinal < () so that after forcing with col(, ) over N
one obtains some x  so that [x  , (P ), (Q ), (L ), $y | < %] holds.
To see this take = and x  = x . The formula holds with this assignment:
Conditions (i)(v) hold directly because x  = x . Condition (vi) holds because of
condition (1) in the conclusion of Claim 3C.7. Conditions (vii) and (viii) hold because
of conditions (2) and (3) in the conclusion of Claim 3C.7.
Using the elementarity of the embedding : M[g] N[g] we may pull the state-
ment () back to M[g]. We obtain some < , some g  which is col(, )-generic/M,
and some x  M[g  ] so that [x  , P , Q , L , $y | < %] holds in M[g  ].
It is easy to see that x  , s , P , Q is a foothold. Conditions (1), (2), (4), and (6)
of Definition 3C.4 hold because of the definition of and, in the case of condition (2),
3C First determinacy result, part II 117

known facts about s . Conditions (3) and (5) hold because of known facts about P and
Q , more precisely because of conditions (2) and (3) in the hypothesis of Claim 3C.7.
Note here the importance of clause (iv) in the definition of , which implies that x and
x  agree to j (s ). #

Corollary 3C.9. Suppose x, s, P , Q is a foothold. Then there exists a foothold x  , s ,


P , Q which extends x, s, P , Q.

Proof. First apply Claim 3C.6. The claim produces a , b , and y which gives rise to
x = $y | < %, y . Now apply Claim 3C.7 with = + 1. Follow this with
an application of Claim 3C.8, using j = lh(s). (Note then that j (s).) It is easy to
see that the resulting foothold extends x, s, P , Q. #

Equipped with the previous claims we can begin our quest for a contradiction. We
know to begin with that both the hypothesis of Lemma 3B.9 and the hypothesis of
Lemma 3C.2 fail, in other words I does not win Gini and II does not win Hini
. We shall

use this knowledge to obtain a foothold. We shall then use the previous claims to extend
this foothold to the point of contradiction.

Claim 3C.10 (assuming that I does not win Gini and II does not win Hini
). There exists

a foothold.

Proof. Begin by constructing a , b M[g] so that:

(i) a is an infinite run of A[L , ], played according to gen [L , ]; and

(ii) b is an infinite run of B[L , ], played according to gen [L , ].

The construction is similar to the construction of y , a , and b in the proof of


Claim 3C.6, only using Gini and Hini instead of G
main and Hmain . We leave the details to
the reader, and only note that the construction is based on the condition that II wins Gini
and I wins Hini . This condition follows from the initial assumption for contradiction

since both games are clopen and hence determined.


Next apply Claim 3C.7 with x = . Let s < be given by that claim. Let
P = a (s ) and let Q = b (s ). Let N = Ult(M, Eord() ) and let be the

ultrapower embedding. The conclusion of Claim 3C.7 is such that:

(iii) player II wins (Gmain )(x , s , (P ), (L )); and


)(x , s , (Q ), ( )).
(iv) player I wins (Hmain L

Since x =  belongs to N, both games above in fact belong to N. Conditions (iii)


and (iv) in fact hold in N. Pulling these conditions down to M using the elementarity
of we get:

(v) player I wins Gmain (x , s , P , L ); and


(x , s , Q , ).
(vi) player II wins Hmain L
118 3 Games of continuously coded length

It is now easy to verify that x = , s , P , Q is a foothold. #


An n-foothold is a foothold x, s, P , Q so that (s) = n. Let n0 be least so that there
exists an n0 -foothold. Note the implicit use of Claim 3C.10 here. Let x0 , s0 , P0 , Q0 be
some n0 -foothold. Working inductively on k < let nk+1 be least so that there exists
an nk+1 -foothold which extends xk , sk , Pk , Qk . Note the implicit use of Corollary 3C.9
here. Let xk+1 , sk+1 , Pk+1 , Qk+1 be some nk+1 -foothold which extends xk , sk , Pk , Qk .
Claim 3C.11. Suppose k l. Then nk nl and xk+1 , sk+1 , Pk+1 , Qk+1 extends
xl , sl , Pl , Ql .
Proof. We work by induction on k. Fix l k < . We first show that nk nl . If
l = 0 this follows immediately from the minimality condition in the definition of n0 .
So suppose l > 0. By induction xk , sk , Pk , Qk extends xl1 , sl1 , Pl1 , Ql1 . By the
minimality condition in the definition of nl it follows immediately that nk nl .
Next let us show that xk+1 , sk+1 , Pk+1 , Qk+1 extends xl , sl , Pl , Ql . If l = k this
follows immediately from the definition of xk+1 , sk+1 , Pk+1 , Qk+1 . So suppose l < k.
We know by induction that:
(i) xk , sk , Pk , Qk extends xl , sl , Pl , Ql .
We know by definition that:
(ii) xk+1 , sk+1 , Pk+1 , Qk+1 extends xk , sk , Pk , Qk .
We know from the previous paragraph that:
(iii) nk nl .
Using conditions (i)(iii) it is easy to verify that xk+1 , xk+1 , Pk+1 , Qk+1 extends xl ,
sl , Pl , Ql . We refer the reader directly to Definition 3C.5, and only comment that
condition (iii) is used in the verification. #
The sequence coded by xk+1 extends the sequence coded by xk . Let $y | < %
be the union of the sequences coded by xk , k < . Intuitively $y | < % is a position
given by a maximal union of footholds, maximal in the sense that {nk | k < } is
made as dense as possible. We plan to get a contradiction by taking the construction
a step further, and extending this maximal sequence. This will ultimately produce a
natural number n which could have been used in {nk | k < }, but wasnt.
Let x = y | < . Let k be the length of the sequence coded by xk . We
have xk = y | < k . Note that (y ) nk for all k . This follows
from Claim 3C.11 and conditions (3) and (4) in Definition 3C.5. Using Claim 3A.4 it
follows that xk and x agree to nk . Pk is a position of nk rounds in A[L , xk ], played
according to gen [L , xk ]. Using the agreement between x and xk it follows that Pk is
also a position of nk rounds in A[L , x ], played according to gen [L , x ]. By similar
reasoning, Qk is a position of nk rounds in B[L , x ], played according to gen [L , x ].
Claim 3C.12. Suppose l < k. Then nl < nk strictly.
3D A slight improvement 119

Proof. We know already that nl nk . But nk = nl is impossible: By condition (2) in


Definition 3C.4, nk  {(y ) | < k }. In particular nk  = (yl ). But (yl ) = (sl )
by condition (3) in Definition 3C.5. So nk  = (sl ). #
Using Claims 3C.11 and 3C.12  we see that Pk strictly
 extends Pl for l < k < ,
and similarly with Q. Let a = k< Pk and let b = k< Qk . Then:
(a) a is an infinite run of A[L , x ], played according to gen [L , x ]; and

(b) b is an infinite run of B[L , x ], played according to gen [L , x ].


Remark 3C.13. Each xk belongs to a small generic extension of M. This follows from
condition (1) in Definition 3C.4. But we cannot expect the whole sequence $xk | k < %
to belong to a small generic extension of M. The reason is the implicit use of g in the
definition of $nk , xk , sk , Pk , Qk | k < %. The maps gen and gen are defined relative
to g, and so g is implicitly used already in the very definition of a foothold. We therefore
cannot expect $y | < % and x to belong to a small generic extension of M. They
need only belong to M[g].
Apply Claim 3C.7 to x , a , and b . (Note that all three belong to M[g], since
the entire construction takes place inside M[g].) Let s < be given by that
claim. Let n = (s ). Note that n  {(y ) | < } by Claim 3C.7. In particular
n  {nk | k < }. This is the first whiff of a contradiction. n should have been put into
{nk | k < } by the minimality conditions in the construction of $nk , xk , sk , Pk , Qk |
k < %.
Let us make the contradiction precise. Let P = a n and let Q = b n . Let k <
be large enough that n < nk+1 . Such a k exists by Claim 3C.12. Apply Claim 3C.8
with = k and j = lh(sk ). The result is some code x  so that x  , s , P , Q is a
foothold, and so that x  , s , P , Q extends xk , sk , Pk , Qk . But n = (s ) is smaller
than nk+1 . This contradicts the minimality condition in the definition of nk+1 .
The contradiction was reached under the assumption that both the hypothesis of
Lemma 3B.9 and the hypothesis of Lemma 3C.2 fail (see Claim 3C.10). We can thus
conclude that one of these hypotheses holds true. Using the corresponding lemma it
follows that Gcont (, C) is determined.

3D A slight improvement
In the previous sections we proved the determinacy of Gcont (, C) for C in the pointclass
<2 11 and a continuous . Here we improve the result slightly. We use the same
large cardinal assumption to handle in the class  02 . Precisely we prove:
Theorem 3D.1. Suppose that there exists a mildly iterable class model M with cardinals
< so that (a) is a Woodin cardinal of M; (b) is strong to + 1 in M; and
(c) M& + 1 is countable in V. Then the games Gcont (, C) are determined for all in
the class  02 and all C in the pointclass <2 11 .
120 3 Games of continuously coded length

It is not clear if this result is optimal. The conclusion could conceivably be strength-
ened to handle in classes above  02 . At the moment it is only known that it cannot be
strengthened beyond  06 , see Neeman [35].
Theorem 3D.1 is the beginning of a hierarchy. For countable 1 we say that
is strong to + in M if for every Z M& + there exists some extender E M
with crit(E) = and Z Ult(M, E). The following result generalizes Theorem 3D.1:

Theorem 3D.2 (for countable 1). Suppose that there exists a mildly iterable class
model M with cardinals < so that (a) is a Woodin cardinal of M; (b) is strong
to + in M; and (c) M& + is countable in V. Then the games Gcont (, C) are
determined for all in the class  01+ and all C in the pointclass <2 11 .

In both theorems the assumptions can be obtained from large cardinals in V using
the methods described at the end of Appendix A.

3D (1)  02 functions. Let us prove Theorem 3D.1. Fix M, , and which satisfy the
large cardinal assumptions listed in Section 3B. Fix a sequence E = $E | < lh(E)% 
as in Section 3B (1). We intend to imitate the determinacy proof of Sections 3B and 3C,
but working now with : R in the class  02 . Fix such a function . As usual we
may assume that a real coding belongs to M.
Work in M[g] where g is col(, )-generic/M. For each n < the pre-image of
{n} under is  02 . Fix a list of closed sets {Ci | i < } large enough to generate
1 {n} forall n. By this we mean that for each n < there is some In so that
1 {n} = iIn Ci . We shall convert into a continuous function by unraveling the
closed sets.
We refer the reader to Martin [25] for the exact definitions of covers which unravel
given sets. Here we shall quote only facts we need, and not even all of these.
Let K0 be the tree < . Construct $Ki+1 , fi+1 , i+1 % so that for each i < ,
$Ki+1 , fi+1 , i+1 % is an i-covering of Ki which unravels fi1 f11 (Ci ). This
can be done in such a way that:

Fact 3D.3. Nodes of length at least 2i in the tree Ki are sequences of the form
$d0 , . . . , dl1 % where d0 , . . . , d2i1 are real numbers and d2i , . . . , dl1 are natural num-
bers.

For a node s Ki of length 2i let Ki (s) denote the tree of sequences t Ki which
extend s, or equal an initial segment of s. Let fi,0 denote f1 fi . We alternate
between thinking of fi,0 as a Lipschitz continuous map from [Ki ] into , and as a
map from Ki into < . From Fact 3D.3 we directly get the following claim:

Claim 3D.4 (for nodes s Ki of length 2i). s, Ki (s), and fi,0 Ki (s) are all coded by
real numbers. #

From Claim 3D.4 and the strength assumptions on the sequence of extenders E (see
Section 3B (1)) we get the following corollary:
3D A slight improvement 121

Corollary 3D.5. For s Ki a node of length 2i, there exists some < lh(E)  so
that s, Ki (s), and fi,0 Ki (s) belong to Ult(M, E )[g]. We use ord(s) to denote the
least such . #

For each i < and each j < i, fi,0 1 (Cj ) is clopen (see Martin [25]). In fact
for w [Ki ], knowledge of w2(j + 1) suffices to determine whether w belongs to
fi,0 1 (Cj ). This, the fact that Kj +1 and Ki agree on nodes of length 2(j + 1), and the
fact that fi,j +1 is the identity on such nodes, allow us to pick sets Sj for j < so that:

Fact 3D.6. Sj is a set of nodes of length 2(j + 1) in Kj +1 . For every j < i < and
every w [Ki ], w fi,0 1 (Cj ) iff w2(j + 1) Sj .

Let K be the inverse limit of the trees Ki . Let f,i be the inverse limit embeddings.
We alternate between thinking of f,i as a Lipschitz continuous map from [K ] into
[Ki ], and as a map from K into Ki . We have the standard commutativity: f,0 =
fi,0 f,i .
Fact 3D.7. K and Ki have the same nodes of length 2i. f,i is the identity on nodes
of length 2i.

Define a partial map : K by setting:


Definition 3D.8. (s) = n iff there is some j so that s2(j + 1) Sj and j In .
(Recall that In are such that for each n, 1 {n} = j In Cj .)

The following claims can be deduced easily from the previous facts:

Claim 3D.9. For each w [K ], ( f,0 )(w) = n just in case that (wi) = n for
all sufficiently large i.

Proof. Note that ( f,0 )(w) = n iff w belongs to f,0 1 (Cj ) for some j In . By
Fact 3D.6, w belongs to f,0 1 (Cj ) iff w2(j + 1) Sj . #

Claim 3D.10. Suppose that w [Ki ], and that (w2i) is defined and equal to n. Then
( fi,0 )(w) = n. #
We can now imitate the determinacy proof of Sections 3B and 3C, only replacing
the function defined above for the one used in those sections. We briefly sketch the
argument.
Fix a payoff set C <1 which is <2 11 in the codes. We aim to show that
Gcont (, C) is determined by imitating the proof in Sections 3B and 3C.
We begin by adapting the definition of the games Gmain (. . . ) of Section 3B (1). Let
M[g] be a small generic extension of M. Let x = y | <  be a code belonging
to M[g]. Let P (M&)< . Let n = lh(P ). Let s be some finite sequence in M[g].
For simplicity assume that lh(s) n. Let K M[g] be a tree which contains s as a
node. Let f M[g] be a Lipschitz embedding from [K] into . We also think of f
as a map from K into < . Suppose that:
122 3 Games of continuously coded length

() For every w [K] which extends s, ( f )(w) is equal to n.

Let s denote the triplet $s, K, f %. We define a game Gmain (x, s, P , ). The game
is played inside M[g] according to Diagram 3.14 and rules (1)(3) below.


I n anI n+1 an+1I
w (n) w (n + 1)
II anII an+1II

Diagram 3.14. The game Gmain (x, s, P , ).

We use m to index rounds in Gmain (x, s, P , ). The game starts with m = lh(P ).
, a
The objects played in round m are m , amI
mII , and w (m). We use Pn to de-

note P , and inductively define Pm+1 = Pm , amI , amII . We let w n = sn,
and inductively define w m + 1 = w m, w (m). Rule (2) below guarantees
that w belongs to [K] and extends s. We let y = f (w ). We let x denote
$y | < %, y .

(1) (Rule for I) m is an ordinal chosen so that:

(a) Pm is a legal position in A[m , x ];


(b) m < if m = n; and

(c) m < m1 if m > n.

(2) (Rule for I and II) w (m), played by I if m is even and by II if m is odd, is subject
to the restriction that w m, w (m) is a node in K. For m < lh(s) we require
further that w (m) = s(m).

(3) (Rule for I and II) amI
and amII are legal moves in the game A[m , x ], fol-

lowing the position Pm .

The first player to violate any of the rules of Gmain (x, s, P , ) loses. Because of
rule (1c) there are no infinite runs of this game.
The reader should compare the definition here with the definition given in
Section 3B (1). The natural number moves y (m) of Section 3B (1) are replaced here
by moves on the tree K. Moves on K are translated to natural numbers via the Lipschitz
map f .
Note how the assumption () above replaces the assumption in Section 3B (1) that
(s) is defined. The current assumption has a similar effect. It guarantees that (y )
is known from the starthere it must in fact equal nand does not depend on the
moves w (m). This implies that x depends in a Lipschitz manner on y , which in
turn depends in a Lipschitz manner on w . To sum, knowledge of w m suffices to
determine x m. Rules (1a) and (3) therefore make sense.
3D A slight improvement 123

Continuing the parallel with Section 3B (1) let us define the games G (
a , x, ). We
work with a (M&) M[g], a code x = y | <  in M[g], and an ordinal .

G (
a , x, ) is played inside M[g] according to Diagram 3.15.

I w (0), 0 w (2), 2 ...


II w (1), 1 w (3), 3 ...

Diagram 3.15. G (
a , x, ), so long at (w i) is not defined or belongs to {n | < }.

Players I and II collaborate to produce an infinite branch w [K ]. K here


is the covering tree constructed above. Player I plays w (m) for even m < , and
player II plays w (m) for odd m < . We set y = f,0 (w ). In addition to playing
w , I and II play auxiliary moves in the game C[x, y ]. (Note that f,0 is Lipschitz, so
knowledge of w m suffices to determine y m. Knowledge of y m in turn suffices
to determine the rules of the first m rounds of C[x , y ].) They continue this way until,
if ever, reaching an i < so that:
(a) (w 2i) is defined;
(b) (w 2i)  {n | < }; and
(c) 2i (w 2i).
The map : K used here is the one given by Definition 3D.8.
The first player to violate any of the rules of C[x, y ] loses. Infinite runs where an
i < satisfying conditions (a)(c) was not reached are won by player I.
If an i < satisfying conditions (a)(c) is reached, the game ends. We set s =
w 2i, n = (s), and P = a n. We let N = Ult(M, Eord(s) ) and let be the
ultrapower map. (See Corollary 3D.5 for the definition of ord(s).) We let K = Ki (s)
and let f = fi,0 Ki (s). Set s = $s, K, f %. I wins just in case that N[g] |=Player I
has a winning strategy in (Gmain )(x, s, (P ), ( )).
Remark 3D.11. Both Corollary 3D.5 and Claim 3D.10 are needed to make sense of the
use of (Gmain )(x, s, (P ), ( )) in the payoff condition. Corollary 3D.5 implies that
s belongs to N [g]. Claim 3D.10 implies that the condition () preceding the definition
of Gmain (. . . ) holds for s.
The reader should compare the game G (x, a , ) described here with the game
described in Section 3B(1). Here instead of directly playing the real y , I and II play
on the tree K . This allows use of the function of Definition 3D.8. By passing to the
cover K we have essentially converted the  02 function to the continuous function
given by .
This is a good place to end our brief sketch of the proof of Theorem 3D.1. The
rest of the proof is a matter of following the argument of Sections 3B and 3C, using
the revised games described above, and using the techniques of Martin [25] to convert
strategies on K or Ki to strategies on natural numbers.
124 3 Games of continuously coded length

Exercise 3D.12. Complete the proof of Theorem 3D.1.

3D (2)  01+ functions. The proof of Theorem 3D.2 is similar to the proof of Theo-
rem 3D.1. But with Theorem 3D.2 one has to unravel sets which reside higher up in
the Borel hierarchy. One therefore has to allow for covers which reside higher up in the
von-Neumann hierarchy; Fact 3D.3 no longer holds when the sets being unraveled are
more complicated than closed.
Fact 3D.3 led us to Claim 3D.4, which in turn combined with our large cardinal
strength assumption to give Corollary 3D.5. Corollary 3D.5 was essential to the defi-
nitions in Section 3D(1), see specifically Remark 3D.11.
To maintain the validity of Corollary 3D.5 when dealing with covers of greater size,
we simply have to increase our large cardinal assumption. We have to make sure that
for each of the trees Ki in the covers used, and for each node s Ki of length 2i, there
is some extender E M so that:

(1) crit(E) = ;

(2) E is at least -strong; and

(3) s, Ki (s), and fi,0 Ki (s) all belong to Ult(M, E)[g].

So long as our large cardinal assumption is strong enough to secure these conditions,
we can run the proof described in Section 3D(1) essentially unchanged.
Condition (3) is the one which relates the necessary large cardinal strength to the
complexity of . Readers familiar with Martin [25] can easily compute the size of the
trees Ki needed to handle a function in a given Borel class, and thereby connect the
complexity of with a large cardinal assumption on M. We leave this computation to
those readers and only note that the end result is Theorem 3D.2.

3E Variation
In the spirit of the previous chapter we develop here a mechanism which reduces games
of continuously coded length to iteration games. This mechanism will allow us to
compose games of continuously length with games of the kind handled in Chapter 2,
and to prove determinacy for games with universally Baire payoff.
Let : R be given. By position we mean a -position and by run we mean a
-run, see Section 3A.
Let M be a ZFC model with a Woodin cardinal . Let A be a name for a subset of
R R in M col(, ) . We define a game G cont = G
cont (M, , , A) in which players I
and II collaborate to construct:

a run $y | %;
3E Variation 125

a sequence of models of the form presented in Diagram 3.16, starting with


M0 = M; and

embeddings j, : M M for , k : M Q for , and


k : Q M , of the form presented in Diagram 3.16.

j0,1 j1,2

kk
SS kkk F0 # kk
SS kkk F1 #
M0 k SSSS k / Q0 M1 k SSSS k / Q1 M2
0 1
T0 T1

j, +1

kkk #
kSkSkSkS
kk
/ Q kSkSkSkS
kk
Sk
Sk F
M k SSSS k / Q M +1 M
SS k SS k
/ M

T T T

cont .
Diagram 3.16. The iterates in G

For limit we let M be the direct limit of the models M for < . M0 is set
equal to M. Mega-round , when is a limit or 0, is played as follows:

(L1) Player I plays a length iteration tree T on M .

(L2) Player II plays a cofinal branch b through T .

This completes mega-round . We let Q be the direct limit along b , and let k be the
corresponding embedding.
At the start of a successor mega-round + 1 we know already the iterate Q .
Mega-round + 1 proceeds as follows:

(S1) Players I and II collaborate in the usual fashion to produce the real y .

If the real y produced following rule (S1) satisfies the condition () (y ) is defined
and (y )  {(y ) | < }, then mega-round + 1 continues according to rules
(S2)(S4).

(S2) Player I plays an extender F Q with critical point smaller than crit(k ). We
set M+1 = Ult(M , F ) and let j,+1 be the ultrapower embedding.

(S3) Player I plays a length iteration tree T+1 on M+1 .

(S4) Player II plays a cofinal branch b+1 through T+1 .


126 3 Games of continuously coded length

This completes mega-round + 1. We let Q+1 be the direct limit along b+1 , and
let k+1 be the direct limit embedding. We are then in a position to begin mega-round
+ 2.
If the real y produced following rule (S1) does not satisfy the condition () above,
then mega-round + 1 continues with rules (F1) and (F2) instead of rules (S2)(S4).
Note that mega-round + 1 finishes the game in this case.
(F1) Player I plays a length iteration tree T on Q .
(F2) Player II plays a cofinal branch b through T .
We let M be the direct limit along b , and let k be the direct limit embedding. We
let j0, : M0 M be equal to k k j0, . By the failure of condition () we
know that $y | % is a runa terminal stage in the game of continuously coded
length associated to . G cont (M, , , A) ends at this point. If M or any of the
models produced along the way is illfounded, then player I wins. Otherwise player I
wins just in case that there exists some h so that:
(1) h is col(, j0, ( ))-generic/M ; and
(2) $x , y % j0, (A)[h], where x = y | < .
The above rules complete the description of the game G cont (M, , , A). As usual
we define a mirrored version of this game. Let B be a name for a subset of R R in
cont (M, , , B) is played according to the rules listed above except
M col(, ) . H
that it is now player II who plays the extenders in rule (S2) and the trees in rules (L1),
(S3) and (F1); while I is the player responsible for branches in rules (L2), (S4) and
(F2). It is now player II who wins if any of the models produced is illfounded. If all
the models produced are wellfounded, then player II wins just in case that there exists
an h which satisfies condition (1) above and condition (2) with A replaced by B.
cont (M, , , A) should be viewed as a reduction mechanism
A strategy for I in G
which converts an iteration strategy for M into a strategy for player I in the game of
continuously coded length associated to . A strategy for II in H cont (M, , , B)
should similarly be viewed as a reduction mechanism which converts an iteration strat-
egy for M into a strategy for player II in the game of continuously coded length. The
reader should compare this to Section 2A. The next theorem should be compared to
Theorem 2A.2.
Theorem 3E.1. Suppose that there are < < so that (a) and are Woodin
cardinals of M; and (b) is strong to + 1 in M. Suppose that belongs to the class
 02 and that (a code for) belongs to M. Suppose that M& + 1 is countable in V.
Fix g V which is col(, )-generic/M. At least one of the following three cases
holds:
cont (M, , , A);
(1) player I has a winning strategy in G
cont (M, , , B); or
(2) player II has a winning strategy in H
3E Variation 127

(3) in M[g ] there is a run $y | % so that $y | < , y % belongs to


neither A[g] nor B[g].
Moreover M can distinguish this. To be precise, there are formulae I and II so that:
if M |= I [ , , A] then case (1) holds; if M |= II [ , , B] then case (2) holds;
and otherwise case (3) holds.
We sketch a proof of Theorem 3E.1 in Section 3E (1). Let us here notice a few
corollaries.
Corollary 3E.2. Suppose that there exists a mildly iterable class model M with car-
dinals < 0 < 1 so that (a) 0 and 1 are Woodin cardinals of M; (b) is strong
to 0 + 1 in M; and (c) M&1 + 1 is countable in V. Then the games Gcont (, C) are
determined for all in the class  02 and all C in the pointclass  12 .
Proof. Fix and C. Let C witness that C is  12 in the codes. Passing to a generic
extension of an iterate of M if needed, we may assume that a code for exists in M
and that C is 21 (z) for some real z which exists in M. Let A M be a col(, 1 )
name for the set of pairs $x, y% so that x is a code for a -position, y is a real, and $x, y%
satisfies the 21 (z) statement which defines C . Let B name the set of pairs which do
not satisfy that 21 (z) statement.
Apply Theorem 3E.1 with = 1 . Suppose case (1) holds. Let  be a winning

strategy for player I in Gcont (M, 1 , , A). Let  be a mild iteration strategy for M.
and  combine in the natural way to produce a strategy for player I in Gcont (, C).

It is easy to check that this is a winning strategy for player I. This uses 21 absolute-
ness, condition (2) in the payoff of G cont , and the fact that -iterates of M are always
wellfounded.
If case (2) of Theorem 3E.1 holds, a similar argument shows that player II has a
winning strategy in Gcont (, C). Case (3) of Theorem 3E.1 is impossible, since A and
B name complementary sets. #
Corollary 3E.3. Suppose that < 0 < 1 are such that (a) 0 and 1 are Woodin
cardinals in V; and (b) is strong to 0 + 1, again in V. Let be  02 and let C be
1 -universally Baire in the codes. Then Gcont (, C) is determined.
Proof sketch. This is simply a matter of applying the methods of Section 2E, more
specifically Exercise 2E.11, to the current context. The argument involves referring
to Theorem 3E.1 where the solution to Exercise 2E.11 referred to Theorem 2A.2, and
using Claim 14 in Appendix A in addition to Theorem 12. #
Corollary 3E.4. Let cont be the game quantifier associated to the class of games Gcont .
Suppose that < 0 < 1 satisfy the assumptions in Corollary 3E.3. Let be  02 and
let C be 1 -universally Baire in the codes. Then cont (, C) is -universally Baire for
each < .
Proof sketch. Refer to Exercises 2E.9 and 2E.12, and apply the arguments there to the
current context. #
128 3 Games of continuously coded length

Let be a countable ordinal. Let C be a subset of R1+ . Gcont1+ (, C ) is


played as follows: Players I and II first follow the game of continuously coded length
associated to , and then continue playing 1 + additional reals. More precisely
players I and II collaborate as usual to produce a sequence of reals $y | < + %
so that (a) for each < , (y ) is defined and is not an element of {(y ) | < };
and (b) (y ) is not defined, or else it is an element of {(y ) | < }. The game is
won by player I iff y | < , $y+ | < % belongs to C .
Combining Theorem 3E.1 with the results connected to Theorem 2A.2 one easily
sees that:

Corollary 3E.5 (for a countable ordinal 1). Suppose that there exists a mildly
iterable class model M, an increasing sequence $ | < %, and < 0 so
that (a) each is a Woodin cardinals of M; (b) is strong to 0 + 1 in M; and
(c) M& sup{ + 1 | < } is countable in V. Then the games Gcont1+ (, C ) are
determined for all in the class  02 and all C in the pointclass <2 11 . #

Similar combinations are possible also with the current results, namely with the
results of this chapter. For example one can combine Theorem 3E.1 with itself to
obtain the determinacy of games of the kind Gcont+cont . (The rules of Gcont+cont are
phrased in the obvious way.)
While we are on the subject of variations let us also mention the following result,
on games where the payoff is homogeneously Suslin (see MartinSteel [18, 2] for
the definition). The result has nothing to do with the methods of the current section.
Rather it involves an adaptation of the argument in Sections 3B and 3C. We give it as
an exercise, meant for readers who are familiar with homogeneously Suslin trees and
their uses in proofs of determinacy.

Exercise 3E.6. Suppose that < are such that (a) is a Woodin cardinal in V; and
(b) is strong to + 1, again in V. Let be  02 and let C be ( + 1)-homogeneously
Suslin in the codes. Prove that Gcont (, C) is determined.

Hint. Let C witness that C is ( + 1)-homogeneously Suslin in the codes. Precisely


this means that C is a subset of R R; $ y , y % C iff y C for any run

y = $y | < + 1%; and C is ( + 1)-homogeneously Suslin.
Let the tree T and the countable collection of measures  witness that C is ( + 1)-
homogeneously Suslin. (In particular the projection of T is equal to C .) Let V be a
rank initial segment of V, large enough that , T , and  all belong to V . Let H be a
countable elementary substructure of V , with , , T , and  all in H . Let M be the
transitive collapse of H , let : M V be the anti-collapse embedding, and let , ,
S, and be such that () = , () = , (S) = T , and () = . 
Let D be the map which assigns to each pair $x, y% R R the following game
D[x, y]: player I plays a sequence u(0), u(1), . . . so that $x, y, u% [S], thereby wit-
nessing that $x, y% belongs to the projection of S, and player II does nothing. (This game
is taken from the determinacy proof for homogeneously Suslin sets, see for example
[18, Theorem 2.3].)
3E Variation 129

Now follow the argument in Sections 3B and 3C, but modify it to use D instead of
Martins games map C.
First modify the definitions in Sections 3B (1) and 3C (1), to use D in the games
G (. . . ) and H (. . . ), instead of C.
You will then have to modify the relevant part of the successor stage in Section 3B (3),
to use D. On the one hand the argument becomes simpler, since only player I has
moves in D[. . . ]. But on the other hand there is an extra detail: Your argument will
show that the code $x , y % belongs to the projection of (k j0, )(S). Through uses
of Theorem 12 and Claim 14 in Appendix A you should be able to make sure that there
is an embedding : Q V so that k j0, = . Use this to conclude that
$x , y % belongs to the projection of T .
Finally you will have to modify the argument made in the proof of Lemma 3C.2,
to use the measures (k j0, )() which witness the homogeneity of (k j0, )(S) in
Q , instead of Silver indiscernibles for Q . Your argument will show that the tower
of measures associated $x , y % is illfounded over Q . Through uses of Theorem 12
and Claim 14 in Appendix A you should again be able to make sure that there is an
embedding : Q V so that k j0, = . Use and the illfoundedness of the
tower associated to $x , y % over Q to conclude that the tower associated to $x , y %
over V is also illfounded, and therefore $x , y % does not belong to the projection of T .
#

3E (1) A sketch of the proof. Let M be a model of ZFC . Let be a Woodin cardinal
of M, and suppose that M& + 1 is countable in V. Let A and B be two names in
M col(, ) for subsets of R R. We work to prove Theorem 3E.1 for these objects.
Let < < be cardinals of M so that is a Woodin cardinal, and is strong to
+ 1 in M. Our tactic is to run the argument of Sections 3B and 3Cor Section 3D (1)
rather, when is  02 and not continuousbut replacing Martins game C with the
auxiliary games maps associated to the names A and B.
Let A name (M& ) A in M col(, ) . Define B similarly. Let A =
(x, y  A [x, y]) be the auxiliary games map associated to A , , and X = M&.
Let B = (x, y  B [x, y]) be the mirrored auxiliary games map associated to B ,
, and X = M&.

Remark 3E.7. The use of X = M& in these definitions is meant to ensure that the
corresponding pivots later on reside above . Note the implicit use here of the fact that
< .

For simplicity suppose that is continuous. We shall follow the argument of Sec-
tions 3B and 3C. The case of in the pointclass  02 is similar, following the modifica-
tions of Section 3D(1).
Define Gmain (. . . , ), G (. . . , ), the names A[ ], and Gini following Sec-
tion 3B (1) and Section 3B(3), except that the definition of G ( a , x, ) is modified.
G (
a , x, ) is now played according to Diagram 3.17. Players I and II collaborate
as usual to produce a real y = $y (i) | i < %. In addition they play auxiliary moves
130 3 Games of continuously coded length

I y (0)
a0I
a1I y (2) ...
II
a0II y (1)
a1II ...

Diagram 3.17. Modified G (


a , x, ), so long at (y i) is not defined or belongs to {n | < }.

in the game A [x, y ]. They continue this way until, if ever, reaching an i < so
that:
(a) (y i) is defined;
(b) (y i)  {n | < }; and
(c) i (y i).
The first player to violate any of the rules of A [x, y ] loses. If an i < satisfying
conditions (a)(c) is reached, the game ends. As in Section 3B(1) we set s = y i and
P = a (s). We let N = Ult(M, Eord(x) ) and let be the ultrapower map. Player I
wins iff N [g] |= main ((A))[x, s, (P ), ( )]. Other than that I is the open player;
infinite runs where an i < satisfying conditions (a)(c) was not reached are won by
player II.
The definition here differs from that in Section 3B (1) in two ways. First, the map
A is used instead of the map C. Secondly, infinite runs of G (. . . , ) are now won
by player II.
Lemma 3E.8. Suppose that I wins G . Then I wins G cont (M, , , A).
ini
Proof sketch. Fix an imaginary opponent willing to play for II in the game
G cont (M, , , A). We work with the imaginary opponent to construct a run of this
game. The construction is a variant of the argument in Section 3B(3).
The use of the iteration strategy in Section 3B (3) is replaced here by a use of the
imaginary opponent. Note how the structure of the iteration trees in the argument of
Section 3B (3) corresponds precisely to the structure of the trees in Diagram 3.16 up
to Q . The moves supplied by the imaginary opponent under rules (L2) and (S4) are
therefore an adequate replacement for the branches supplied by the iteration strategy in
Section 3B (3).
Other than that only the successor stage in Section 3B (3) requires modification. So
suppose we are at a successor stage + 1. We have the model Q and the embedding
k j0, : M Q . We have x coding the current position $y | < %. We have
a and given by the construction so far, see the successor stage in Section 3B (3).
We know that conditions (S1) and (S2) of Section 3B (3) hold true. Thus player I wins
the game G = (k j0, )(G )( a , x , ). Let be a strategy witnessing this. We

may pick in the model Q [h ]. Recall that h is generic for col(, (k j0, )()).
Let Apiv and piv be the pivot games and strategies maps associated to A ,
, and X = M&. In Section 3B(3) we used the method of ascribing indiscernibles
to form a run of G played according to . Here we use piv instead.
3E Variation 131

Using the initial assumption that M& + 1 is countable in V, it is easy to check


that Q &(k j0, )( ) + 1 is countable in V. Fix in V some surjection  :
Q &(k j0, )( ) + 1.
We work to form a real y and a run a, , T, of the pivot game
(k j0, )(Apiv )[x , y ]. The construction proceeds along the usual lines:

(P1) the imaginary opponent produces y (i) for odd i;


,
(P2) (k j0, )(piv )[, x , y ] produces the moves aiII for all i, and the extenders
which give rise to T, ;

(P3) and its shifts along the even branch of T, produce y (i) for even i, and the
,
moves aiI for all i.

Remark 3E.9. As usual the construction requires shifting along the even branch of
T, . It is important for this shifting that belongs to a generic extension of Q of
size (k j0, )(), and that the extenders used in T, all have critical points above
(k j0, )(). The latter fact traces back to Remark 3E.7, and to Lemma 1C.7.
If an i < which satisfies conditions (a)(c) in the definition of G (. . . ) is reached,
we must end the construction; the shift of G ends at the first such i, and the shift of
ceases to supply moves.
We know in this case that condition (S3) in Section 3B (3) holds true. (Note the use
of the fact that the extenders in T, have critical points above (k j0, )(). Initially
we only know that a version of condition (S3) shifted to the 2i-th model of T, holds
true. But the critical point involved in the shift is high enough that this can be pulled
back to Q .) Once we have condition (S3) we can simply return to the construction of
Section 3B (3).
So suppose that an i < which satisfies conditions (a)(c) in the definition of

G (. . . ) is not reached. Note that + 1 is the final mega-round of G(M, , , A) in
this case. We have so far produced the real y . To complete this final mega-round we
must follow rules (F1) and (F2). We must then verify that the run constructed is won
by player I. This will complete the proof of Lemma 3E.8.
The construction in items (P1)(P3) produces an infinite run a, , T, of
(k j0, )(Apiv )[x , y ], played according to (k j0, )(piv )[, x , y ].
Play T, as the move T for player I in rule (F1). The imaginary opponent,
playing according to rule (F2), produces a cofinal branch b through this tree. Let
M be the direct limit along this branch, and let k : Q M be the direct limit
embedding.
If M or any other model along the way is illfounded, then player I wins and our
task is complete. So suppose that M and all the models along the way are wellfounded.
Now is a strategy for player I, the open player, in G . There are thus no infinite
runs according to . But a, is an infinite run according to the image of along the
even branch of T . It follows that the even branch leads to an illfounded direct limit.
b must therefore be an odd branch.
132 3 Games of continuously coded length

I y (0)
b0I
b1I y (2) ...
II
b0II y (1)
b1II ...

Diagram 3.18. Modified H (b,


 x, ), so long at (y i) is not defined or belongs to {n | < }.

a, , T, is a (k j0, )(A )-pivot for $x , y % by Lemma 1C.5. By Defi-


nition 1C.1 and since b is an odd branch it follows that there must exist an h so
that:
(1) h is col(, (k k j0, )( ))-generic/M ; and
(2) $
a, , x , y % (k k j0, )(A )[h].
The definition of A is such that condition (2) can be re-written as:
(2 ) $x , y % (k k j0, )(A)[h].

This precisely is the winning condition in G(M, , , A). #
(. . . , ), H (. . . , ), the names B[ ], and H by mirroring the def-
Define Hmain ini
initions given before Lemma 3E.8. The mirrored games H (. . . , ) use the map B .
Player II is the open player in these games. Other than that the mirroring process is
similar to the one in Section 3C(1). By mirroring the proof of Lemma 3E.8 we get:
Lemma 3E.10. Suppose that II wins Hini cont (M, , , B).
. Then II wins H #
Lemmas 3E.8 and 3E.10 cover cases (1) and (2) in Theorem 3E.1. Let us set the
ground for a proof of case (3) when the hypotheses of these lemmas fail. We intend in
this case to follow a modification of the argument in Section 3C (2).
Let L and H be as in Section 3C(2). Fix g V which is col(, )-generic/M.
Fix g M[g ] which is col(, )-generic/M. We work below with reference to these
fixed objects.
Recall that the definition of Gmain (. . . , ), modified to the current settings from
Section 3B (1), involves the definition of col(, ) names A[ ]. Recall further that
A[ ] is the auxiliary games map associated to A[ ], , and M&. The definition
(. . . , ) similarly involves names B[ ] and the associated mirrored auxiliary
of Hmain
games maps B[ ]. Let gen [L ] be the generic strategies map associated to A[L ], ,
and M&. Similarly let gen [L ] be the mirrored generic strategies map associated to
B[L ], , and M&. Note that both gen [L ] and gen [L ] belong to M[g].
Let gen be the generic strategies map associated to A , , and X = M&.
gen is defined using the generic g , and belongs to M[g ]. Let gen M[g ]
be the mirrored generic strategies map associated to B , , and X = M&.
The next claim is an adaptation of Claim 3C.7 to the current definitions of A and B,
definitions which rest on the modified games G (. . . ) and H (. . . ) described above.
Conditions (1)(3) in the conclusion of Claim 3C.7 are replicated here in conditions
(B1)(B3). But here there is an extra clause, clause (A) in the conclusion of Claim 3E.11.
3E Variation 133

Claim 3E.11. Suppose x , a , b M[g] are such that:

(1) x is a code, x = y | <  say;

(2) a is an infinite run of A[L , x ], played according to gen [L , x ];

(3) b is an infinite run of B[L , x ], played according to gen [L , x ].

Then either (A) there exists some y M[g ] so that $y | < % is a complete run
of the game of continuously coded length associated to and so that $x , y % belongs
to neither A[g ] nor B[g ]; or else there exists some s < so that:

(B1) (s ) is defined, (s )  {(y ) | < }, and lh(s ) (s );

(B2) player II wins (Gmain )(x , s , (P ), (L )) in N[g], where P denotes


a (s );
)(x , s , (Q ), ( )) in N[g], where Q denotes
(B3) player I wins (Hmain L
b (s ).

In conditions (B2) and (B3), N denotes Ult(M, Eord(x ) ) and : M N denotes the
ultrapower embedding. Note then that x belongs to N[g].

Proof. We adapt the proof of Claim 3C.7 to the current situation.


By condition (2) and Lemma 1B.2, $ a , x %  A[L ][g]. Since both a and
x belong to M[g], this means that I does not win G (
a , x , L ) in M[g]. Since

G ( a , x , L ) is an open game, it must be that II wins. Fix then M[g], a winning

strategy for II in G (a , x , L ). Working similarly with condition (3) fix M[g],
a winning strategy for I in H (b , x , L ).
As a side remark let us note that II is the closed player in G ( a , x , L ), and that
I is the closed player in H (b , x , L ). and are thus strategies for the closed
players in their respective games.
Working inside M[g ] we construct a real y , an infinite run a of A [x , y ],
and an infinite run b of B [x , y ]. We construct these objects by combining moves
given by gen , , gen , and :
for all i;
gen [x , y ] produces the auxiliary moves aiI
for all i and the natural number moves
produces the auxiliary moves aiII
y (i) for odd i;
for all i; and
gen [x , y ] produces the auxiliary moves biII
for all i and the natural number moves y (i)
produces the auxiliary moves biI
for even i.
134 3 Games of continuously coded length

The reader may consult Diagrams 3.17 and 3.18 to check that these conditions cover all
the moves in the games G ( a , x , L ) and H (b , x , L ). Note that gen , gen ,
, and all belong to M[g ]. ( and in fact belong to M[g] M[g ].) The

construction can thus be carried out inside M[g ].


If ever an i < is reached which satisfies conditions (a)(c) in the definition of
G (a , x , L ), we must stop the construction. The reason is that both G ( a , x , L )
and H (b , x , L ) end at the first such i, and so both and cease to supply moves
for the construction at the first such i.
But if an i as above is reached, we can simply imitate the argument in Claim 3C.7
to find an s which satisfies conditions (B1)(B3).
So suppose an i < which satisfies conditions (a)(c) is not reached. Lemma 1B.2
and the use of gen during the construction guarantee that $ a , x , y %  A [g ].
In other words $x , y %  A[g ]. Similarly $b , x , y %  B [g ], so $x , y % 
B[g ]. The construction takes place inside M[g ], so certainly y M[g ]. We
have obtained clause (A) of the claim. #

Lemma 3E.12. Suppose that both the hypothesis of Lemmas 3E.8 and the hypothesis of
Lemma 3E.10 fail. Then in M[g ] there is a run $y | % so that $y | < , y %
belongs to neither A[g ] nor B[g ].

Proof. Suppose for contradiction that a run $y | % M[g ] which satisfies the
conclusion of the lemma does not exist. Note that in this case clause (A) of Claim 3E.11
can never actually occur. Claim 3E.11 therefore becomes an exact replica of Claim 3C.7.
We can now follow the argument of Section 3C(2), using Claim 3E.11 as a replacement
for Claim 3C.7, to obtain a contradiction. #

Remark 3E.13. The arrangement of the proof of Lemma 3E.12 as an argument by


contradiction is just a matter of typographical convenience. In fact it is a constructive
proof. One follows the argument of Section 3C (2) with Claim 3E.11 substituting for
Claim 3C.7, until reaching a point where an application of Claim 3E.11 yields clause (A).

Lemmas 3E.8, 3E.10, and 3E.12 together show that at least one of the cases in
Theorem 3E.1 must hold. To complete the proof of Theorem 3E.1 we must check that
the hypotheses of Lemmas 3E.8 and 3E.10 (or at least something sufficiently close
to these hypotheses) can be expressed in M using as parameters only , , and the
appropriate name A or B.
The game Gini is definable over M using the parameters , , , , and A; our
description earlier gives the definition. Let I [, , , , A, G ] be a formula which
expresses the statement G is the game Gini defined using , , , , and A.
Let I [ , , A] stand for the statement there exist some , , and G so that:
< is a Woodin cardinal; < is strong to + 1; I [, , , , A, G ] holds;
and I wins G .
Define II and II similarly. It is clear from Lemmas 3E.8, 3E.10, and 3E.12 that
I and II witness the conclusion of Theorem 3E.1.
Chapter 4
Pullbacks

The determinacy proofs in Chapter 3 may be viewed as combining the basic methods
of Chapter 1 with linear iterations. This is most clearly evident in Diagram 3.8; the
diagram folds the methods of Section 1F into a linear iteration by the extenders F .
Here we begin the process of upgrading the combination. Our intention is to fold the
methods of Chapter 1 into non-linear iterations. We have in mind very specific designs
for the non-linearity. We intend to fold the methods of Chapter 1 into iterations dictated
by the extender algebra of Woodin [45]. Our hope is to reach a point where runs of
long games are made generic, not over collapse algebras as in the previous chapters,
but over Woodin algebras.
The method we begin to engineer here, a combination of the techniques of Chapter 1
with non-linear iterations governed by Woodins extender algebra, will eventually allow
us to handle long games which run to local cardinals. But it will be a while before we get
there. This chapter only sets the grounds. Much of this setting of the grounds involves
definitions of names; we have to arrange all sorts of auxiliary games in all sorts of
particular manners. The reader may find these arrangements, presented in Sections 4C,
4D, and 4E, a little difficult to digest in a first reading. Our advice is to skim through
them, and then return to go over them more carefully while reading Chapter 5.

4A Codes
Fix a transitive ZFC model M. Let W denote the set (or class) of M which are
Woodin cardinals in M, but not limits of Woodin cardinals in M. We say that W is
a relative successor if is the first Woodin cardinal of M, or there is a largest Woodin
cardinal of M below . Otherwise W is a relative limit.
For W let e() denote sup{ + 1 | W < }. Note that e() < strictly;
otherwise would be a limit of Woodin cardinals in M, but limits of Woodin cardinals
were excluded from W .
Definition 4A.1. A position is a function F which satisfies:
(1) The domain of F is an initial segment of W . Precisely, dom(F ) = W for
some ordinal .
(2) For dom(F ) a relative successor, F () is a real.
(3) For dom(F ) a relative limit, F () is an element of (M& e() + 1) .
The least witnessing condition (1) is the relative domain of F , denoted rdm(F ).
136 4 Pullbacks

Formally we think of F () as a sequence $F ()(n) | n < %. F ()(n) is a natural


number if is a relative successor, and an element of M& e() + 1 if is a relative
limit. In the case of a relative limit it is convenient to split F () into two sequences,
$F ()(2n) | n < %, and $F ()(2n + 1) | n < %. Each of these sequences is an
element of (M& e() + 1) . We refer to them as FI () and FII () respectively.
Remark 4A.2. Let F be a position of relative domain . Suppose there are Woodin
cardinals of M above and let be the first one. Let y be a real if is a relative
successor, and an element of (M& + 1) if is a relative limit.
F and y together determine a position of relative domain + 1, namely the position
F which extends F with the value F () = y. We use F , y to denote this extended
position.
We would like to code positions using sequences of length . To do this we need
an enumeration in order type of the domain of the position. Let us start by describing
a specific method of witnessing that the domain of a position is countable in V.
Let L denote the collection of ordinals so that W is cofinal in .
Remark 4A.3. When we say that a set B is cofinal in a successor ordinal , we mean
that 1 belongs to B. We further adopt the convention that is cofinal in the ordinal 0.
Under our definition L contains 0, all ordinals in { + 1 | W }, and all limit points
of W .
Note that L contains precisely those ordinals which may serve as relative domains
(see Definition 4A.1).
Definition 4A.4. A witness for an ordinal L is a set w ( W ) which
satisfies the following conditions:
(1) for every n there exists at most one W so that $n, % w;
(2) for every W there exists at most one n so that $n, % w;
(3) the set { | n so that $n, % w} is cofinal in .
A natural number n is used in w if there exists some so that $n, % w. We write
dom(w) to denote the set of numbers which are used in w. For n dom(w) we write
w(n) for the unique so that $n, % w. Uniqueness is given by condition (1). We
write range(w) to denote the set {w(n) | n dom(w)}. This is a subset of W . By
condition (3) it is cofinal in . range(w) is countable, so the existence of a witness for
witnesses that has countable (or finite) cofinality in V.
We say that a witness w is cofinal in an ordinal just in case that range(w)
is cofinal in . (In particular must belong to L.) We allow for the case that is a
successor ordinal or 0; see Remark 4A.3. We use L(w) to denote the set of so that w
is cofinal in .
For a witness w and an ordinal we write w& to denote w ( ). If belongs
to L(w) then w& is a witness for . The converse is also true; if w& is a witness for
then it is cofinal in by condition (3) in Definition 4A.4, and from this it follows that
belongs to L(w).
4A Codes 137

Definition 4A.5. Let w and w be witnesses, for and say. w is said to extend w
just in case that w & = w.

There is another way to look at witnesses which may be intuitively clearer. Let w
be a witness for an ordinal L. range(w) is a countable set of ordinals. Let w be
the order type of this set (with the natural order on ordinals). For < w let w be
the -th ordinal in range(w), so that  w is an order preserving bijection from w
onto range(w). For < w let nw be the unique n so that w(n) = . Uniqueness
w

is given by condition (2) in Definition 4A.4. The sequences w = $w | < w % and


nw = $nw | < w % satisfy the following conditions:

(1) $w | < w % is a strictly increasing sequence of elements of W ;

(2) $w | < w % is cofinal in ; and

| < w % is a sequence of distinct natural numbers.


(3) $nw

Condition (3), which corresponds to condition (1) in Definition 4A.4, should remind
the reader of the coding used in Chapter 3.
We think of w and nw as an alternative presentation for the witness w. It is useful
to keep this alternative presentation in mind. For example Definition 4A.5 is intuitively
clearer when phrased for the alternative presentation. It is easy to check that w extends
w iff the sequences w and nw extend the sequences w and nw .

Definition 4A.6. Let L. An annotated position of relative domain is a function t


which satisfies:

(1) The domain of t is (W L).

(2) t( W ) is a position (in the sense of Definition 4A.1). Precisely: if W


is a relative successor then t () is real, and if W is a relative limit then
t () is an element of (M& e() + 1) .

(3) For each L, t () is a witness for .

There are no conflicts between the last two clauses since W and L are disjoint.
We write rdm(t) to denote the relative domain of t, namely . An annotated posi-
tion t is said to extend t just in case that t  rdm(t) is equal to t.

An annotated position thus divides into two parts: a position F = t( W ),


and a sequence of witnesses $t () | L%. Letting w = t () we often abuse
 = $w | < %, instead of
notation and refer to the second part as a sequence w
 = $w | L%.
w

Remark 4A.7. Let t be an annotated position of relative domain . Suppose there are
Woodin cardinals of M above and let be the first one. Let w be a witness for .
Let y be a real if is a relative successor, and an element of (M& + 1) if is a relative
limit.
138 4 Pullbacks

t, w, and y together determine an annotated position of relative domain + 1,


namely the annotated position t which extends t with the assignments t () = w and
t () = y. We use t, w, y to denote this extended annotated position.
Let t be an annotated position of relative domain . Let F denote the position
t( W ), and for each L let w denote t (), a witness for . Let w be a
witness for . The witnesses $w | < % and w together witness that the domain
of F is countable. They can thus be used to code F by a sequence of length . We
describe this coding more precisely below. We define a code t, w for the pair $t, w%.
We shall make sure that:
Property 4A.8. t, w is an element of (M& rdm(t)) .
Property 4A.9. t and w can be recovered from the code t, w .
The definition of t, w proceeds by induction on rdm(t). We assume knowledge
of t, w for t of relative domain smaller than rdm(t). We also assume Properties 4A.8
and 4A.9 for t of relative domain smaller than rdm(t). Given these assumptions we
make the following definitions:

Let q(w) be the sequence $


q (w)n | n < % given by:

w(n) if n dom(w);
q(w)n =
otherwise.

For n dom(w) let n = e(w(n)) and let xn = tn , wn  . (w(n) is an


element of W smaller than . e(w(n)) is therefore an element of L smaller than
= rdm(t). The definition of xn is thus meaningful by induction.)
Let x(t, w) be the sequence $x (t, w)n | n < % given by:

xn if n dom(w);
x(t, w)n =
otherwise.

Let f(t, w) be the sequence $f(t, w)n | n < % given by:



F (w(n)) if n dom(w);
f(t, w)n =
otherwise.

We think of as some low rank element of M M standing for undefined.


q (w), x(t, w), f(t, w)%. This uses
It is clear that t and w can be recovered from $
the induction hypothesis which implies that tn and wn can be recovered from xn
(Property 4A.9), the fact that range(w) is cofinal in , and the fact that w can be
recovered from q(w).
4A Codes 139

The induction hypothesis (specifically Property 4A.8) implies that xn is an element


of (M&n ) and so certainly an element of (M&) . It follows that x(t, w) is an element
of (M&) . q(w) is clearly an element of (M&) . By Definition 4A.1, F (w(n)) is
at worst an element of (M&n + 1) . n = e(w(n)) is smaller than w(n) which in turn
is smaller than . So F (w(n)) is certainly an element of (M&) . It follows from this
that f(t, w) is an element of (M&) . The entire triple $ q (w), x(t, w), f(t, w)% is
thus an element of (M&)(1++) . Using some recursive bijection b : 2
we may convert $ q (w), x(t, w), f(t, w)% to an element of (M&) . Define t, w to
be this element. It is clear that Properties 4A.8 and 4A.9 hold for $t, w%.
The precise bijection b used to convert $ q (w), x(t, w), f(t, w)% into the code t, w
is not so important, so long as it yields the following property:
Property 4A.10. For any natural number l, t, wl depends only on q(w)l,
x(t, w)0 , . . . , x(t, w)l1 , and f(t, w)0 , . . . , f(t, w)l1 .
Any reasonable choice of the bijection b will give rise to a coding which satisfies
Property 4A.10.
Definition 4A.11. A -code is a code x = t, w where t is an annotated position of
relative domain (and w is a witness for ).

Remark 4A.12. Suppose x = t, w is a -code. Suppose there are Woodin cardinals
of M above and let be the first one. Let y be a real if is a relative successor, and
an element of (M& + 1) if is a relative limit.
t, w, y is then an annotated position of relative domain + 1 (see Remarks 4A.7
and 4A.2). We use x  y to denote this annotated position. Note that t and w can be
recovered from x, so the notation x  y makes sense.
So far we saw how a position can be coded as a sequence of length , granted enough
witnesses. Let us next see how to generate witnesses. Let t be an annotated position
of relative domain . Recall that t splits into two parts, a position F = t( W )
and a sequence of witnesses $w = t () | L%. We show how to generate an
additional witness, for . We consider two cases: successor; and limit.
Successor case. Suppose that is a successor, say = +1. Suppose we are handed
and n so that:
(1) is an element of L, smaller than ; and
(2) n is a natural number which is not used in w = t ().
Let w = w {$n, %}. w is then a witness for = + 1. # (Successor case)
Definition 4A.13. We use w + $n, % to denote w {$n, %}.
Definition 4A.14 (for an annotated position t of successor relative domain = + 1).
A witness w for is amenable to t if it is equal to t () + $n, % for some and n which
satisfy the conditions of the successor case above.
140 4 Pullbacks

Remark 4A.15. Among the conditions placed on and n in the successor case let us
specifically underline the requirement that n is not used in w . This requirement is
crucial. Without this requirement w + $n, % would not be a witness; it would fail to
satisfy condition (1) in Definition 4A.4.

Claim 4A.16 (under the conditions of the successor case). The codes t, w  and
t, w + $n, % agree to n.

Proof. Let w denote w + $n, %. It is easy to see that q(w) and q(w ) agree to n. (In
fact q(w)i = q(w )i for all i except i = n.) It is easy to see that x(t, w) and x(t, w )
agree to n. (In fact x(t, w)i = x(t, w )i for all i except i = n.) It is easy to see that
f(t, w) and f(t, w ) agree to n. (In fact f(t, w)i = f(t, w )i for all i except
i = n.) Using Property 4A.10 it follows that t, w and t, w  agree to n. #

Definition 4A.17. Let w and w be witnesses. Let n be a natural number. w is said


to n-extend w just in case that:

(1) w extends w; and

(2) dom(w ) n = dom(w) n.

Note that the two conditions in Definition 4A.17 together imply that w (i) = w(i)
for all i < n in the relevant domains. Using this and an argument similar to the proof
of Claim 4A.16 one can show that:

Claim 4A.18. Let t and t be annotated positions, of relative domains and say.
Suppose that t extends t. Let w and w be witnesses for and respectively. Let n
be a natural number. Suppose that w is an n-extension of w. Then t , w  agrees
with t, w to n. #

Limit case. Let t be an annotated position of relative domain . For each L


let w denote t (). Suppose that is a limit ordinal. Suppose we are handed a set A
so that:

(1) A is contained in L;

(2) A is cofinal in ; and

(3) for all < both in A, w extends w .


 
Let w = A w = A t (). w is then a witness for . # (Limit case)

Definition 4A.19 (for an annotated position


 t of limit relative domain ). A witness
w for is amenable to t if it is equal to A t () for some A which satisfies the
conditions of the limit case above.

Claim 4A.20 (under theconditions of the limit case above). The codes t, w , for
A, converge to t, A w  as approaches .
4B Woodins extender algebra 141


Proof. Let w denote A w . Fix n < . We must find some n < so that for all
A [n , ), t, w agrees with t, w  to n.
Note that the union A w is increasing by condition (3) in the limit case. So
$dom(w ) n | A% is an increasing sequence of subsets of n. Since is a limit
ordinal and since n is finite, there must exist some point on which this sequence reaches
a terminal value. More precisely there must exist some n A so that dom(w ) n =
dom(w) n for all A [n , ). But then w is an n-extension of w for all
A [n , ). Using Claim 4A.18 it follows that t, w agrees with t, w  to n,
for all A [n , ). #
Claims 4A.16, 4A.18, and 4A.20 are similar to the agreement claims of Chapter 3,
specifically to Claims 3A.3, 3A.4, and 3A.5. The similarities become more evident
if one works with the alternative presentation described following Definition 4A.5.
The diligent reader may wish to phrase Claim 4A.20 in the language of alternative
presentations, to see more clearly the connection to Chapter 3.
We end this section with a few words on the connection of our definitions to se-
quences of reals, and on the dependence of our definitions on M.
Definition 4A.21. Let t be an annotated position of relative domain . Define z(t) to be
the sequence $t ( ) | < %, where is the order type of the set of relative successors
below and $ | < % enumerates this set in increasing order.
z(t) is a sequence of real numbers, by condition (2) in Definition 4A.6. We refer to
it as the real part of the annotated position t.
Remark 4A.22. All the notions defined in this section are dependent on M. When
we wish to emphasize this dependence we write: M-position instead of position,
annotated M-position instead of annotated position, etc.

4B Woodins extender algebra


This section is expository, presenting a forcing notion due to Woodin.
Work as usual with a transitive ZFC model M. A basic identity of type 1 is any
expression of the kind t()(n) = a where:
(1) belongs to W ;
(2) n is a natural number;
(3) if is a relative successor then a is a natural number; and
(4) if is a relative limit then a belongs to M& e() + 1.
t is just a decorative symbol which we use for aesthetic reasons. The parameters which
define a basic identity of type 1 are , n, and a.
A basic identity of type 2 is any expression of the kind a t() where:
142 4 Pullbacks

(1) belongs to L; and

(2) a belongs to ( W ).

Again t is just a decorative symbol which we use for aesthetic reasons. The parameters
which define a basic identity of type 2 are and a.
A basic identity is any expression fitting one of the two types above.
An identity is anything generated from basic identities using negations and transfinite
disjunctions.
Formally we represent identities as follows:

The basic identity t()(n) = a is represented by the triple $, n, a%.

The basic identity a t() is represented by the pair $, a%.

The identity is represented by the pair $1, r% where r is the formal represen-
tation of .

The identity < is represented by the sequence $r | < % where for each
< , r is the formal representation of .

One can check easily that the map which assigns to identities their formal representations
is injective.
When we say that the identity belongs to the model Q we mean to say that the
formal representation of belongs to Q.
The height of an identity , denoted ht( ), is defined by induction as follows:

The height of the basic identity t()(n) = a is + 1.

The height of the basic identity a t() is + 1.

The height of is equal to the height of .



The height of < is equal to sup{ht( ) | < }.

If is an identity of height , then the only expressions of the kinds t() and

t () which appear in are ones where and are smaller than . Given an annotated
position t of relative domain at least we can then evaluate the truth value of
under t.
Let us be more precise. For an identity and an annotated position t with
rdm(t) ht( ) we define the truth value of under t. The definition is by induc-
tion on the complexity of , following the natural rules:

(T1) Suppose is a basic identity of type 1, say the expression t()(n) = a. Then
is true under t just in case that t ()(n) = a.

(T2) Suppose is a basic identity of type 2, say the expression a t(). Then is
true under t just in case that a t ().
4B Woodins extender algebra 143

(T3) Suppose = . Then is true under t just in case that is not true under t.

(T4) Suppose = < . Then is true under t just in case that there exists some
< so that is true under t.

For condition (T1) recall that t () is a sequence of length , see Definitions 4A.1 and
4A.6. For condition (T2) recall that t () is a subset of ( W ), see Definitions 4A.4
and 4A.6. In both conditions we are using the assumption that rdm(t) ht( ); to
make sense of t () in the case of condition (T1) and to make sense of t () in the case
of condition (T2).
We say that is false under t just in case that ht( ) rdm(t) and is not true
under t. We write t |= to indicate that is true under t, and t  |= to indicate that
is false under t. We only use this notation when rdm(t) ht( ). Trying to evaluate
the truth value of under t when rdm(t) < ht( ) creates difficulties with conditions
(T1) and (T2). So we shall not talk about the truth value at all in such cases.

Remark 4B.1. Note that t |= and t |= can never both hold true.

We write t |= A just in case that A is a set of identities, each with height at most
the relative domain of t, and each true under t.
It is convenient to use some standard abbreviations when discussing identities. We
 to abbreviate (
use  ). We use to abbreviate ( ). We use
< to abbreviate
< . We use t ()(n)  = a to abbreviate t ()(n) = a.
Finally we use a  t() to abbreviate a t().
Fix now some which is a Woodin limit of Woodin cardinals in M. We aim to define
a forcing notion for introducing an annotated M-position of relative domain . This
forcing notion is a version of Woodins extender algebra, with generators, customized
to fit our needs.
Woodins extender algebra is a refinement of the Lindenbaum algebra on identities
in M& . To start let us work toward a definition of the Lindenbaum algebra. We work
with identities which belong to M&. Note that each identity in M& certainly has
height smaller than . So its truth value can be evaluated under any annotated position
of relative domain .
By inference relation we mean some relation  relating sets of identities with
identities. We write A  , where A is a set of identities and is an identity, to
indicate that $A, % belongs to the relation .

Definition 4B.2. Let  be an inference relation. Define an inference relation +


by throwing into + all the pairs in , and all the pairs needed to satisfy conditions
(I1)(I6) below, and nothing more.

(I1) If A  and A { }  , then A + .

(I2) If A  , then A + .

(I3) If A  , then A + .
144 4 Pullbacks

(I4) If $ | < % M is a sequence of identities


 of length < , and if there
exists some < so that A  , then A +
< .

(I5) If $ | < % M is a sequence


 of identities of length < , and if A 
for all < , then A + < .

(I6) If and are identities so that A { }  and A { }  , then A + .

We refer to conditions (I1)(I6) as inference rules. Rule (I6) corresponds to proof


by contradiction.
Remark 4B.3. If  belongs to M then so does + . In fact the map  + (acting
on inference relations in M) belongs to M.
Working by induction on define a sequence of inference relations " as
follows:
A "0 iff A M is a set of identities, A M&, and A;
for each < , " +1 = (" )+ ; and

for limit , " = < " .
Note that the entire definition can be carried out inside the model M.
Claim 4B.4 (for ). Let t be an annotated position of relative domain . Suppose
that t |= A and A " . Then t |= .
Proof. The proof is by induction on . The limit case and the case of = 0 are trivial.
The successor case too is quite easy. Let us only remark that the inference rules (I2)
and (I3) correspond to condition (T3), the inference rules (I4) and (I5) correspond to
condition (T4), and the inference rule (I6) corresponds to Remark 4B.1. #
Let " denote the relation " . Note that " belongs to M. We have:
Claim 4B.5. "+ is equal to ".
Proof. This is immediate using the restriction to sequences of length smaller than in
the inference rules (I4) and (I5). #
The relation " therefore satisfies the inference rules (I1)(I6), with both  and +
replaced by ". Because " contains "0 it satisfies the following additional rule:
(I0) If A M& in M is a set of identities and belongs to A, then A " .
We say that A infers just in case that A " . We say that an identity infers (and
write " ) just in case that the singleton set { } infers . Let us prove some simple
claims concerning this inference relation. In these claims and always stand for
identities in M&, and A and B always stand for sets of identities, in M and contained
in M& .
4B Woodins extender algebra 145

Claim 4B.6. Suppose A " and B A. Then B " .

Proof. Prove the claim for each " , working by induction on . #

Claim 4B.7. Suppose " . Then " .

Proof. We have {, } " by the inference rule (I0). Since " we have
{, } " by Claim 4B.6. Now by the inference rule (I6) it follows that { } " .
#

Claim 4B.8. " and " .

Proof. We show that " . By the inference rule (I4) we know that " .
Applying Claim 4B.7 we get " . By the inference rule (I3) it follows that
" . #

Claim 4B.9. Suppose A " ( ). Then A { } " .

Proof. Using Claim 4B.6 and the assumption that A " ( ) we get A {, } "
( ). On the other hand using the inference rules (I2) and (I5) we get {, } " ,
so certainly A {, } " . Applying the inference rule (I6) we get A { } " .
#

Claim 4B.10. Suppose A " . Then A { } " .

Proof. Recall that abbreviates ( ). Suppose A " . Using


Claim 4B.9 we see that A { } " . By the inference rule (I3), A { } " . #

Let us now proceed with the definition of Woodins extender algebra. The starting
point is the Lindenbaum algebra on identities in M& .

Definition 4B.11. For two identities , M&, put  iff " .

Claim 4B.12. The relation  is reflexive and transitive.

Proof. Reflexivity follows from the inference rule (I0). Transitivity is proved using the
inference rule (I1). #

Definition 4B.13. For identities , M&, put iff  and  .

The relation is an equivalence relation. Both the relations  and belong to M,


because " belongs to M.
For each identity M& let [ ] denote the equivalence class of . More precisely
[ ] = { M& | } (M&) where is least so that this intersection is not
empty. (We are employing here the usual Scott trick to convert from an equivalence
class which is potentially an unbounded subset of M&, to an element of M& which
represents this equivalence class in a canonical manner.) Let A M be the set of
equivalence classes. For a, b A put a A b iff  for some (any) a and some
(any) b. Which and we pick does not matter, because of the transitivity of .
146 4 Pullbacks

There is a dependence on in our definitions, which begins already in Defini-


tion 4B.2. When we wish to take note of this dependence we write A instead of A,
and write [ ] instead of [ ].
$A, A % is a partial ordering in the model M. It is a Lindenbaum algebra, computed
in M on identities in M&. Woodins extender algebra is the result of restricting this
partial order to some subset W of A. W is the set of equivalence classes which are
consistent with certain axioms. Let us describe these axioms.

Definition 4B.14. A basic axiom is any identity which fits one of the descriptions
below:

(B1) Suppose
 W is a relative successor and n is a natural number. Then
i<

t ()(n) = i is a basic axiom.

(B2) Suppose W is a relative successor, n is a natural number, and i, j are two


distinct elements of . Then t()(n) = i t()(n)  = j is a basic axiom.

(B3) Suppose W is a relative limit, n is a natural number, is an ordinal


, and $a | < % is an enumeration of M& e() + 1 which belongs
smaller than 
to M. Then < t()(n) = a is a basic axiom.

(B4) Suppose W is a relative limit, n is a natural number, and a, b are two


distinct elements of M& e() + 1. Then t()(n) = a t()(n)  = b is a basic
axiom.

(B5) Suppose belongs to L, n is a natural number, and 1 , 2 are two distinct


elements of W . Let a1 = $n, 1 % and a2 = $n, 2 %. Then a1  t()a2  t()
is a basic axiom.

(B6) Suppose belongs to L, is an element of W , and n1 , n2 are two distinct


natural numbers. Let a1 = $n1 , % and a2 = $n2 , %. Then a1  t() a2  t()
is a basic axiom.

(B7) Suppose belongs to L, and is an element of W . Suppose is an


ordinal smaller than and $a
| < % is an enumeration in M of {$n, %
( W ) | }. Then < a t() is a basic axiom.

Note how axioms (B5)(B7) correspond precisely to the conditions in Defini-


tion 4A.4. Axioms (B1)(B4) correspond to the conditions in Definition 4A.1.

Claim 4B.15. Suppose t is an annotated position of relative domain . Then all the
basic axioms are true under t. #

To generate Woodins algebra W we need some additional axioms, more restrictive


than the basic axioms, and slightly more complicated.
Suppose that and are cardinals of M, both smaller than , with W and
< . Suppose that is not a Woodin cardinal in M.
4B Woodins extender algebra 147

Suppose that E is an extender in M, with crit(E) = and strong enough that


M& + 1 Ult(M, E). Suppose further that E belongs to M& + . Let : M
Ult(M, E) be the ultrapower embedding by E.
Suppose that  = $ ( ) | < % M is a sequence of identities, of length .
Suppose that for each ,  ( ) belongs to M&. Applying to  produces a sequence
of identities ( ), of length () > . Let be the -th identity in this sequence.
Precisely, let = ( )().
Suppose that the height of is less than or equal to + 1.
Under these circumstances, and only under these circumstances, we define an iden-
tity (, , E,  ). We define:

(, , E,  ) =  ( ).
<

Definition 4B.16. An extender axiom is any identity (, , E,  ) of the kind described


above.

Remark 4B.17. The following list summarizes precisely the conditions under which
(, , E,  ) is defined:

(1) W , < , and is not a Woodin cardinal in M;

(2) E M& + is an extender of M, crit(E) = , and M& + 1 Ult(M, E);

(3)  = $
( ) | < % M is a sequence of identities, each one in M&;

(4) the identity (


)(), where is the ultrapower embedding by E, has height at
most + 1.

Let us try to explain the motivation behind extender axioms. Work with , , E, and
 as above. Let be the ultrapower embedding by E and let = ( )(). Suppose
for a moment that we are given an annotated position t (with  relative domain above all
relevant heights). If t |= ( )() then certainly t |= <() ( )( ). Since is
an
 elementary embedding we feel that morally we should be able totake the fact that
<() ( )( ) holds true, pull it down by , and conclude
 that  ( ) holds
<
true. Thus we feelthat morally t |= should imply t |= <  ( ). In other words
we feel that <  ( ) should be true. Given these feelings it seems reasonable
to take this last identity as an axiom.
The moral argument of the previous paragraph seems to indicate that any annotated
position should satisfy the extender axioms. Of course this is false; the flawed reasoning
above fails to take account of the possible effect of on t. But something close enough is
true. By passing to ultrapowers of M we can somehow eliminate those extender axioms
which fail under t. (A precise formulation of this process is given by Lemma 4B.32.)
There is therefore no harm in restricting our attention to annotated positions which
satisfy the extender axioms. In other words there is no harm in restricting our attention
to identities which are consistent with all the extender axioms.
148 4 Pullbacks

Let us carry out this restriction. Let Ax be the set of all the axioms described in
Definitions 4B.14 and 4B.16. Each identity in Ax is either a basic axiom or an extender
axiom, and all identities of this form are in Ax. Note that Ax belongs to M. This is
because both Definition 4B.14 and Definition 4B.16 can be phrased inside M.
An identity M& is bad if Ax " . An identity is good if M& and
is not bad. Since both Ax and the relation " belong to M, these notions can be (and
are) defined inside M.
Note that if then is bad iff is bad. (This uses Claim 4B.7.) Working in
M define W A by setting:

[ ] W is a good identity.

The definition makes sense because implies that is good iff is good. Let W
be the restriction of A to W. Woodins extender algebra is the partial order $W, W %.

Remark 4B.18. As noted this is a particular version of the extender algebra, specifically
customized to our own needs. We comment more on the relationship with the standard
version of Woodins extender algebra in the historical remark at the end of the section.

We use W alternately to denote the set W defined above, or the partial order
$W, W %, or the set of good identities. Which one we mean should be clear from
the context. We think of W as a forcing notion, and of elements of W as conditions.
Note that W is defined in M relative to the parameter . When we wish to emphasize
which parameter is used we write W instead of W.
Let us develop some properties of the forcing notion W. We follow the ideas of
Woodin [45], only adapted and customized to our context. We work throughout under
the assumption that is a Woodin limit of Woodin cardinals in M. This assumption
remains valid to the end of the section, even when not explicitly mentioned.

Lemma 4B.19 (assuming is a Woodin limit of Woodin cardinals). W has the chain
condition in M.

Proof. Work in M. By induction we may assume that the lemma holds for < .
Precisely:

(1) For every < , if is a Woodin limit of Woodin cardinals in M then W has
the chain condition in M.

Suppose for contradiction that T W = W is an anti-chain of length . We


view T as a sequence $ | < % where each is a good identity. The fact that T is
an anti-chain tells us that for any < < , and are incompatible in W. This
means that there is no good identity with both  and  . Using Claim 4B.8
it follows that is not good. Thus:

(2) (For any < < .) Ax " ( ).


4B Woodins extender algebra 149

Let H be the set H = {$, % | < }. We view H alternately as a sequence and


as a function, writing H ( ) for . Each identity is an element of M&. So H is a
subset of M& . Using the assumption that is a Woodin cardinal fix a cardinal <
which (in M) is <-strong wrt H . (See Definition 1A.11.)
Let < be large enough that for each , $, % belongs to M&. Using the
assumption that is a limit of Woodin cardinals in M fix < so that W and
. (Hence in particular > .)
Using the fact that is <-strong wrt H fix an extender E in M so that:

(i) crit(E) = ;

(ii) M& + 1 Ult(M, E); and

(iii) (H ) M& = H M& where is the ultrapower embedding by E.

By passing to a restriction of E if needed we may further assume that:

(iv) E belongs to M& + .

It is implicit in condition (ii) that () > (see Fact 2 in Appendix A). Thus:

(v) < < ().

Claim 4B.20. For each , (H )( ) = .

Proof. Note to begin with that (H ) is a function (in other word for each there is
at most one object z so that $, z% (H )). This follows from the fact that H is a
function using the elementarity of . Fix now . From condition (iii) and the fact
that $, % belongs to H M& if follows that $, % is an element of (H ). But this
means that (H )( ) = . #

Claim 4B.21. For each < , belongs to M&.

Proof. Fix < . By the elementarity of , (H )(( )) = ( ). Since is smaller


than it is not moved by . So (H )( ) = ( ). We already know, by the previous
claim, that (H )( ) = . So ( ) = . In particular ( ) belongs to M&.
M& = Ult(M, E)& by condition (ii), and < () by condition (v). So ( )
belongs to Ult(M, E)&(). Pulling this statement back via we get M&. #

Claim 4B.22. is not a Woodin cardinal in M.

Proof. Condition (v) and the fact that is a Woodin cardinal in Ult(M, E) together
imply that is a limit of Woodin cardinals in M. Suppose for contradiction that is
itself a Woodin cardinal in M. Thus is a Woodin limit of Woodin cardinals in M.
Using the fact that is <-strong in M one can verify that W simply equals
W (M&). But then the fact that $ | < % is a sequence of incompatible
identities in W (M&) implies that it is an anti-chain in W . This contradicts
condition (1) above with = . #
150 4 Pullbacks

Let  = H . We have ( )() = (H )() = . (The last equality uses


Claim 4B.20.) was chosen large enough that M&. Thus ( )() M&. This
certainly implies that the height of ( )() is at most + 1.
We have verified by now that the tuple $, , E,  % falls under the circumstances
required for the definition of an extender axiom (, , E, ). Since ( )() = the
extender axiom (, , E,  ) is simply the identity < . We have therefore:

(3) < belongs to Ax.
Using this and Claim 4B.10 we see that:

(4) Ax { } " < .
On the other hand using condition (2) above and Claim 4B.9 we see that Ax { } "
for all < . Using the inference rule (I5) we get:

(5) Ax { } " < .
Using conditions (4), (5), and the inference rule (I6) we conclude that Ax " .
But then is a bad identity, contradicting the fact that it (or rather its equivalence
class) is a condition in W. # (Lemma 4B.19)

Definition 4B.23. Given an annotated M-position t V with relative domain let


G(t) be defined by:

G(t) = {[ ] | is an identity in M& and t |= }.

Lemma 4B.24 (assuming is a Woodin limit of Woodin cardinals). Suppose t V is


an annotated M-position with relative domain , and suppose t |= Ax. Then G(t) is
W-generic/M.
Proof. From the fact that t |= Ax it follows that G(t) W. Using Claims 4B.4 and
4B.8 it is easy to see that G(t) is a filter. Let us check that G(t) is generic over M. Let
T = $[ ] | < % be a maximal anti-chain in W, with T M. We must find some
< so that [ ] belongs to G(t).
By
 Lemma 4B.19 we know that , the length of T , is strictly smaller than .
= < is therefore an identity in M&.

Claim 4B.25. Ax " .


Proof. Suppose not. Then is a good identity, and [ ] is a condition in W. Since
T is a maximal anti-chain there must be some < so that and are compatible.
So there must be some good identity so that " and " . From " we
get " using the inference rule (I4) and the definition of . Then using the inference
rule (I6) we get " . But then certainly Ax " , contradicting the fact that is
good. #
4B Woodins extender algebra 151

Using the last claim and the initial assumption that t |= Ax we get t |= . This
means that there exists some < so that t |= . We have [ ] G(t) for that .
# (Lemma 4B.24)
Definition 4B.26. Suppose is an element of W and suppose t V is an annotated
M-position of relative domain + 1. An M-obstruction for t is a pair $E,  % which
satisfies conditions (1)(6) below:
(1) E is a + 1-strong extender in M& + , with crit(E) < .
Let < denote crit(E) and let denote the ultrapower embedding by E.
(2) is not a Woodin cardinal in M;
(3)  M is a sequence of identities of length ,  = $
( ) | < % say, and each
identity  ( ) belongs to M&;
(4) for each < , t |= 
( );
)() is at most + 1; and
(5) the height of the identity (
(6) t |= (
)().
t is obstruction free over M if it has no M-obstructions.
Note that Definition 4B.26 is made with no reference to . This is important; we
will later use the definition in general contexts with no specific in mind.
Remark 4B.27. By a -iseq ( identities sequence) we mean a sequence  of the kind
described in condition (3) of Definition 4B.26. A -iseq  = $ ( ) | < % is realized
by t just in case that there is some < so that t |=  ( ).
Remark 4B.28. Suppose $E,  % is an M-obstruction for t. We use (E) to denote the
first Woodin cardinal of M above crit(E). Since itself is a Woodin cardinal of M
we have certainly (E) . Note that (E) is a relative limit in W . This is because
crit(E) is a limit of Woodin cardinals of M by condition (1) in Definition 4B.26, but
not a Woodin cardinal of M by condition (2).
Definition 4B.29. An annotated position t is M-clear if t + 1 is obstruction free over
M for all rdm(t) W .
Corollary 4B.30. Suppose t V is an annotated M-position of relative domain ,
where is a Woodin limit of Woodin cardinals. Suppose further that t is M-clear. Then
G(t) is W -generic/M.
Proof. From the assumption that t is clear it follows easily that all extender axioms are
true under t. (Note that if an extender axiom (, , E,  ) fails under t then $E,  % is an
obstruction for t + 1.) Certainly all position axioms are true under t, see Claim 4B.15.
So t |= Ax. Now apply Lemma 4B.24. #
152 4 Pullbacks

Definition 4B.31. An M-obstruction $E,  % for t is called minimal if it satisfies the


following condition in addition to the conditions of Definition 4B.26:

(7) t is obstruction free over Ult(M, E).

Lemma 4B.32. Suppose belongs to W and suppose t V is an annotated M-position


of relative domain + 1. Then either t is obstruction free over M, or else there exists
a minimal M-obstruction for t.

Lemma 4B.32 says that either t is obstruction free, or else it can be made obstruction
free by passing to a certain ultrapower of M.

Proof (of Lemma 4B.32) . If t is obstruction free over M we are done. So suppose
t is not obstruction free. Fix an obstruction $E0 , 0 % M. We continue to pick
pairs $En+1 , n+1 % working by induction on n < . If ever an n < is reached so
that t is obstruction free over Ult(M, En ) we end the construction. Otherwise we let
$En+1 , n+1 % be some Ult(M, En )-obstruction for t and continue the construction.
Note that an Ult(M, En )-obstruction for t is also an M-obstruction for t. It is thus
enough to show that our construction ends at some finite n.
Suppose for contradiction the construction continues through all n < . We have
En+1 Ult(M, En ) by construction. In other words, $En | n < % is an infinite
descending chain in the Mitchell order on short extenders in M. But this order is
wellfounded by the results of Steel [42]. #

So far we considered how an annotated position t can induce a generic G(t). Let
us now consider the opposite direction. Suppose for the rest of the section that G is
W-generic/M.

Claim 4B.33. Let be an identity in M&. Suppose [ ] belongs to G. Then [ ] does


not belong to G.

Proof. Suppose for contradiction that both [ ] and [ ] belong to G. Since G is a


filter there must exist some [] G which is below both [ ] and [ ]. So " and
" . From this and the inference rule (I6) it follows that " . So certainly
Ax " . But then []  W. This contradicts the fact that [] G W. #

Claim 4B.34. Let$ | < % be a sequence of identities in M& , of length less than
. Suppose that < is one of the basic axioms. Then there exists some < so
that [ ] G.

Proof. Let D W be defined by:

[] D iff is a good identity and  for some < .

It is enough to check that D is dense in W. So fix some condition [ ] W. We wish


to find [] W so that  and [] D. If for some < , is good, then
= works. So suppose for contradiction that for each < , is bad. In
4B Woodins extender algebra 153

other words Ax infers ( ) for each < . By Claim 4B.9 then Ax {} "
for each < .Using the inference rule (I5) it follows that Ax { } " < .
By assumption < is a basic axiom, so certainly Ax { } " < . Using the
inference rule (I6) it follows that Ax " . So is a bad identity. But this contradicts
the fact that [ ] is a condition in W. #

Claim 4B.35. Suppose n is a natural number and suppose W is a relative


successor. Then there exists some natural number i so that [t()(n) = i] belongs to G.

Proof. This is a direct application of Claim 4B.34 with the basic axiom (B1) of Defini-
tion 4B.14. #

Claim 4B.36. Suppose n is a natural number and suppose W is a relative


successor. Then there exists at most one natural number i so that [t()(n) = i] belongs
to G.

Proof. Suppose for contradiction that both [t()(n) = i] and [t()(n) = j ] belong to
G, where i and j are distinct natural numbers. By Claim 4B.33, neither [t()(n)  = i]
nor [t()(n) = j ] belongs to G. Now apply Claim 4B.34 with the basic axiom (B2) to
get a contradiction. #

Claim 4B.37. Suppose n is a natural number and suppose W is a relative limit.


Then there exists exactly one a M& e() + 1 so that [t()(n) = a] belongs to G.

Proof. Similar to the last two claims, but using the basic axioms (B3) and (B4). #

Claim 4B.38. Suppose is an element of L. Then:

(1) For every n there is at most one W so that [$n, % t()] belongs
to G.

(2) For every W there is at most one n so that [$n, % t()] belongs
to G.

(3) The set { | n so that [$n, % t()] belongs to G} is cofinal in .

Proof. Similar to the previous claims, using the basic axioms (B5)(B7). #

For W a relative successor define a real y = $y (n) | n < % by letting


y (n) be the unique natural number i so that [t()(n) = i] belongs to G. Existence and
uniqueness are given by Claims 4B.35 and 4B.36.
For W a relative limit define y = $y (n) | n < % in (M& e() + 1) by
letting y (n) be the unique element a of M& e() + 1 so that [t()(n) = a] belongs to
G. Existence and uniqueness are given by Claim 4B.37.
For L define w ( W ) by putting a w iff [a t()] belongs
to G. By Claim 4B.38, w is a witness for .
154 4 Pullbacks

Working inside the generic extension M[G] let t (G) be the annotated position, of
relative domain , defined by:

y if belongs to W ; and
t (G)() =
w if belongs to L.

Definition 4B.39. Let t M be the canonical name, in the forcing W, for the annotated
position t (G) defined above.

Remark 4B.40. We are using the symbol t for two purposes. On the one hand it is
used as a decorative symbol in the notation for identities. On the other hand it is used
as the canonical name for t (G). The two uses are related. For example: the condition
[t()(n) = a] forces the statement t()(n) = a, the condition [a t()] forces the
statement a t(), etc.

To each M-clear annotated position t of relative domain we associated earlier a


generic filter G(t). To each generic filter G we now have associated an annotated posi-
tion t[G] = t (G) of relative domain . The following claim relates these associations
to one another.

Claim 4B.41. Suppose t V is an M-clear annotated position of relative domain .


Then t[G(t)] = t. In particular t belongs to M[G(t)]. #

To end let us record the following claim:

Claim 4B.42. Suppose that Y M is a W-name for a set of annotated positions of


relative domain . Then there is some ordinal , strictly smaller than , so that:

() For any G which is W-generic/M, t[G] Y [G] iff this is forced by some condition
in G (M&).

Proof. Let K W be the set of conditions which force t Y . Let T K be


a maximal anti-chain in K. By Lemma 4B.19, T has size strictly smaller than .
Since each element of T belongs to M&, and since is a Woodin cardinal and hence
inaccessible in M, it follows that T is contained in M& for some strictly smaller
than . It is easy to see that this witnesses (). #

Definition 4B.43. An ordinal < which satisfies () of Claim 4B.42 is said to


seal Y .

Historical Remark. It was Woodin [45] who introduced extender axioms; used them to
refine the Lindenbaum algebra on identities in M& to the point of having the chain
condition; and noted that any extra hardship imposed by the extender axioms can be
removed through the use of iterated ultrapowers. The end result was a forcing notion
in M which could, through the formation of an iteration tree on M, be made to admit
any real as generic.
4C Names, part I 155

All the definitions and claims presented in Section 4B trace back directly to the work
of Woodin [45]. W is essentially Woodins algebra for identities on generators, using
only extenders which overlap Woodin cardinals and have non-Woodin critical points.
We introduced some cosmetic changes to the definitions. For example our notion
of an identity is specifically tailored to create a forcing notion which adds an annotated
position, rather than an arbitrary sequence. Our basic axioms are also specifically
tailored for this purpose.
We made one change which is a little more substantial. A direct adaptation of Woodin
[45] would define an extender axiom (, , E,  ) under the conditions summarized
in Remark 4B.17, except that condition (4) would have an additional clause:
(4)(b) (
)() has hereditary cardinality at most (so that it is not affected by ele-
mentary embeddings with critical points above ).
Our ability to manage without this clause ultimately traces to the fact that is a Woodin
cardinal.

4C Names, part I
Fix a transitive ZFC model M. We work in M and in generic extensions of M through-
out this section. When we say a Woodin cardinal we mean a Woodin cardinal of M.
When we say generic we mean generic over M. By annotated position we always
mean an annotated M-position. We occasionally write position instead of annotated
position. But we always mean annotated position.
Definition 4C.1. Suppose belongs to W . A -sequence is an M-clear annotated
position of relative domain + 1. A -name is a col(, )-name C for a set of -
sequences.
We intend to define a certain pullback operation on names. Given < both in W
and a -name C , we intend to define a -name which we refer to as the (, )-pullback
of C , denoted Back(, )(C ) (read back , of C ). The complete definition of
the pullback operation stretches over Sections 4C and 4D. Roughly speaking we intend
Back(, )(C ) to name the set of -sequences from which player I can win to produce
a -sequence which belongs to (an interpretation of a shift of) C . But this intuition
is not entirely accurate; we fold into the definition of Back(, )(C ) some extra
ingredient, intended to remove obstructions to the positions encountered along the way
to C .

4C (1) Removing obstructions. We begin by making precise our plans for dealing
with obstructions. Suppose W , and suppose t is an annotated position of relative
domain + 1, which exists in some generic extension of M. We work to define a
property of t which tells us that obstructions for t can be removed while making t
winning for I. We will later fold this property into the definition of the pullback.
156 4 Pullbacks

Definition 4C.2. Let Q M be a forcing notion. Let u M be a name in Q. For


a set a we write a u[] to mean that a belongs to u[G] for some G which is Q-
generic/M. We write a  u[] to mean that a does not belong to u[G] for any G which
is Q-generic/M (possibly for the trivial reason that a does not belong to M[G]).

Remark 4C.3. Any extension N of M which has a as an element can figure out whether
a u[] or not. To see this first use absoluteness to note that a u[] iff a G witnessing
this exists in the generic extension of N by col(, 2|Q| + |trcl(u, a)|). Then use the
homogeneity of the collapse. (trcl(u, a) here is the transitive closure of {u, a}.)

Definition 4C.4. Let be an element of L. A -functor is a function c which satisfies


the following conditions:

(1) c is a function from I = {$, % ( L) ( W ) | < } into M&; and

(2) for each pair $, % in I , c(, ) is a col(, )-name.

Recall that we are working with some W and some annotated position t of
relative domain + 1. Let be a relative limit in W . By Definition 4A.6, t ( ) =
$t ( )(n) | n < % is an element of (M& e( ) + 1) . We split t ( ) into two sequences,
denoted tI ( ) and tII ( ), by setting tI ( )(n) = t ( )(2n) and tII ( )(n) = t ( )(2n + 1).
We only use tI in this section, leaving tII for the mirrored definitions of Section 4E.
tI ( ) is an element of (M& e( ) + 1) .

Definition 4C.5 (for t an annotated position of relative domain + 1, and a


relative limit in W ). t is suitable at just in case that for every n < , tI ( )(n) is an
e( )-functor.

An e( )-functor is formally a set of pairs taken from (( L) ( W )) (M&)


where = e( ). In particular it is a subset of M& = M& e( ). Definition 4C.5 is
thus compatible with the fact that tI ( )(n) is an element of M& e( ) + 1. It places some
structural restrictions on the format of tI ( )(n), saying that it is not just any element of
M& e( ) + 1, but that in fact it is a function of a particular kind.

Definition 4C.6 (for t an annotated position of relative domain + 1). $E,  % is an


acceptable obstruction for t just in case that:

(1) $E,  % is a minimal M-obstruction for t, see Definition 4B.31;

(2) t is suitable at (E); and

(3) there exists a natural number n so that t belongs to (c)(, )[] where stands
for the ultrapower embedding by E, c stands for tI ( (E))(n), and stands for
crit(E).

Some words are needed to explain Definition 4C.6. Let us start with the structural
aspects. Recall that (E) is the first Woodin cardinal of M above crit(E). (E) W is
a relative limit by Remark 4B.28. The reference to suitable in condition (2) therefore
4C Names, part I 157

makes sense. Having demanded that t be suitable at (E) we know that c = tI ( (E))(n)
is a crit(E)-functor. In particular the domain of c is equal to the set I of Definition 4C.4
with = crit(E). Now , the ultrapower embedding by a + 1-strong extender, sends
above which remains a Woodin cardinal in the ultrapower. From this it follows
that $, % belongs to ( L) ( W ), and hence to the domain of (c). The
restrictions placed on c by Definition 4C.4 imply that (c)(, ) is a name in col(, ).
The reference to (c)(, )[] in condition (3) of Definition 4C.6 therefore makes sense.
Definition 4C.6 says that t belongs to a shifted interpretation of a name recorded as
part of t. The record referred to is tI ( (E)). This record is shifted by E. A particular
name (c)(, ) in the shifted record is taken. Definition 4C.6 says that t belongs to an
interpretation of this name.
As it stands Definition 4C.6 says very little; we know nothing of what kind of
names are recorded in tI ( (E)). To make the definition useful when reaching t of
relative domain + 1 we should make sure to have recorded useful information in tI ( ),
at relative limits earlier on in the construction of t.
In Section 4C(3) we shall say precisely what kind of names our constructions
record in tI ( ). Roughly speaking we intend to use tI ( ) to record names for winning
positions for I (whatever that means). Passing to Ult(M, E) then makes t into such
a winning position; this is the point of condition (3) in Definition 4C.6. Now passing
to Ult(M, E) also removes all obstructions for t; this is the point of the minimality
demand in condition (1). So passing to Ult(M, E) makes t both obstruction free and
winning for I. (This is assuming that E is part of an acceptable obstruction for t, and t
has useful information recorded at relative limits.)
Definition 4C.7 (for t an annotated position of relative domain + 1). t is acceptably
obstructed if there exists an acceptable obstruction for t.
From now on we intuitively think of acceptably obstructed positions t as winning
for I. Our justification for this intuition is the paragraph preceding Definition 4C.7.
Remark 4C.8. Conditions (1) and (2) in Definition 4C.6 can clearly be phrased over
any extension of M which has t as an element. Condition (3) can also be phrased over
any extension of M which has t as an element, using Remark 4C.3. So any extension
of M which has t as an element can figure out whether t is acceptably obstructed.

4C (2) Relative successors. Let be a relative successor in W . Fix a -name C .


In other words fix a col(, )-name C M for a set of M-clear annotated positions
of relative domain + 1.
Let = e( ). Fix (in some generic extension of M) an annotated position t of
relative domain , and a witness w for .
For expository simplicity fix g which is col(, )-generic/M. Working in M[g ]
define A (M&) by:
A = {$x, y% M[g ] | x is a -code, y is a real, and x  y is either:
(a) an element of C [g ]; or (b) acceptably obstructed}.
158 4 Pullbacks

(See Remark 4A.12 for the definition of x  y.) A is a subset of (M&) . Note
that clause (b) in the definition of A can be phrased over M[g ] using Remark 4C.8.
Let A M be the canonical name for (M& ) A . We view A as naming a
subset of (M& ) ((M&) ) . Let A = (x, y  A [x, y]) be the auxiliary
games map associated to A , , and X = (M&) .
Remark 4C.9. For future reference let us record the fact that we are using the definitions
of Chapter 1 with the set X = (M&) , where = e( ).
Define the game Gsuc (t, w, C ) to be played as follows: I and II play natural
numbers y(i), creating together the real y = $y(i) | i < %. In addition they play
moves in the auxiliary game A [t, w, y]. Infinite runs of Gsuc (t, w, C ) are won by
player II.

I y(0) a0I a1I y(2) ...


II a0II y(1) a1II ...

Diagram 4.1. The game Gsuc (t, w, C ).

As usual we use the fact that A is Lipschitz continuous. Knowledge of (t, w,


and) yi suffices to determine the rules for round i of A [t, w, y]. The definition of
Gsuc (t, w, C ) therefore makes sense.
Moves in Gsuc (t, w, C ) are elements of M (in fact of M& ). But the game itself is
defined modulo knowledge of t and w. So the game itself only exists in models which
have t and w as elements.
We shall refer to the games Gsuc (. . . ) in our future definitions of the pullback
operation. Informally Gsuc (t, w, C ) can be viewed as a game where players I and II
collaborate to produce y, and in addition player I works against the auxiliary moves
played by II to try and witness that the annotated positions t, w, y is either:
(a) an element of C [h ] for some col(, )-generic h ; or
(b) acceptably obstructed.
Let suc (t, w, C ) be the statement player I has a winning strategy in

Gsuc (t, w, C ). suc (t, w, C ) can be evaluated in any model which contains M and
has t and w as elements. Since Gsuc (t, w, C ) is an open game, and since winning an
open game is absolute, we get:
Claim 4C.10. suc (t, w, C ) is absolute between any two generic extensions of M
which have t and w as elements. #

4C (3) Relative limits and records. Fix a relative limit in W . We work to define the
pullback operation on -names. More precisely we define the (, )-pullback operations
for our fixed and all relevant . We assume complete knowledge of the pullback oper-
ation below . More precisely we assume knowledge of the (, )-pullback operations
for all < in W , and all relevant .
4C Names, part I 159

Remark 4C.11. This inductive assumption encompasses all W , relative limits


and relative successors alike. But the definition of the pullback is limited here to the
case of a relative limit W . We shall define the pullback operation in the case of
relative successors later on, in Section 4D.
Fix a -name Y . Let = e( ). We make various supporting definitions before
defining the pullback operation on Y :
For each ordinal we define a col(, )-name R[ ] for a subset of (M& )
(M&) .
For each ordinal , each P (M& )< , each W , and each L,
we define a -name C(, , P , ).
These objects are defined with reference to Y . Really we should call them RY [ ] and
CY (, , P , ). But Y is fixed here, and so we suppress its mention in the notation.
The supporting definitions are made by induction on the ordinal . Let us fix and
work to define C(. . . , ) and R[ ]. We assume knowledge of C(. . . , ) and R[ ] for
all < .

4C(3)(a) The definition of C(. . . , ). Fix P (M& )< . Fix W . Fix


L. We work to define C(, , P , ).
Recall that by induction we know R[ ] for each < . R[ ] is a col(, )-name
for a subset of (M& ) (M&) . Let R[ ] be the auxiliary games map associated
to R[ ], , and X = M&.
Remark 4C.12. For future reference let us record the fact that we are using the defini-
tions of Chapter 1 with the set X = M&, where = e( ).
For expository simplicity fix some g which is col(, )-generic/M. Fix in M[g] an
M-clear annotated position t of relative domain + 1. Let n = lh(P ). Suppose that
n is not used in t (). Let w = t () + $n, % (see Definition 4A.13). Let x = t, w.
Let be the first Woodin cardinal above .
Define Glim (t, , , P , ) to be the game played according to Diagram 4.2, rules
(1)(3), and the payoff condition (P) below.

I , n n anI ... (n 1) a(n 1)I


II anII ... a(n 1)II

Diagram 4.2. The game Glim (t, , , P , ).

(1) (Rule for I)


(a) is an element of W , greater than and smaller than .
160 4 Pullbacks

(b) n is a natural number which is strictly greater than n, and not used in w.

After I plays and n subject to rule (1), the game continues with n n rounds subject
to rules (2) and (3) below. We use i to index these rounds, starting with i = n. We
set ai = $aiI , aiII % following round i. We use Pn to denote P and inductively define
Pi+1 = Pi , ai as the game proceeds.

(2) (Rule for I) i is an ordinal chosen so that:

(a) Pi is a legal position in R[i , x];


(b) i < if i = n; and
(c) i < i1 if i > n.

(3) (Rule for I and II) aiI and aiII are legal moves in the game R[i , x], following
the position Pi .

(We remind the reader that in the above rules, is the first Woodin cardinal above ,
n = lh(P ), w = t () + $n, %, and x = t, w.)
Once the game is over we let P = Pn and let = n 1 . We set = + 1
and C = C( , , P , ). We finally set C = Back( , )(C ).

(P) The run of Glim (t, , , P , ) described above is won by player I just in case that
suc (t, w, C ) holds true.

Remark 4C.13. We made various uses of the inductive assumptions in the definition of
Glim (t, , , P , ). We used knowledge of the names R[ ] for < in rules (2a) and
(3). (Note that each i is smaller than by rules (2b) and (2c).) We used knowledge of
the names C(. . . , ) for < to define the name C in preparing to phrase the payoff
condition. (Note that < .) Finally we used knowledge of the pullback operation
below to define the name C . (We used knowledge of the ( , )-pullback operation.
Note that < by rule (1a).)

Definition 4C.14. Let C(, , P , ) be the canonical name for the set of t M[g] so
that:

(1) t is an M-clear annotated position or relative domain + 1;

(2) n = lh(P ) is not used in t (); and

(3) I wins the game Glim (t, , , P , ).

Remark 4C.15. For future reference let us observe that the dependence of
C(, , P , )[g] on is monotone increasing: if  then C(, , P , )[g]
C(, , P ,  )[g]. To see this simply note that Glim (t, , , P , ) becomes easier for I
as increases.
4C Names, part I 161

4C(3)(b) The definition of R[ ]. Continue to work with a fixed ordinal . We have


at our disposal the names C(. . . , ), defined above. Our next goal is to define R[ ].
Recall that all our definitions here are made with reference to a -name Y fixed at the
outset of Section 4C(3). We shall refer to Y in condition (3a) of Definition 4C.18 below.
For expository purposes fix h which is col(, )-generic/M. We define a set R[ ]
(M& ) (M&) working inside M[h]. We shall later let R[ ] be the canonical name
for this set.
Definition 4C.16. Given two sequences u = $u(n) | n < % and v = $v(n) | n < %,
let uv be the length sequence defined by (uv)(2n) = u(n) and (uv)(2n+1) =
v(n) for each n < .
Let I = {$, % ( L) ( W ) | < }.
Definition 4C.17 (for a (M& ) ). For each n < let ua, (n) : I M& be the
function defined by ua, (n)(, ) = C(, , a n, ). Let ua, denote the sequence
$ua, (n) | n < %.
ua, is the record we attach to a and . It records, in the manner of Definition 4C.4,
the names C(. . . , P , ) for P -s which are initial segments of a .
Definition 4C.18. Working in M[h] let R[ ] be the set of all pairs $
a , x% so that:
(1) a belongs to (M& ) ;
(2) x (M&) is a -code; and
(3) there exists some v (M& + 1) so that x  (ua, v) is either:

(a) an element of Y [h], or


(b) acceptably obstructed.
(Clause (3b) can be phrased over M[h] using Remark 4C.8.) Let R[ ] be the canonical
name for R[ ].
Remark 4C.19. Let t denote the position x  (ua, v) of condition (3) in Defini-
tion 4C.18. This is an annotated position of relative domain + 1. Note how tI ( ),
which is equal to ua, , records the names C(, , a n, ) ranging over , , and n.
This is in line with our stated intention in Section 4C (1) to use tI ( ) to somehow record
winning positions for player I.

4C(3)(c) The pullback operation on Y . We defined so far the following objects:


For each ordinal , we defined a col(, )-name R[ ] for a subset of (M& )
(M&) .
For each ordinal , each P (M& )< , each W , and each L,
we defined a -name C(, , P , ).
162 4 Pullbacks

These objects were defined with reference to a fixed -name Y . The dependence of the
definitions on Y came in through condition (3a) of Definition 4C.18. Directly it may
seem that only the definition of R[ ] depends on the fixed -name Y . But remember
that the definition of C(. . . , ) is made with reference to the names R[ ] for < ,
and hence with indirect dependence on Y .
Let $L , H % be the least pair of local indiscernibles of M relative to , see Defini-
tion 1A.15 for details.

Definition 4C.20. Define Back(, )(Y ) to be the name C(, , P , ) taken with = 0,
P = , and = L .

4D Names, part II
We continue to work with a fixed transitive ZFC model M, as we did in Section 4C.
As before Woodin means Woodin in M.

Definition 4D.1. Suppose is a Woodin limit of Woodin cardinals. A -sequence is an


M-clear annotated position of relative domain . A -name is a W -name C for a set
of -sequences.

Definitions 4C.1 and 4D.1 together give the notions of -sequences and -names
for all Woodin cardinals , whether limits of Woodin cardinals or elements of W . Note
that there is no conflict between the two definitions, since Woodin limits of Woodin
cardinals are excluded from W .
In Section 4C we initiated the definition of a pullback operation on names. We
already defined the (, )-pullback operations (for W ) in the case that is a
relative limit in W . We have yet to handle the case of relative successors. But let us
already enlarge our goals. We wish to define a (, )-pullback operation not only for
W , but also in the case that is a Woodin limit of Woodin cardinals.

4D (1) Woodin limits of Woodin cardinals. Let be a Woodin limit of Woodin cardi-
nals. Fix a -name Y . We work to define the ( , )-pullback of Y for each W .
We assume knowledge of the pullback operation below during the definition.
Let < be the least ordinal which seals Y , see Definition 4B.43.

Definition 4D.2 (for [, ) W , where < is the least ordinal which seals Y ).
Let s be an M-clear annotated position of relative domain + 1. s is hopeful wrt Y if
there exists some good identity M& so that:

(1) s |= ; and

(2) [ ] W t Y .

(We are employing here the notation and terminology of Section 4B.)
4D Names, part II 163

Definition 4D.2 is restricted to . Note that condition (1) makes sense granted
this restriction: belongs to M& and therefore has height less than . The relative
domain of s is + 1, which is greater than . So the truth value of can be evaluated
under s.

Claim 4D.3. Suppose  W is greater than . Suppose s  is an M-clear annotated


position of relative domain  + 1 which extends s. Then s  is hopeful wrt Y iff s is
hopeful wrt Y .

Proof. Since s  extends s and since , we have s   = s. Using this we see that
s |= s  |= for any M&. The claim follows. #

Continue to work with the fixed -name Y , and with standing for the least ordinal
which seals Y . We make the following supporting definition before defining the pullback
operation on Y :

For each ordinal , each [, ) W , each L, and each n < we


define a -name C(, , n, ).

The name is defined with reference to Y and really we should call it CY (, , n, ). But
Y is fixed and so we suppress its mention in the notation. As usual the definition is by
induction on .
Let us begin the definition. Fix an ordinal . Fix [, )W . Fix L. Fix
n < . For expository simplicity fix some g which is col(, )-generic/M. Fix in M[g]
an M-clear annotated position s of relative domain + 1. Suppose that n is not used in
s(). Let w denote the witness s() + $n, %. We define a game Gwl (s, , , n, ) as
follows:
Players I and II play according to Diagram 4.3 and rules (1)(3) below.

I , n I , n
II II

(If s is hopeful wrt Y ) (If s is not hopeful wrt Y )

Diagram 4.3. The game Gwl (s, , , n, ).

(1) (Rule for II) < .

(2) (Rule for I) W is greater than , and greater than the first Woodin
cardinal above . n is a natural number which is not used in w.

(3) (Rule for I or II) < .

Once the game is over we let = + 1 and let C = C( , , n , ). We let be


the first Woodin cardinal above , and let C = Back( , )(C ).
164 4 Pullbacks

(P) The run of Gwl (s, , , n, ) described above is won by player I just in case that
suc (s, w, C ) holds true.
Remark 4D.4. The move corresponding to rule (3) is played by II if s is hopeful
with respect to Y , and by I otherwise.
The game Gwl (s, , , n, ) is very simple. The two players together come up with
[, ) W , n dom(w), and < . They look at C( + 1, , n , )
and essentially ask whether I can win to enter this name.
The moves and n are made by player I but player II can force to be arbitrarily
large below . The move we think of as a step towards a descending chain. Which
player is burdened with this step depends on whether or not s is hopeful with respect
to Y . The placement of the burden is in line with our intuitive view of Y as a payoff set
for I. If s is hopeful wrt Y then I is making good progress towards this payoff set, so
we punish II with the burden of the descending ordinal. If s is not hopeful wrt Y then
we feel that I is making very bad progress towards the payoff set, and so we punish I.
Definition 4D.5. Define C(, , n, ) to be the canonical name for the set of s M[g]
so that:
(1) s is an M-clear annotated position of relative domain + 1;
(2) n is not used in s(); and
(3) I wins the game Gwl (s, , , n, ).
Recall that < is the least ordinal which seals Y . Definition 4D.5 determines
a -name C(, , n, ) for each [, ) W , each L, each n < , and
each ordinal . The definition is inductive in two ways. It assumes knowledge of the
pullback operation below , and it assumes knowledge of the names C(. . . , ) for
< . Both assumptions were used in preparing to phrase the payoff condition (P);
the first inductive assumption was used in the definition of C , and the second inductive
assumption was used in the definition of C .
Let us continue to work with the fixed -name Y . We continue to use to denote
the least ordinal which seals Y . Fix now some W . may be an element
of [, ) or it may be smaller than . For expository simplicity fix some q which is
col(, )-generic/M.
Let $L , H % be the least pair of local indiscernibles of M relative to . Let be the
first Woodin cardinal above . For each > in [, ) W let C = C(, , n, )
where = 0, n = 0, and = L . Let A = Back( , )(C ). (This last assignment
requires knowledge of the ( , )-pullback operation. We have this knowledge by
induction since is smaller than .)
Definition 4D.6. Define Back(, )(Y ) to be the canonical name in col(, ) for the
set of r M[q] so that:
(1) r is an M-clear annotated position of relative domain + 1; and
4D Names, part II 165

(2) in M[q] there exists some witness w for + 1 and there exists some > in
[, ) W so that suc (r, w, A ) holds true.

4D (2) Relative successors. Let us now begin to discharge our obligation to define
the pullback operation in the case of relative successors in W . Fix a relative successor
W . Let be the largest Woodin cardinal below .

Remark 4D.7. Recall that the first Woodin cardinal is considered a relative successor.
When we say let be the largest Woodin cardinal below we make the implicit
assumption that is not the first Woodin cardinal.

Let C be a -name. We work to define Back(, )(C ). We divide the definition


into two cases, depending on whether is a Woodin limit of Woodin cardinals, or an
element of W .

Case 1. If belongs to W . For expository simplicity fix some g which is col(, )-


generic/M. Define Back(, )(C ) to be the canonical name for the set of t M[g]
so that:

(A1) t is an M-clear annotated position of relative domain + 1; and

(A2) in M[g] there exists some witness w for + 1 so that suc (t, w, C ) holds true.

We refer the reader to Section 4C(2) for the definition of suc . # (Case 1)

Case 2. If is a Woodin limit of Woodin cardinals. A -name in this case is a name in


the forcing W , see Definition 4D.1. Fix for expository simplicity some G which is W -
generic/M. Define Back(, )(C ) to be the canonical name for the set of t M[G]
so that:

(B1) t is an M-clear annotated position of relative domain ; and

(B2) the statement there exists some witness w for so that suc (t, w, C ) holds true
is forced in col(, ) over M[G].

Forced in condition (B2) means forced by some condition. But this is the same as
forced by the empty condition, since col(, ) is homogeneous. # (Case 2)

The differences between the two cases above stem from the differences between
Definitions 4C.1 and 4D.1. In condition (A1) we talk about t of relative domain + 1;
while in condition (B1) we talk about t of relative domain . In condition (A2) we work
over the model M[g] where g collapses ; while in condition (B2) we work over the
model M[G] where G is generic for W .
In the case of condition (B2) note that we cannot hope to find w inside the model
M[G]. This is because W has the chain condition, and so maintains uncountable
cofinality in M[G]. Thus instead of searching for w in M[G] we search for it in
M[G]col(,) .
166 4 Pullbacks

4D (3) Compositions. So far we defined the pullback operation in the following three
basic cases:
(I) We defined the (, )-pullback operation in the case that is a relative limit in
W and < is an element of W .
(II) We defined the (, )-pullback operation in the case that is a Woodin limit of
Woodin cardinals and < is an element of W .
(III) We defined the (, )-pullback operation in the case that is a relative successor
in W and is the largest Woodin cardinal below .
We deal with the remaining cases by composing the basic cases. The exact compositions
are defined in cases (IV) and (V) below.
(IV) From successor. Suppose is a relative successor in W and C is a -name. Let
be the largest Woodin cardinal below . Suppose < is an element of W .
Let C = Back(, )(C ). This assignment uses the definitions of Section 4D (2),
listed above as case (III). Let A = Back(, )(C). This assignment is possible by
induction, since is smaller than . Define Back( , )(C ) to be A.
(V) To Woodin limit. Suppose is a Woodin limit of Woodin cardinals. Let be the
first Woodin cardinal above . Suppose is a Woodin cardinal greater than . Suppose
C is a -name.
Let C = Back( , )(C ). Since belongs to W this assignment can be made
using one of the cases (I)(IV) listed so far. Which case is used depends on whether
is a relative limit, a relative successor, or a Woodin limit of Woodin cardinals, and
on the distance between and . Let C = Back(, )(C ). This assignment uses
the definitions of Section 4D(2), listed above as case (III). Define Back(, )(C ) to
be C.

4D (4) Summary. Our work in Sections 4C and 4D defines a general pullback opera-
tion on names.
Suppose is a Woodin cardinal, either a Woodin limit of Woodin cardinals or
an element of W . Suppose C is a -name. (The notion of a -name is given by
Definition 4D.1 when is a Woodin limit of Woodin cardinals, and by Definition 4C.1
when belongs to W .) Suppose that is a Woodin cardinal, again either a Woodin
limit of Woodin cardinals or an element of W , and smaller than .
Back(, )(C ), the (, )-pullback of C , is defined under these general assump-
tions. The definition is by induction on , using cases (I)(V) above. The reader may
easily verify that exactly one of the cases fits each given pair $, %. The particular
case used defines Back(, )(C ), assuming knowledge of the pullback below .
Intuitively speaking, Back(, )(C ) names the set of -sequences from which
player I can win to either enter a shift of C , or at least reach an acceptably obstructed
position along the way to C . Acceptably obstructed positions are themselves winning
for I, not in the game for C but in a different game, recorded roughly at the critical
4E Mirror images 167

point of the acceptable obstruction. This intuition will be given precise meaning in
Chapter 6.
Remark 4D.8. It is convenient sometimes to have Back(, )(C ) defined not only
for < , but also for = . Let us set Back(, )(C ) = C in the case that
= .

4E Mirror images
We work as before with a fixed transitive ZFC model M. Given two Woodin cardinals
< , and given a -name D , we intend to define the -name Back II (, )(D ). We
refer to this name as the mirrored (, )-pullback of D . Our definitions here precisely
mirror the definitions in Sections 4C and 4D.
Remark 4E.1. For the sake of symmetry we sometimes write Back I rather than plain
Back for the pullback operations defined in Sections 4C and 4D.

4E (1) Removing obstructions. We start by mirroring the definitions of Section 4C (1).


Definition 4E.2 (for W , t an annotated position of relative domain + 1, and
a relative limit in W ). t is II-suitable at just in case that for every n < , tII ( )(n) is
an e( )-functor.
Note the use of tII ( ) in Definition 4E.2, compared to the use of tI ( ) in Defini-
tion 4C.5. This is the only difference between the two definitions. Thus Definition 4E.2
places on tII ( ) the same structural requirements that were placed on tI ( ) in Defini-
tion 4C.5.
Definition 4E.3 (for W and t an annotated position of relative domain + 1).
$E,  % is a II-acceptable obstruction for t just in case that:
(1) $E,  % is a minimal M-obstruction for t, see Definition 4B.31;
(2) t is II-suitable at (E); and
(3) there exists a natural number n so that t belongs to (d)(, )[] where stands
for the ultrapower embedding by E, d stands for tII ( (E))(n), and stands for
crit(E).
Definition 4E.3 is similar to Definition 4C.6, but uses tII instead of tI . It says that
t belongs to a shifted interpretation of a name recorded as part of t. But the record
consulted here is tII ( (E)) rather than tI ( (E)). Later on when we define the mirrored
pullback operation in the case of relative limits W we shall use tII ( ) to record
names for winning positions for II, just as in Section 4C (3) we used tI ( ) to record
winning positions for I.
168 4 Pullbacks

Definition 4E.4 (for W and t an annotated position of relative domain + 1). t is


II-acceptably obstructed if there exists a II-acceptable obstruction for t.
We intuitively think of II-acceptably obstructed positions t as winning for II, just
as in the previous sections we thought of I-acceptably obstructed positions as winning
for I.
Remark 4E.5. We occasionally attach the prefix I- to the terms of Sections 4C and
4D, to distinguish them more clearly from the mirrored terms of the current section. For
example we say I-acceptably obstructed instead of acceptably obstructed, to differentiate
more clearly references to Definition 4C.7 from references to Definition 4E.4.

4E (2) Mirrored successor game. Let us next mirror the definitions of Section 4C (2).
Let be a relative successor in W . Fix a -name D . Let = e( ). Fix (in some
generic extension of M) an annotated position t of relative domain and a witness w
for .
For expository simplicity fix g which is col(, )-generic/M. Working in M[g ]
define B (M&) by:

B = {$x, y% M[g ] | x is a -code, y is a real, and x  y is either:


(a) an element of D [g ]; or (b) II-acceptably obstructed}.

Let B M be the canonical name for (M& ) B . We view B as naming a


subset of (M& ) ((M&) ) . Let B = (x, y  B [x, y]) be the mirrored
auxiliary games map associated to B , , and X = (M&) .
Define the game Hsuc (t, w, D ) to be played as follows: I and II play natural

numbers y(i), creating together the real y = $y(i) | i < %. In addition they play
moves in B [t, w, y]. Infinite runs of Hsuc (t, w, C ) are won by player I.

I y(0) b0I b1I y(2) ...


II b0II y(1) b1II ...

(t, w, D ).
Diagram 4.4. The game Hsuc

(t, w, D ) differs from the game defined in Section 4C(2) in several ways:
Hsuc
infinite runs here are won by player I; the auxiliary moves here are in the mirrored
auxiliary games associated to the name B ; and clause (b) of the definition of B asks
for II-acceptable obstructions rather than I-acceptable obstructions.
Informally Hsuc (t, w, D ) can be viewed as a game where players I and II collab-

orate to produce y, and in addition player II works against the auxiliary moves played
by I to try and witness that the annotated positions t, w, y is either:
(a) an element of D [h ] for some col(, )-generic h ; or
(b) II-acceptably obstructed.
4E Mirror images 169

Let suc (t, w, D ) be the statement player II has a winning strategy in


(t, w, D ). (t, w, D ) is the natural mirror to the formula (t, w, C )
Hsuc suc suc
of Section 4C (2). We have:

Claim 4E.6. suc (t, w, D ) is absolute between any two generic extensions of M which
have t and w as elements. #

4E (3) Relative limits and records. Let be a relative limit in W . We work to define
the mirrored (, )-pullback operations for this fixed and for W . Our
definitions here mirror the definitions in Section 4C (3).
Fix a -name Z. Let = e( ). As in Section 4C (3) we start with some supporting
definitions:

For each ordinal we define a col(, )-name S[ ] for a subset of (M& )


(M&) .

For each ordinal , each Q (M& )< , each W , and each L,


we define a -name D(, , Q, ).

These objects are defined with reference to Z and really we should call them SZ [ ] and
DZ (, , Q, ). But Z is fixed here, and so we suppress its mention in the notation.

4E(3)(a) The definition of D(. . . , ). Fix Q (M& )< . Fix W . Fix


L. We work to define D(, , Q, ).
For each < let S[ ] be the mirrored auxiliary games map associated to S[ ],
, and X = M&.
For expository simplicity fix some g which is col(, )-generic/M. Fix in M[g] an
M-clear annotated position t of relative domain + 1. Let n = lh(Q). Suppose that
n is not used in t () and let w = t () + $n, %. Let x = t, w. Let be the first
Woodin cardinal above .
(t, , , Q, ) to be the game played according to Diagram 4.5, rules
Define Hlim
(1)(3), and the payoff condition (P) below.

I bnI ... b(n 1)I


II , n n bnII ... (n 1) b(n 1)II

(t, , , Q, ).
Diagram 4.5. The game Hlim

(1) (Rule for II)

(a) is an element of W , greater than and smaller than .


(b) n is a natural number which is strictly greater than n, and not used in w.
170 4 Pullbacks

After II plays and n subject to rule (1), the game continues with n n rounds
subject to rules (2) and (3) below. We use i to index these rounds, starting with i = n.
We set bi = $biII , biI % following round i. We use Qn to denote Q and inductively
define Qi+1 = Qi , bi as the game proceeds.
(2) (Rule for II) i is an ordinal chosen so that:
(a) Qi is a legal position in S[i , x];
(b) i < if i = n; and
(c) i < i1 if i > n.
(3) (Rule for I and II) biII and biI are legal moves in the game S[i , x], following
the position Qi .
Once the game is over we let Q = Qn and let = n 1 . We set = + 1
and D = D( , , Q , ). We finally set D = Back II ( , )(D ).

(P) The run of Hlim (t, , , Q, ) described above is won by player II just in case

that suc (t, w, D ) holds true.

The definition of Hlim (. . . ) is the natural mirror to the definition of G (. . . ) in


lim
Section 4C (3). The auxiliary moves here are in the mirrored auxiliary games associated
to the names S[ ]; the roles of the players in Diagram 4.5 are precisely dual to their
roles in Diagram 4.2; the name D is defined using the mirrored pullback operation;
and the payoff condition is phrased with reference to the formula suc of Section 4E (2)
rather than the formula suc of Section 4C(2).
Definition 4E.7. Let D(, , Q, ) be the canonical name for the set of t M[g] so
that:
(1) t is an M-clear annotated position of relative domain + 1;
(2) n = lh(Q) is not used in t (); and
(t, , , Q, ).
(3) II wins the game Hlim

4E(3)(b) The definition of S[ ]. Continue to work with a fixed ordinal . We have


at our disposal the names D(. . . , ), defined above. Our next goal is to define S[ ].
Let I = {$, % ( L) ( W ) | < }.
Definition 4E.8 (for b (M& ) ). For each n < let vb,
 (n) : I M& be the
function defined by vb, 
 (n)(, ) = D(, , bn, ). Let vb,
 denote the sequence
$vb,
 (n) | n < %.


 is the II-record we attach to b and . It records, in the manner of Definition 4C.4,
vb,
the names D(. . . , Q, ) for Q-s which are initial segments of b.
4E Mirror images 171

Definition 4E.9. For expository purposes fix h which is col(, )-generic/M. Working
 x% so that:
in M[h] let S[ ] be the set of all pairs $b,

(1) b belongs to (M& ) ;

(2) x (M&) is a -code; and

(3) there exists some u (M& + 1) so that x  (u vb,


 ) is either:

(a) an element of Z[h], or


(b) II-acceptably obstructed.

Let S[ ] be the canonical name for S[ ].

Definition 4E.9 is the natural dual to Definition 4C.18. Note particularly how
condition (3a) here talks about x  (u vb,
 ), while in Definition 4C.18 the condition
talked about x  (ua, v). For t = x  (u v) we have tI ( ) = u and tII ( ) = v.
So in Definition 4C.18 the record ua, was attached on tI ( ), while here the record
vb,
 is attached on tII ( ). This distinction is in line with our declared intention in
Section 4E (1), to use tII for the mirrored definitions just as previously we used tI .

Remark 4E.10. The definitions above are inductive in several ways. They assume
knowledge of the pullback operation below , they assume knowledge of the names
D(. . . , ) for < , and they assume knowledge of the mirrored auxiliary games
maps S[ ] for < . The inductive assumptions were used in the definition of the
game Hwl (. . . , ). The rules of the game are phrased with reference to the maps S[ ]

for < . The payoff condition (P) is phrased with reference to a name D(. . . , )
for < and with reference to the ( , )-pullback operation for some < .

4E(3)(c) The pullback operation on Z. We defined so far the following objects:

For each ordinal , we defined a col(, )-name S[ ] for a subset of (M& )


(M&) .

For each ordinal , each Q (M& )< , each W , and each L,


we defined a -name D(, , P , ).

The definitions were made with reference to a fixed -name Z. (The dependence on Z
came in through condition (3b) of Definition 4E.9.)
Let $L , H % be the least pair of local indiscernibles of M relative to .

Definition 4E.11. Define Back II (, )(Z) to be the name D(, , Q, ) taken with
= 0, Q = , and = L .

This definition mirrors Definition 4C.20.


172 4 Pullbacks

4E (4) Woodin limits of Woodin cardinals. Let be a Woodin limit of Woodin car-
dinals. Fix a -name Z. We work to define Back II ( , )(Z) for W . Our
definitions here mirror those of Section 4D(1).
Let < be the least ordinal which seals Z. We make the following supporting
definition before defining the pullback operation on Z.
For each ordinal , each [, ) W , each L, and each n < we
define a -name D(, , n, ).
The name is defined with reference to Z. Since Z is fixed we suppress it in the notation.
Let us begin the definition. Fix an ordinal . Fix [, )W . Fix L. Fix
n < . For expository simplicity fix some g which is col(, )-generic/M. Fix in M[g]
an M-clear annotated position s of relative domain + 1. Suppose that n is not used in
s() and let w denote the witness s() + $n, %. We define a game Hwl (s, , , n, ) as

follows:
Players I and II play according to Diagram 4.6 and rules (1)(3) below.

I I
II , n II , n

(If s is hopeful wrt Z) (If s is not hopeful wrt Z)

(s, , , n, ).
Diagram 4.6. The game Hwl

(1) (Rule for I) < .


(2) (Rule for II) W is greater than , and greater than the first Woodin
cardinal above . n is a natural number which is not used in w.
(3) (Rule for I or II) < .

Once the game is over we let = + 1 and let D = D( , , n , ). We let be


the first Woodin cardinal above , and let D = Back II ( )(D ).
(P) The run of Hwl (s, , , n, ) described above is won by player II just in case that

suc (s, w, D ) holds true.
Remark 4E.12. The move corresponding to rule (3) is played by I if s is hopeful
with respect to Z, and by II otherwise.
The game Hwl (s, , , n, ) is the natural dual to the game G (s, , , n, ) of
wl
Section 4D (1). Note how the roles of the players in Diagram 4.6 are precisely dual to
their roles in Diagram 4.3. Note how Remark 4E.12 is precisely dual to Remark 4D.4;
Z is intuitively regarded as a payoff for II, and so it is player II who is punished here
with the burden of if s is not hopeful. Note finally how the payoff condition (P) uses
the formula suc of Section 4E(2) rather than the formula suc of Section 4C (2).
4E Mirror images 173

Definition 4E.13. Define D(, , n, ) to be the canonical name for the set of s M[g]
so that:
(1) s is an M-clear annotated position of relative domain + 1;
(2) n is not used in s(); and
(s, , , n, ).
(3) II wins the game Hwl
Recall that < is the least ordinal which seals Z. Definition 4E.13 determines
D(, , n, ) for any [, ) W , any L, any n < , and any ordinal .
As usual the definition is inductive. It assumes knowledge of the mirrored pullback
operation below , and it assumes knowledge of the names D(. . . , ) for < .
Both assumptions were used in preparing to phrase the payoff condition (P); the first in
the definition of D , and the second in the definition of D .
Let us continue to work with the fixed -name Z, and continue to use to denote
the least ordinal which seals Z. Fix now some W smaller than . For expository
simplicity fix some q which is col(, )-generic/M. Let $L , H % be the least pair of
local indiscernibles of M relative to . Let be the first Woodin cardinal above . For
each > in [, ) W let D = D(, , n, ) where = 0, n = 0, and = L .
Let B = Back II ( , )(D ).
Definition 4E.14. Define Back II ()(Z) to be the canonical name for the set of r M[q]
so that:
(1) r is an M-clear annotated position of relative domain + 1; and
(2) in M[q] there exists some witness w for + 1 and there exists some > in
[, ) W so that suc (r, w, B ) holds true.
Definition 4E.14 establishes the action of the mirrored ( , )-pullback operation
on Z. It mirrors Definition 4D.6.

4E (5) Relative successors. Fix a relative successor W . Let D be a -name.


Let be the largest Woodin cardinal below . We work to define Back II (, )(D ).
Our definitions here mirror the definitions in Section 4D (2).
Case 1. If belongs to W . For expository simplicity fix some g which is col(, )-
generic/M. Define Back(, )(D ) to be the canonical name for the set of t M[g]
so that:
(A1) t is an M-clear annotated position of relative domain + 1; and
(A2) in M[g] there exists some witness w for + 1 so that suc (t, w, D ) holds true.
# (Case 1)

Case 2. If is a Woodin limit of Woodin cardinals. Fix for expository simplicity some
G which is W -generic/M. Define Back(, )(C ) to be the canonical name for the
set of t M[G] so that:
174 4 Pullbacks

(B1) t is an M-clear annotated position of relative domain ; and

(B2) the statement there exists some witness w for so that suc (t, w, D ) holds
true is forced in col(, ) over M[G]. # (Case 2)

suc in conditions (A2) and (B2) is the formula defined in Section 4E (2).

4E (6) Compositions. So far we defined the mirrored pullback operation in the fol-
lowing three basic cases:

(I) We defined the mirrored (, )-pullback operation in the case that is a relative
limit in W and < is an element of W .

(II) We defined the mirrored (, )-pullback operation in the case that is a Woodin
limit of Woodin cardinals and < is an element of W .

(III) We defined the mirrored (, )-pullback operation in the case that is a relative
successor in W and is the largest Woodin cardinal below .
The remaining cases are defined by composition using the basic cases. Let us just
say that the compositions trivially mirror those of Section 4D (3), and omit the exact
phrasing.

4E (7) Summary. Our work in Section 4E defines a mirrored pullback operation on


names.
Suppose is a Woodin cardinal, either a Woodin limit of Woodin cardinals or an
element of W . Suppose D is a -name. Suppose that is a Woodin cardinal, again
either a Woodin limit of Woodin cardinals or an element of W , and smaller than .
Back II (, )(D ), the mirrored (, )-pullback of D , is defined under these as-
sumptions. The definition is by induction on via a process which precisely mirrors
that of Sections 4C and 4D.
Both the pullback operation of Sections 4C and 4D and the mirrored pullback
operation of the current section will be used in Chapters 5, 6, and 7. The nature of these
operations should become clearer with use.
Chapter 5
When both players lose

Fix a transitive ZFC model M. We work in M and in generic extensions of M through-


out this chapter. We follow the notational conventions of Sections 4C, 4D, and 4E. In
particular Woodin means Woodin in M. suc and suc below are the formulae defined
in Sections 4C (2) and 4E(2) respectively.
Definition (5G.2). Let be a Woodin cardinal. Let C be a -name. Define ini (, C)
to be the formula suc (t0 , w0 , C0 ) holds true with:
t0 equal to the empty annotated position of relative domain 0;
w0 equal to the empty witness for 0; and
C0 equal to Back(0 , )(C) where 0 is the first Woodin cardinal of M.
For a -name D let ini (, D) be defined similarly, but using suc and the mirrored
pullback.
If ini (, C) holds then we intuitively expect player I to be able to win to enter a
shift of Back(0 , )(C) where 0 is the first Woodin cardinal of M, and from there we
expect player I to be able to win to enter a shift of C. Similarly if ini (, D) holds then
we expect player II to be able to win to enter a shift of D. We will turn these intuitive
expectations into precise arguments in Chapter 6. Here we handle the remaining case,
the case that both ini (, C) and ini (, D) fail. This is the analogue to case (3) in
Theorems 2A.2 and 3E.1, and we produce in this case a -sequence t which avoids C
and D.
We construct t working by induction on . The induction divides into five cases
corresponding to the ones listed in Section 4D(3) and mirrored in Section 4E (6). The
three main cases are: relative limits, to be handled in Section 5C; Woodin limits of
Woodin cardinals, to be handled in Section 5D; and relative successors, to be handled
in Section 5E.

5A Saturation
Let us recall some definitions from Section 4B. A -iseq is a -length sequence of
identities  = $ ( ) | < % so that  ( ) belongs to M& for each < , and so that
the entire sequence  belongs to M. A -iseq  is realized by an annotated position t
just in case that there exists some < so that t |=  ( ). It is implicit in the statement
t |=  ( ) that the relative domain of t is at least ht( ( )). But other than this we make
176 5 When both players lose

no assumptions on the relative domain of t. It may be smaller than , and it may be


smaller than ht(
( )) for many < .
The terminology reviewed above is derived from Definition 4B.26. Without repro-
ducing that definition let us point out that the more iseqs t realizes, the less likely it is
to be obstructed. Our main aim in this section is to phrase a definition which states that
t realizes enough iseqs so as to avoid minimal unacceptable obstructions.

Definition 5A.1. An annotated position t is nice if it satisfies the following conditions:

(N1) For each rdm(t) W , t + 1 belongs to a generic extension of M by


col(, ).

(N2) For each relative limit rdm(t)W and each i < , t ( )(i) is an e( )-functor.

(N3) For each rdm(t) L and each L(t ()), t () extends t ().

(N4) For each rdm(t) L, there are infinitely many n which are not used in
t ().

Condition (N2) should be read in conjunction with definition 4C.4. It implies that t is
both I-suitable and II-suitable at all relative limits in its domain. Conditions (N3) and
(N4) should be read in conjunction with the definitions of Section 4A. Condition (N3)
should be viewed as a coherence requirement on the witnesses in t. Condition (N4)
says that the domain of t () leaves enough free room for extension.

Remark 5A.2. Definition 5A.1 depends on M. (Condition (N1) is phrased with refer-
ence to M, and the very notion of a position depends on M through the definition of W
and L.) When we wish to take note of this dependence we write nice over M instead
of nice.

Claim 5A.3. Suppose t is nice over M. Let M be a transitive model of ZFC and
suppose that M & rdm(t) is equal to M& rdm(t). Then t is nice also over M .

Proof. Only condition (N1) requires some work. Note that if t + 1 belongs to a
generic extension of M by col(, ) then the canonical name for t + 1 can be coded
as a subset of M&. The agreement between M and M is sufficient that this name
belongs also to M . #

Definition 5A.4. Suppose belongs to W . Let C and D be -names. A -sequence t


is said to avoid C and D if there exists some g so that:

(1) g is col(, )-generic/M;

(2) t belongs to M[g]; and

(3) t belongs to neither C[g] nor D[g].


5A Saturation 177

We shall refer to Definition 5A.4 often throughout this chapter. It should be read in
conjunction with Definition 4C.1. Recall that a -sequence, in the case that belongs to
W , is an M-clear annotated position of relative domain + 1. A -name is a col(, )-
name for a set of such sequences.

Definition 5A.5. A step is a triple $, , t% where:

(1) W , L, and < ;

(2) t is a -sequence;

(3) t is nice; and

(4) t is saturated (see Definition 5A.8).

We call $, , t% a -step if + 1 (the relative domain of t) is smaller than .

Definition 5A.6. Let s be a nice annotated position, and let w be a witness for rdm(s).
Let a be a finite subset of . A step $, , t% is said to extend the pair $s, w% if:

(1) t extends s strictly;

(2) rdm(s);

(3) t(rdm(s)) = w; and

(4) t() extends w.

$, , t% is said to a-extend $s, w% if in addition:

(5) no number in a is used in t().

Note particularly the last three conditions in Definition 5A.6. a-extending the pair
$s, w% requires not only finding some t which extends s, but also making sure that
t(rdm(s)) = w, and finding some rdm(s) so that t() extends w and does not use
any number in a.
We are now ready to phrase the main definition of this section:

Definition 5A.7 (for a nice annotated position t). Let be a relative limit in rdm(t)W .
Let denote e( ). t is saturated at just in case that: for every -iseq  which is not
realized by t, there is an m < , a < in L(t ()), and a finite a dom(t ()),
so that there is no -step $, , t% which satisfies the following conditions:

(1) $, , t% is an a-extension of $t, t ()%;

(2) t avoids c(, ) and d(, ), where c = tI ( )(m) and d = tII ( )(m); and

(3) t realizes  .
178 5 When both players lose

Roughly speaking Definition 5A.7 states that if a -iseq  is not realized by t,


then it is also not realized by any smaller t which is sufficiently similar to t and
appropriate. Smaller means of relative domain + 1 for some < . Similar
means satisfying the conditions of Definition 5A.6 with $s, w% given by $t, t ()% for
some L(t ()) and with a disjoint from dom(t ()). Note how these condi-
tions translate in the context of Definition 5A.7 to conditions on t which are satisfied
by t. Appropriate means satisfying conditions (2), (3), and (4) in Definition 5A.5, and
condition (2) in Definition 5A.7. With these interpretations of smaller, similar, and
appropriate, Definition 5A.7 precisely states that if a -iseq  in not realized by t then
there is some level of similarity (given by and a) and some notion of appropriateness
(given by m) so that  is also not realized by any t which is smaller than t, similar to t
to the degree given by and a, and appropriate.
Definition 5A.7 is made for a nice annotated position t. Since t is nice it certainly
satisfies condition (N2). From this it follows that c = tI ( )(m) = t ( )(2m) and
d = tII ( )(m) = t ( )(2m + 1) are functors of the kind described in Definition 4C.4.
The domain of these functors is I = {$, % ( L) ( W ) | < }. It
is easy to check that the pairs $, % which come up in Definition 5A.7 belong to
this I . From the fact that c and d are functors it follows that c(, ) and d(, ) are
-names. Condition (2) in Definition 5A.7, which talks about avoiding c(, ) and
d(, ), therefore makes sense.
Definition 5A.8 (for a nice annotated position t). t is saturated if it is saturated at all
relative limits in rdm(t) W .
Definition 5A.8 at last gives meaning to condition (4) in Definition 5A.5. Defi-
nitions 5A.7 and 5A.8 are made together by induction. To define saturation at we
require, through Definition 5A.5, knowledge of Definition 5A.8 for nice annotated po-
sitions t of relative domain smaller than = e( ). For this we require knowledge of
saturation at , but only for < = e( ) < .
Remark 5A.9. The definition of saturation is dependent on M. When we wish to
take note of this dependence we write saturated over M instead of saturated. We
similarly talk about steps over M when we wish to take note of the dependence of
Definition 5A.5 on M. We write avoids over M when we wish to take note of the
dependence of Definition 5A.4 on M.
Claim 5A.10. Let t be nice, and saturated at over M. Let M be a transitive model of
ZFC . Suppose that t is nice also over M . Let denote e( ). Suppose that M & + 1
is equal to M& + 1. Then t is saturated at also over M .
Proof. Work by induction on . Using the claim with < it follows that any -step
over M is also a -step over M and (by symmetry) vice versa. Using this and the fact
that all quantifiers in Definition 5A.7 range over M& + 1 at worst, it is easy to see
that -saturation over M is the same as -saturation over M . Let us only comment
that the most sizable quantifier in Definition 5A.7 is the first one, ranging over -iseqs.
A -iseq is a -sequence of identities in M&. The agreement between M and M is
sufficient to imply that M and M have the same -iseqs. #
5A Saturation 179

Recall the intuition behind the definition of saturation: saturation is meant to say
that t realizes enough iseqs so as to avoid minimal unacceptable obstructions. The next
lemma gives this intuition a precise meaning. It states that a minimal obstruction to a
saturated position cannot be unacceptable.

Lemma 5A.11. Let belong to W , let t be a nice annotated position of relative domain
+ 1, and suppose that t e() is M-clear. Let $E,  % be a minimal obstruction for t.
Suppose that t is saturated. Then the obstruction $E,  % is either I-acceptable, or
II-acceptable, or both.

Proof. Let = (E). (Recall that this is the first Woodin cardinal above crit(E).) Let
= e( ). Let M = Ult(M, E) and let : M M be the ultrapower embedding.
It is easy to see that e( ) = crit(E). So the critical point of is .
Since t is nice, t = t + 1 belongs to a generic extension of M by col(, ). Fix
some g, col(, )-generic/M, so that t M[g]. t is a function with domain contained
in + 1 and range contained in (M&) . From this it follows easily that the canonical
name for t can be coded as a subset of M&. Since M agrees with M on subsets of
M&, t belongs also to M [g].
Since $E,  % is an obstruction for t we have:

(i)  is a -iseq;

(ii) t does not realize  ; and

(iii) (in M [g]) t realizes (


).

) is a ()-iseq over M [g]. To see that it is


In the case of condition (iii) note that (
realized by t note that t |= ( )(), by condition (6) of Definition 4B.26.
By Definition 4B.26, E is + 1 strong. Implicit in this is the following condition
(see Fact 2 in Appendix A):

(iv) () > .

Since t is nice over M we get using the strength of E and Claim 5A.3:

(v) t is nice over M .

Since t is saturated over M we get by Claim 5A.10:

(vi) t is saturated over M .

Since $E,  % is a minimal obstruction for t we get by Definition 4B.31:

(a) t is obstruction free over M .

Combining this with the assumption that t e() is M-clear, and therefore M -clear
since M and M agree past e(), we get:

(vii) t is M -clear.
180 5 When both players lose

t is assumed saturated, and so it is certainly saturated at . Applying Definition 5A.7


with the -iseq  we obtain an m < , a < in L(t ()) and a finite a
dom(t ()) so that:
(A) For each -step $, , t%, at least one of the conditions (1)(3) in Definition 5A.7
fails.
The fact that a dom(t ()) implies that:
(viii) No number in a is used in t ().
From condition (N3) and the fact that L(t ()) it follows that:
(ix) t () extends t ().
For the record let us note that:
(x) .
We trivially have the following condition:
(xi) t extends t strictly and t () = t ().
Suppose now for contradiction that $E,  % is not I-acceptable, and not II-acceptable.
Let c = tI ( )(m) and let d = tII ( )(m). (Recall that m is the number given by
Definition 5A.7 applied with the iseq  .) Looking at Definitions 4C.6 and 4E.3 we see
that:
(b) t  (c)(, )[] and t  (d)(, )[].
(Note that t is both I-suitable and II-suitable at . This follows from condition (N2) in
Definition 5A.1.) Recall that we fixed earlier some col(, )-generic g so that t belongs
to M [g]. We did this using condition (N1) and the agreement between M and M .
Combining the fact that t belongs to M [g] with condition (b) we see that:
(xii) t avoids (c)(, ) and (d)(, ) over M .
Conditions (iii)(xii) combined directly give:
(B) Over M , $, , t% is a ()-step which satisfies each of the following conditions:
(1) $, , t% is an a-extension of $t, t ()%;
(2) t avoids (c)(, ) and (d)(, ); and
(3) t realizes (
).
But condition (B) contradicts the shift of condition (A) to M via the elementary
embedding . (For the sake of precision let us note that condition (A) can be stated
over any extension of M which has t and t () as elements. Since t is nice, and since
is smaller than , t and t () belong to a generic extension of M of size less than .
can be extended to act on this generic extension, and used to shift condition (A) to
the corresponding generic extension of M .) #
5B Successors, basic step 181

Corollary 5A.12. Let belong to W and let t be a nice annotated position of relative
domain + 1. Suppose that t e() is M-clear. Suppose that t itself is not I-acceptably
obstructed and not II-acceptably obstructed. Suppose that t is saturated. Then t is
M-clear.
Proof. Suppose for contradiction that t is not M-clear. Since t e() is M-clear this
must mean that t itself is obstructed. Using Lemma 4B.32 fix a minimal obstruction
$E,  % for t. Now apply Lemma 5A.11. The lemma says that $E,  % is I-acceptable,
or II-acceptable, or both. But this contradicts the assumption that t is not acceptably
obstructed. #

5B Successors, basic step


Let be a relative successor in W . Fix names C and D . Let = e( ). Fix, in
some generic extension of M, an annotated position t of relative domain . Suppose
that:
(S1) t is nice, saturated, and M-clear.
Fix a witness w for . Suppose that:
(S2) w is amenable to t (see Definitions 4A.14 and 4A.19), and there are infinitely
many numbers which are not used in w.
Let g be col(, )-generic/M. Suppose that:
(S3) t and w belong to M[g ].
We work under these assumptions to prove the following lemma:
Lemma 5B.1 (under the assumptions (S1)(S3)). Suppose that suc (t, w, C ) and
suc (t, w, D ) both fail. Then there exists some real y M[g ] so that t, w, y is
nice, saturated, M-clear, and belongs to neither C [g ] nor D [g ].
Regarding the statement of Lemma 5B.1 let us note that suc (t, w, C ) and
suc (t, w, D ) are absolute for generic extensions of M which have t and w as el-
ements. We can therefore simply state that suc (t, w, C ) and suc (t, w, D ) fail,
without having to say where. We will use the fact that they fail in M[g ].
Proof of Lemma 5B.1. We follow the notation of Sections 4C (2) and 4E (2). The as-
sumption that suc (t, w, C ) fails implies that I does not have a winning strategy in
Gsuc (t, w, C ). Gsuc (t, w, C ) is an open game, and hence determined. Since the
game is not won by I, it must be won by II (the closed player). Let be a winning
strategy for II in Gsuc (t, w, C ). Since winning an open game is absolute we may pick
M[g ]. Working similarly with the assumption that suc (t, w, D ) fails let us
pick a strategy M[g ] which is winning for I (the closed player) in Hsuc
(t, w, D ).
182 5 When both players lose

Let gen M[g ] be the generic strategies map associated to the name A of
Section 4C (2). Let gen M[g ] be the mirrored generic strategies map associated to
the name B of Section 4E(2).
Working inside M[g ] we construct y, a , and b so that:

(1) y = $y(n) | n < % is a real;

(2) y and a form an infinite play of Gsuc (t, w, C ) played according to both and
gen [t, w, y];

(3) y and b form an infinite play of Hsuc


(t, w, D ) played according to both and

gen [t, w, y].

The construction follows the usual lines. gen [y] produces the moves anI for all n;
produces the numbers y(n) for odd n and the moves anII for all n; produces the
numbers y(n) for even n and the moves bnI for all n; and gen [y] produces the moves
bnII for all n. As always the construction uses the Lipschitz continuity of the maps
gen and gen .
From condition (2) above and Lemma 1B.2 we get $t, w, y%  A . From this and
the definition of A in Section 4C(2) it follows that:

(i) t, w, y = t, w  y is not an element of C [g ]; and

(ii) t, w, y is not I-acceptably obstructed.

Working similarly with condition (3), Lemma 1D.2, and the definition of B in
Section 4E (2) we get:

(iii) t, w, y is not an element of D [g ]; and

(iv) t, w, y is not II-acceptably obstructed.

Claim 5B.2. t, w, y is nice.

Proof. Conditions (N1)(N4) of Definition 5A.1 are easy to verify using assump-
tions (S1)(S3) and the fact that y belongs to M[g ]. Let us only note that condi-
tion (N3) in the case = is proved using the assumption that w is amenable to t in
condition (S2), and using the assumption that t itself is nice in condition (S1). #

Claim 5B.3. t, w, y is saturated.

Proof. Saturation only imposes conditions on relative limits (see Definition 5A.8).
Since is a relative successor, the set of relative limits below rdm(t, w, y) is pre-
cisely equal to the set of relative limits below rdm(t). The saturation of t therefore
directly implies the saturation of t, w, y. #

Claim 5B.4. t, w, y is M-clear.


5C Relative limits 183

Proof. This follows from the assumption that t is M-clear in condition (S1), conditions
(ii) and (iv) above, and Claim 5B.3, using Corollary 5A.12 on the annotated position
t, w, y. #

We established now that t, w, y is nice, saturated, and M-clear. Conditions (i)


and (iii) above state that it belongs to neither C [g ] nor D [g ]. # (Lemma 5B.1)

5C Relative limits
Fix a relative limit W . Let Y and Z be -names. We plan to prove the following
lemma:

Lemma (5C.9). Let < be an element of W and let t be a nice, saturated -sequence.
Suppose that t avoids Back I (, )(Y ) and Back II (, )(Z). Then there exists a nice,
saturated -sequence s which extends t and avoids Y and Z.

Remark 5C.1. By a nice, saturated -sequence we of course mean a -sequence t which


is nice (Definition 5A.1) and saturated (Definition 5A.8) as an annotated position.

The complete proof of Lemma 5C.9 is an induction which stretches to Section 5G.
For the time being let us just say that we use the following hypothesis during the proof,
but only for < :

Hypothesis 5C.2. Let belong to W . Let C and D be -names. Let <


belong to W . Let r be a nice, saturated -sequence.
Suppose that r avoids Back I (, )(C ) and Back II ( , )(D ). Then there exists
a nice, saturated -sequence r which extends r and avoids C and D .

5C (1) Basic step. Recall that we are working with a fixed relative limit W
and fixed -names Y and Z. We follow the notation of Sections 4C (3) and 4E (3).
Section 4C (3) defines (among other things) a list of names R[ ] and associated auxiliary
games maps R[ ]. Section 4E(3) defines names S[ ] and mirrored auxiliary games
maps S[ ].
Let $L , H % be the least pair of local indiscernibles of M relative to .

Claim 5C.3. R[L ] is equal to R[H ].

Proof. Note that R[ ] is uniformly definable in M& + from the parameters:

(1) ;

(2) c0 coding the canonical equivalent to Y (by this we mean the canonical name for
the set named by Y );

(3) c1 coding the pullback operation below ;


184 5 When both players lose

(4) c2 coding the map which assigns to each canonical col(, )-name R for a subset
of (M& ) (M& e( )) its corresponding auxiliary games map R; and
(5) c3 coding the map which assigns to each relative successor < and each
canonical -name C , the canonical name in the collapse to the largest Woodin
cardinal below for the set of pairs $t, w% so that suc (t, w, C ) holds true.
The uniform definition of R[ ] from these parameters is the formalization of
Section 4C (3). Y is a col(, )-name for a set of -sequences. Its canonical equiv-
alent can be coded as an element of M& + 2. The parameters c1 and c3 are subsets of
M& . The parameter c2 is a map from M& + 2 into M& + 1. Thus all the parameters
c0 , . . . , c3 belong to M& + . The claim now follows from the local indiscernibility
of L and H . #
Claim 5C.4. For each P (M& )< , each W , and each L,
C(, , P , L ) is equal to C(, , P , H ).
Proof. This again follows from the local indiscernibility of L and H . #
Claim 5C.5. Let k < be given. There exists a sequence $i | i < k% so that:
(1) i < i1 for each i (0, k 1], and 0 < L ;
(2) for each i k 1, R[i ] is equal to R[L ]; and
(3) for each i k 1, each P (M& )< , each W , and each L,
C(, , P , i ) is equal to C(, , P , L ).
Proof. Let c4 equal R[L ]. Let c5 be the function , , P  C(, , P , L ), defined
on P (M& )< , W , and L. Note that c4 and c5 are both elements
of M& + . Let c0 , . . . , c3 be the parameters used in the proof of Claim 5C.3. Let
k ( , c0 , . . . , c5 ) be the formula which says: there exists below a descending se-
quence 0 > 1 > > k1 so that R[i ] is equal to c4 for each i k 1, and the
map , , P  C(, , P , i ) is equal to c5 for each i k 1. (The references to
R[i ] and C(. . . , i ) can be made precise using the parameters c0 , . . . , c3 .)
From the local indiscernibility of L and H it follows that k [ , c0 , . . . , c5 ] holds
for = L iff it holds for = H .
It is easy from this to prove Claim 5C.5 by induction on k. #
Remark 5C.6. Indiscernibility claims similar to Claims 5C.3 and 5C.5 can be made
for the mirrored names S[ ] and D(. . . , ) of Section 4E (3).
Fix some h which is col(, )-generic/M. Let gen M[h] be the generic strategies
map associated to the name R[L ]. Let gen M[h] (not to be confused with the fixed
Woodin cardinal W ) be the mirrored generic strategies map associated to the
name S[L ].
Fix W . Fix W . Fix n < . Fix P , Q (M& )< , both of length n,
say P = $a0 , . . . , an1 % and Q = $b0 , . . . , bn1 %. Fix an annotated position t. Suppose
that:
5C Relative limits 185

(L1) t is a nice, saturated -sequence; and

(L2) t () does not use n.

Let w denote t () + $n, %. This is a witness for + 1 = rdm(t). Suppose that:

(L3) P is a position in R[L , t, w], played according to gen [t, w];

(L4) Q is a position in S[L , t, w], played according to gen [t, w]; and

(L5) t avoids C(, , P , L ) and D(, , Q, L ).

We work under these assumptions to prove the following claim, which we will use
later on in the proof of Lemma 5C.9.

Claim 5C.7 (under the assumptions (L1)(L5)). Let < e( ) and m be given.
Suppose that m is greater than n and not used in w.
Then there is a W greater than , an n greater than m, a
L, an annotated position t , and finite sequences P and Q , so that the following

conditions, (1)(9), hold:

(1) t is a nice, saturated -sequence;

(2) t extends t strictly;

(3) t ( + 1) = w;

(4) t ( ) is an m-extension of t ( + 1) (in particular + 1);

(5) t ( ) does not use m;

(6) t ( ) does not use n .

Let w denote t ( ) + $n , %. This is a witness for + 1 = rdm(t ).

(7) P is a position of length n in R[L , t , w ], extending P , and played ac-


cording to gen [t , w ];

(8) Q is a position of length n in S[L , t , w ], extending Q, and played ac-


cording to gen [t , w ]; and

(9) t avoids C( , , P , L ) and D( , , Q , L ).

Conditions (1) and (6)(9) should be viewed as parallels of conditions (L1)(L5),


setting the grounds for an iterated use of Claim 5C.7. Conditions (2)(5) should be
viewed as stating the particular way in which the objects , , and t extend , ,
and t.
186 5 When both players lose

Proof of Claim 5C.7. Fix some g so that: g is col(, )-generic/M; t belongs to M[g];
and t belongs to neither C(, , P , L )[g] nor D(, , P , L )[g]. This is possible using
condition (L5). Using the fact that t belongs to M[g] but not to C(, , P , L )[g],
Definition 4C.14, the fact that t is a -sequence hence in particular M-clear, and as-
sumption (L2), we see that I does not win Glim (t, , , P , L ). But the game is clopen,
hence determined. So II must win.
Fix a winning strategy for player II in Glim (t, , , P , L ). Using condition (N4)
in Definition 5A.1 fix some n > m which is not used in t (). Let $i | i < n % be
given by Claim 5C.5 applied with k = n . Fix some in W , larger than the first
Woodin cardinal above and larger than . This is possibly since and are both
smaller than e( ), and is a relative limit.
, n , $i | n i < n %, gen [t, w] and combine to produce a run of

Glim (t, , , P , L ) as presented in Diagram 4.2. We leave it to the reader to verify that
this combination satisfies all the necessary rules stated in Section 4C (3). Let us only
note that one uses condition (L3) to start the construction off, and uses the fact that
R[i ] = R[L ] as the construction progresses through rounds i n.
Let $ai | n i < n % be the auxiliary moves in the run of Glim (t, , , P , L )
constructed above. Following the notation of Section 4C (3) let Pn denote P , let Pi+1 =
Pi , ai , and let P = Pn . We have:
(ii) P is a position of length n in R[L , t, w], extending P , and played according
to gen [t, w].
In phrasing this condition we are making use the indiscernibility given by Claim 5C.5.
The indiscernibility allows us to switch from R[n 1 , . . . ] to R[L , . . . ].
Continuing to follow the notation of Section 4C (3) set = + 1 and let
denote the first Woodin cardinal above . Since P is part of a run of Glim (. . . ) played
according to , and since is a winning strategy for II, we have:
(iii) suc (t, w, C ) fails with the assignment C = Back I ( , )(C ), where C =
C( , , P , L ).
This condition is essentially the negation of the payoff condition (P) in
Section 4C (3). By a strict reading of that payoff condition we should have set C =
C( , , P , n 1 ) in condition (iii). But Claim 5C.5 states that this is the same as
C( , , P , L ).
Working as we did above but with Q, the condition (L4), and the fact that t 
D(, , Q, L )[g], we obtain Q so that:
(iv) Q is a position of length n in S[L , t, w], extending Q, and played according
to gen [t, w]; and
(v) suc (t, w, D ) fails with the assignment D = Back II ( , )(D ), where D =
D( , , Q , L ).
The collapse to is absorbed by the collapse to . So we may fix g which is
col(, )-generic/M, with g M[g ]. Since t belongs to M[g] we have t M[g ].
5C Relative limits 187

Note that w then also belongs to M[g ], since it is easily obtained from t (). By
conditions (iii) and (v) we are in a position to use Lemma 5B.1. An application of the
lemma produces a nice, saturated -sequence t M[g ] so that:

(vi) t extends t and t ( + 1) = w; and

(vii) t M[g ] belongs to neither C [g ] nor D [g ].

Rephrasing condition (vii) we see that t avoids C = Back I ( , )(C ) and D =


Back II ( , )(D ). From this an application of Hypothesis 5C.2 produces a nice,
saturated -sequence t so that:

(viii) t extends t ; and

(ix) t avoids C and D .

, n , = + 1, t , P , and Q are easily seen to satisfy the requirements in


the conclusion of Claim 5C.7. Let us only comment on conditions (7) and (8). w by
definition is equal to t ( ) + $n , %. It follows from this that w is an n -extension
of t ( ). t ( ) is just t ( + 1), which is equal to t ( + 1), which in turn is equal
to w. w is thus an n -extension of w. Using Claim 4A.18 we now see that t, w
and t , w  agree to n . Conditions (7) and (8) therefore follow from conditions (ii)
and (iv). #

Remark 5C.8. The proof of Claim 5C.7 used Hypothesis 5C.2. We used the hypothesis
with = , r = t , and = . was chosen smaller than . So we only needed
the hypothesis for < .

5C (2) Construction. We continue to work with a fixed relative limit W , and


fixed -names Y and Z. The next lemma achieves the intention stated at the start of
Section 5C:

Lemma 5C.9 (using Hypothesis 5C.2, but only for < ). Let < be an element
of W . Let t be a nice, saturated -sequence. Suppose that t avoids Back I (, )(Y ) and
Back II (, )(Z). Then there exists a nice, saturated -sequence s which extends t and
avoids Y and Z.

Proof. By condition (N1) in Definition 5A.1, t belongs to a generic extension of M by


col(, ). col(, ) is absorbed into col(, ). We can therefore fix some h which is
col(, )-generic/M with t M[h]. We work inside M[h] throughout the proof.
We follow the notation of Section 5C(1). gen M[h] is the generic strategies
map associated to the name R[L ]. gen M[h] is the mirrored generic strategies map
associated to the name S[L ]. R and S are defined in Sections 4C (3) and 4E (3).
By assumption t avoids Back I (, )(Y ) and Back II (, )(Z). Looking at Defini-
tions 4C.20 and 4E.11 we see that:

(i) t avoids C(0, , , L ) and D(0, , , L ).


188 5 When both players lose

Starting with condition (i) and using repeated applications of Claim 5C.7 one can
easily construct an s which satisfies the requirements of Lemma 5C.9, except that s may
not be saturated at and obstruction free. We expect saturation to inhibit obstructions
using Corollary 5A.12. So the key to the construction is saturation at . We intend to
secure saturation by making s realize as many e( )-iseqs as possible.
Let denote e( ). Let J be the set of all -iseqs.
Claim 5C.10. The set {
J | no extension of t realizes  } is infinite.
Proof. A -iseq  = $ ( ) | < % is constant if  ( ) =  (0) for all < . Let I be
the set of constant -iseqs  so that ht( (0)) + 1 and t |=  (0). It is easy to see
that I is infinite. The claim follows; if  belongs to I then it cannot be realized by any
extension of t. #
J is a subset of M& + 1, and hence countable in M[h]. Fix in M[h] some enu-
meration $k | k < % of J .
= e( ) is also countable in M[h]. Since is a relative limit in W , is a limit
ordinal. Fix in M[h] some increasing sequence $k | k < % so that k < for each k,
and supk< k = .
We work to construct for each k < the objects listed in conditions (A)(E) below:
(A) k W and k L with k < k ;
(B) a nice, saturated k -sequence tk ;
(C) a natural number nk which is not used in tk (k ).
Let wk denote tk (k ) + $nk , k %. This is a witness for k + 1 = rdm(tk ).
(D) a position Pk of length nk in R[L , tk , wk ], played according to the generic
strategy gen [tk , wk ]; and
(E) a position Qk of length nk in S[L , tk , wk ], played according to the generic
strategy gen [tk , wk ].
Let 0 = 0, let 0 = , let n0 = 0, let P0 = Q0 = , and let t0 = t. This assignment
trivially satisfies the above condition. We continue the construction by induction on k.
Having defined the objects of conditions (A)(E) for k we fix:
(F) a number mk which is not used in tk (k ), with mk > nk .
This is possible by condition (N4) in Definition 5A.1, which states that there are infinitely
many numbers not used in tk (k ). We then define the objects of conditions (A)(E) for
k + 1 according to one of the two cases described below. In both cases we make sure
that:
(1) tk+1 extends tk strictly;
(2) tk+1 (k + 1) = wk ;
5C Relative limits 189

(3) tk+1 (k+1 ) is an mk -extension of tk+1 (k + 1) (in particular k+1 k + 1);

(4) Pk+1 extends Pk ;

(5) Qk+1 extends Qk ; and

(6) tk+1 avoids C(k+1 , k+1 , Pk+1 , L ) and D(k+1 , k+1 , Qk+1 , L ).

Case 1. If there exists an assignment which satisfies conditions (A)(E) for k + 1,


satisfies conditions (1)(6) above, and satisfies in addition the case condition (I) below.
In this case we fix some assignment which satisfies all these conditions.

(I) nk+1 = mk ; and tk+1 realizes k . # (Case 1)

Case 2. Otherwise. Fix in this second case an assignment which satisfies conditions
(A)(E) for k + 1, satisfies conditions (1)(6) above, and satisfies in addition:

(II) k+1 > k ; nk+1 > mk ; and mk is not used in tk+1 (k+1 ). # (Case 2)

The construction described by the two cases above is a matter of folding a bias
for realizing -iseqs into the natural construction suggested by Claim 5C.7. Note that
for each k the objects k , k , Pk , Qk , and tk satisfy the assumptions (L1)(L5) in
Section 5C (1). Using Claim 5C.7 it is always possible to find an assignment for k + 1
which satisfies the requirements of case 2. (By routine absoluteness it is possible to
pick this assignment in M[h].) But the construction first searches for an assignment
which satisfies the conditions of case 1. Only if no such assignment is found does the
construction retreat to a use of Claim 5C.7 to obtain the conditions of case 2.
Let us develop some properties of the objects constructed above. Some of these
properties are trivial and we list them for the record, without proof.
Claim 5C.11. The sequence $mk | k < % is strictly increasing. #
Claim 5C.12. For each k < , wk+1 is an mk -extension of wk .

Proof. wk+1 by definition is equal to tk+1 (k+1 ) + $nk+1 , k+1 %. tk+1 (k+1 ) is an mk -
extension of wk by conditions (2) and (3). nk+1 mk by the case conditions (I) and
(II). The claim follows. #

Define T1 by setting k T1 iff the objects of conditions (A)(E) for k + 1 were


constructed according to case 1. The next claim formalizes the bias towards case 1 in
the construction.
Claim 5C.13 (for k < ). Suppose that there is a W , a < in L, a nice,
saturated -sequence t, and P and Q in (M& )< , so that the following conditions,
(C1)(C8), hold:

(C1) t extends tk strictly;

(C2) t(k + 1) = wk ;
190 5 When both players lose

(C3) t() is an mk -extension of wk ;

(C4) t() does not use mk .

Let w = t() + $mk , %. This is a witness for + 1 = rdm(t).

(C5) P is a position of length mk in R[L , t, w], extending Pk and played according
to gen [t, w];

(C6) Q is a position of length mk in S[L , t, w], extending Qk and played according
to gen [t, w]; )

(C7) t avoids C(, , P , L ) and D(, , P , L ); and

(C8) t realizes k .

Then k belongs to T1 .

Proof. Simply note that the conditions assumed in the claim correspond precisely to
the requirements of case 1 of the construction. The use of mk in conditions (C4)(C6)
is not an error; it corresponds to the requirement nk+1 = mk in the case condition (I).
#

Claim 5C.14 (for k < ). If k T1 then tk+1 realizes k .

Proof. If k T1 then the case condition (I) holds for tk+1 . #

Define T2 by setting k T2 iff the objects of conditions (A)(E) for k + 1 were


constructed according to case 2. (In other words define T2 = T1 .)
Claim 5C.15. T2 is infinite.

Proof. By Claim 5C.10 there are infinitely many k < so that no extension of t realizes
k . Every such k must belong to T2 , since it couldnt possibly admit an assignment
which satisfies the case condition (I). #

Claim 5C.16. supk< k = supk< k = .

Proof. The fact that supk< k = follows from Claim 5C.15 and the requirement
k+1 > k in the case condition (II). The fact that supk< k = supk< k follows from
conditions (A) and (3). #

Let t = k< tk . The union makes sense by condition (1). t is a nice annotated
relative domain . For each k we have t (k + 1) = wk by condition (2).
position of 
Let w = k< wk . The union makes sense by Claim 5C.12. w is a witness for ,
amenable to t (see Definition 4A.19).
Claim 5C.17. t is M-clear.
5C Relative limits 191

Proof. For each k < , tk is a k -sequence and so in particular M-clear. It is easy to


check that the increasing union of M-clear annotated positions is M-clear. The claim
follows. #

Claim 5C.18 (for k < ). w is an mk -extension of wk .

Proof. Immediate from Claims 5C.11 and 5C.12. #

Claim 5C.19. Suppose k T2 . Then mk is not used in w .

Proof. By the case condition (II), mk is not used in tk+1 (k+1 ) and nk+1 > mk . From
this it follows that mk is not used in wk+1 = tk+1 (k+1 ) + $nk+1 , k+1 %. Since w is an
mk+1 -extension of wk+1 , and since mk+1 > mk , it follows that mk is not used in w .
#

Let x = t , w . Using Claims 5C.18 and 4A.18 we see that x and tk , wk 
agree to mk . By condition (F), mk is greater than nk . So x and tk , wk  certainly agree
to nk . From this and condition (D)it follows that Pk is a position in R[L , x ], played
according to gen [x ]. Let a = k< Pk . The union makes sense by condition (4).
a is an infinite run of R[L , x ], played according to gen [x ]. Using Lemma 1B.2 it
follows that $ a , x %  R[L ][h]. Looking at Definition 4C.18 we conclude that:

(iv) For every v (M& + 1) M[h], x  (ua,L v) is neither:

(a) an element of Y [h], nor


(b) I-acceptably obstructed.

ua,L here is the I-record attached to a and L , see Definition 4C.17.



Let b = k< Qk . By reasoning similar to the above but using conditions (E) and
(5), Lemma 1D.2, and Definition 4E.9 we get:

(v) For every u (M& + 1) M[h], x  (u vb,


 ) is neither:
L

(a) an element of Z[h], nor


(b) II-acceptably obstructed.

vb,
 here is the II-record attached to b and L , see Definition 4E.8.
L
Let y = ua,L vb,
 . Let s = x  y . This is the same as setting s =
L
t , w , y .

Claim 5C.20. s M[h] is a nice annotated position of relative domain + 1, extend-


ing t.
192 5 When both players lose

Proof. t is an annotated position of relative domain = e( ), w is a code for ,


and y belongs to (M& + 1) . From this it follows that s = t , w , y is an
annotated position of relative domain + 1. s extends t0 by construction, and t0 was
set equal to t, so s extends t. s belongs to M[h] since the entire construction takes
place inside M[h]. It is easy to verify that s is nice. Let us only make two comments
on this matter. From Claims 5C.15 and 5C.19 it follows that there are infinitely many
numbers not used in w , and this secures condition (N4) in Definition 5A.1 for = .
Condition (N3) in the case = is proved using the fact that w is amenable to t ,
and the fact that t itself satisfies condition (N3). #
Claim 5C.21. s is saturated at .
Proof. This claim is the reason for the inclusion of case 1 in the construction above.
Any -iseq which is not realized by s corresponds to a stage in the construction where
case 2 was taken. In other words it corresponds to a stage where objects satisfying the
conditions of case 1 could not be found. We shall use this to argue for saturation.
Fix a -iseq  . Let us verify the statement of Definition 5A.7 for this particular
iseq. If  is realized by s then Definition 5A.7 requires nothing. So we may assume:
(a)  is not realized by s.
Using the fact that $k | k < % enumerates J we may find k < so that  is
equal to k . By condition (a), k is not realized by s and so certainly not realized by
any initial segment of s. In particular k is not realized by tk+1 . Using Claim 5C.14 it
follows that:
(b) k T2 .
From this and Claim 5C.19 it follows that mk is not used in w . This in turn implies
that {mk } {i < mk | i  dom(w )} is (a finite set) contained in dom(w ) =
dom(s()).
Let m = mk , let = k + 1, and let a = {mk } {i < mk | i  dom(w )}. We
prove that this particular assignment witnesses the truth of the existential statement in
Definition 5A.7, for the iseq k .
Fix a -step $, , t%. We have to verify that at least one of the conditions in
Definition 5A.7 fails. Suppose otherwise. In other words suppose that:
(c) $, , t% is an a-extension of $s, s()% = $tk , wk %;
(d) t avoids c(, ) and d(, ) where c = ua,L (m) and d = vb,
 (m); and
L

(e) t realizes k .
We work to obtain a contradiction from these conditions.
In translating from condition (2) in Definition 5A.7 to condition (d) here we already
factored in the fact that s( ) = y and y = ua,L vb,  . When we further factor
L
in Definitions 4C.17 and 4E.8 we get:
5C Relative limits 193


(f) t avoids C(, , a m, L ) and D(, , bm, L ).
Recall that a was set equal to {mk } {i < mk | i  d(w )}. By Claim 5C.18, w is
an mk -extension of wk . So an i < mk is used in w iff it is used in wk . Thus:
(g) a = {mk } {i < mk | i  dom(wk )}.
Using condition (g) and the fact that $, , t% is an a-extension of $tk , wk % it is easy
to check that , and t satisfy conditions (C1)(C4) in Claim 5C.13. Following the
notation of Claim 5C.13 let w = t() + $mk , %. w is an mk -extension of t(), which
by condition (C3) is an mk -extension of wk . So w is an mk -extension of wk . Using
this, condition (C1), and Claim 4A.18 it follows that t, w agrees with tk , wk  to mk .
We saw already that tk , wk  agrees with x to mk . So t, w and x agree to mk .
It is now easy to verify that conditions (C5) and (C6) of Claim 5C.13 hold for , , t,
P = a mk , and Q = bm k . The remaining conditions, (C7) and (C8), follow from
conditions (e) and (f). Applying Claim 5C.13 we get k T1 . But this contradicts
condition (b) above. # (Claim 5C.21)
Claim 5C.22. s is saturated.
Proof. s is saturated below since each tk is saturated. s is saturated at by Claim 5C.21.
#
So far we proved that s M[h] is a nice annotated position of relative domain +1,
extending t, and saturated. Let us now continue to show that s is M-clear, and that s
avoids Y and Z. Conditions (iv) and (v) above provide the starting point. Applying
condition (iv) with v = vb,
 we get:
L

(vi) s  Y [h]; and


(vii) s is not I-acceptably obstructed.
Applying condition (v) with u = ua,L we get:

(viii) s  Z[h]; and


(ix) s is not II-acceptably obstructed.
Claim 5C.23. s is M-clear.
Proof. Use Corollary 5A.12 on the nice annotated position s. The assumptions of Corol-
lary 5A.12 hold for s by Claim 5C.17, conditions (vii) and (ix) above, and Claim 5C.22.
#
By construction s extends t0 which is equal to t. Claims 5C.20, 5C.22, and 5C.23
combine to state that s is a nice, saturated -sequence. Conditions (vi) and (viii)
combined with the fact that s M[h] state that s avoids Y and Z. So s satisfies the
requirements in the conclusion of Lemma 5C.9. # (Lemma 5C.9)
194 5 When both players lose

Remark 5C.24. Note that Claim 5C.7 is used in the proof of Lemma 5C.9. The claim
is needed to guarantee the existence of objects satisfying the requirements of case 2,
which is used as a fall-back option during the construction. Hypothesis 5C.2 is used
through the appeal to Claim 5C.7. But the hypothesis is only needed for < , by
Remark 5C.8.

5D Woodin limits of Woodin cardinals


Our intention next is to prove a parallel of Lemma 5C.9 for the case of a Woodin limit
of Woodin cardinals.

Definition 5D.1. Suppose is a Woodin limit of Woodin cardinals. Let C and D be


-names. A -sequence t is said to avoid C and D if it belongs to neither C[G(t)] nor
D[G(t)].

Definition 5D.1 is a companion to Definition 5A.4. There is no conflict between


the two since Woodin limits of Woodin cardinals are excluded from W . The filter
G(t) referred to in Definition 5D.1 is the one given by Definition 4B.23. In our
context, the context of an annotated position t of relative domain , G(t) is equal to
{[ ] | is an identity in M& and t |= }. When t is a -sequence (hence in particular
M-clear) and is a Woodin limit of Woodin cardinals, G(t) is W -generic/M by Corol-
lary 4B.30. Moreover t belongs to M[G(t)] by Claim 4B.41. The requirement in
Definition 5D.1 that t belongs to neither C[G(t)] nor D[G(t)] is therefore non-trivial;
t does belong to the generic extension M[G(t)].
Fix a Woodin limit of Woodin cardinals . Let Y and Z be -names. We work
with these fixed objects for the rest of Section 5D. We shall use Hypothesis 5C.2, but
only for < . We follow throughout the notation of Sections 4D (1) and 4E (4). In
particular < is the least ordinal which seals Y and < is the least ordinal which
seals Z. C(, , n, ) and D(, , n, ) are the names determined by Definitions 4D.5
and 4E.13. The former is defined for [, ) W , L, n < , and ON.
The latter is defined for [, ) W and , n, and as before.

5D (1) Basic step. Let $L , H % be the least pair of local indiscernibles of M relative
to .

Claim 5D.2. For each [, ) W , each L, and each n < , C(, , n, L )


is equal to C(, , n, H ).

Proof. Note that C(, , n, ) is uniformly definable in M& + from the parameters:

(1) ;

(2) c0 coding the canonical equivalent to Y ;


5D Woodin limits of Woodin cardinals 195

(3) c1 coding the pullback operation below ; and


(4) c2 coding the map which assigns to each relative successor < and each
canonical -name C , the canonical name in the collapse to the largest Woodin
cardinal below for the set of pairs $t, w% so that suc (t, w, C ) holds true.
The definition of C(, , n, ) from these parameters is simply the formalization of the
definition given in Section 4D(1). With the exception of all the parameters listed
are elements of M& + . The claim now follows from the local indiscernibility of L
and H . #
Claim 5D.3. For each [, ) W , each L, and each n < , D(, , n, L )
is equal to D(, , n, H ).
Proof. Similar to the proof of the previous claim. #
Fix now some [max{, }, ) W , some L, some n < , and some
annotated position s of relative domain + 1. Since the relative domain of s is greater
than and we can ask whether s is hopeful with respect to each of Y and Z. Let us
here work under the following assumptions:
(W1) s is not hopeful, not with respect to Y and not with respect to Z;
(W2) s is a nice, saturated -sequence;
(W3) s() does not use n; and
(W4) s avoids C(, , n, L ) and D(, , n, L ).
Claim 5D.4 (under the assumptions (W1)(W4)). Let < and m < be given.
Then there is a W greater than , an n greater than m, and an annotated
position s so that:
(1) s extends s strictly;
(2) s ( + 1) = s() + $n, %;
(3) s is not hopeful, not with respect to Y and not with respect to Z;
(4) s is a nice, saturated -sequence;
(5) n not used in s ( + 1); and
(6) s avoids C( + 1, , n , L ) and D( + 1, , n , L ).
Condition (3) in the conclusion of Claim 5D.4 is actually redundant; it follows from
conditions (W1) and (1) by Claim 4D.3. We include condition (3) in the statement
of Claim 5D.4 only to make clearer the connection between the assumptions of the
claim and its conclusion. Conditions (3)(6) precisely match assumptions (W1)(W4),
setting the grounds for an iterated use.
196 5 When both players lose

Proof of Claim 5D.4. Using Claims 5D.2 and 5D.3 we can switch from L to H and
rephrase condition (W4) to say that s avoids C(, , n, H ) and D(, , n, H ). We may
thus fix some g so that:
g is col(, )-generic/M;
s belongs to M[g]; and
s belongs to neither C(, , n, H )[g] nor D(, , n, H )[g].
Using the fact that s belongs to M[g] but not C(, , n, H )[g], Definition 4D.5, the
fact that s is a -sequence, and condition (W3), we see that:
(i) II wins the game Gwl (s, , , n, H ) of Section 4D (1).
Similar reasoning using Definition 4E.13 and the fact that s does not belong to
D(, , n, H )[g] shows that:
(s, , , n, ) of Section 4E (4).
(ii) I wins the game Hwl H

Gwl (s, , , n, H ) is a finite game consisting of moves , , n , and . The rules


of the game are such that condition (i) translates to the following:
(iii) There exists some < ; so that for every and n which satisfy rule (2) in
Section 4D(1); and for every < H ; the payoff condition (P) in Section 4D (1)
fails.
Remark 5D.5. We are assuming here that s is not hopeful with respect to Y . The
move in Gwl (s, , , n, H ) therefore corresponds to player I, see Diagram 4.3 and
Remark 4D.4. The phrasing of condition (iii) takes account of this fact; , like the
other moves for I, is bounded by a universal quantifier in condition (iii).
Condition (ii) similarly translates to:
(iv) There exists some < ; so that for every and n which satisfy rule (2) in
Section 4E(4); and for every < H ; the payoff condition (P) in Section 4E (4)
fails.
Here too is bounded by a universal quantifier, taking account of the fact that s is not
hopeful with respect to Z.
Let a and b respectively witness the existential statements in conditions (iii) and
(iv). Following the notation of Sections 4D(1) and 4E (4) let be the first Woodin
cardinal greater than , let w = s()+$n, %, and let = +1. Using the fact that is a
limit of Woodin cardinals fix some W which is greater than max{a , b }, greater
than , and greater than . Using condition (N4) in Definition 5A.1 fix some n <
which is not used in s(), larger than m, and distinct from n. Since w = s() + $n, %
it follows that n is also not used in w.
Condition (iii), applied with the objects and n fixed above and with = L ,
implies that:
5D Woodin limits of Woodin cardinals 197

(v) suc (s, w, C ) fails with the assignment C = Back I ( , )(C ), where C =
C( , , n , L ).
Condition (iv), similarly applied with the objects and n fixed above and with
= L , implies that:
(vi) suc (s, w, D ) fails with the assignment D = Back II ( , )(D ), where D =
D( , , n , L ).
Condition (v) and (vi) allow for an application of Lemma 5B.1. The lemma pro-
duces a nice, saturated -sequence s which extends s and avoids C and D , with
s ( + 1) = w. An application of Hypothesis 5C.2 then produces a nice, saturated
-sequence s which extends s and avoids C and D . It is easy to check that , n ,
and s satisfy the conditions stated in Claim 5D.4. #
Remark 5D.6. The proof of Claim 5D.4 uses Hypothesis 5C.2 with = , r = s ,
and = . Since was chosen smaller than , Hypothesis 5C.2 is only needed for
< .

5D (2) Impossibility. Let us again work with [max{, }, ) W , L,


n < , and an annotated position s of relative domain + 1. But let us now assume
that:
(I1) s is hopeful, both with respect to Y and with respect to Z.
(I2) s is a nice, saturated -sequence; and
(I3) s() does not use n.
Note how assumption (I1) differs from the assumption (W1) in Section 5D (1).
Let a , b ON be given. Assume further that:
(I4) s avoids C(, , n, a ) and D(, , n, b ).
Claim 5D.7 (under the assumptions (I1)(I4)). There is a W , an n , an
annotated position s , and ordinals a and b so that:
(1) s extends s strictly;
(2) s ( + 1) = s() + $n, %;
(3) s is hopeful, both with respect to Y and with respect to Z;
(4) s is a nice, saturated -sequence;
(5) n is not used in s ( + 1);
(6) s avoids C( + 1, , n , a ) and D( + 1, , n , b ); and
(7) a < a and b < b , strictly.
198 5 When both players lose

Claim 5D.7 is a parallel of Claim 5D.4 for the case that s is hopeful with respect
to Y and Z. Here too condition (3) is redundant, since it follows from conditions (I1)
and (1) using Claim 4D.3. We include condition (3) in the statement of Claim 5D.7 to
make clearer the connection between the conclusion of the claim and its assumptions.
The claim is specifically geared for an iterated use.

Proof of Claim 5D.7. From assumption (I4) we get:

(i) II wins the game Gwl (s, , , n, a ) of Section 4D (1); and


(s, , , n, ) of Section 4E (4).
(ii) I wins the game Hwl b

The rules of Gwl (s, , , n, a ) are such that condition (i) translates to:

(iii) There exists some < ; so that for every and n which satisfy rule (2) in
Section 4D(1); there exists some < a ; so that the payoff condition (P) in
Section 4D(1) fails.

Remark 5D.8. We are assuming in Claim 5D.7 that s is hopeful with respect to Y .
It follows from this assumption that the move in Gwl (s, , , n, a ) corresponds to
player II, see Diagram 4.3 and Remark 4D.4. In phrasing condition (iii) we therefore
bound by an existential quantifier.
Condition (ii) similarly translates to:

(iv) There exists some < ; so that for every and n which satisfy rule (2) in
Section 4E(4); there exists some < b ; so that the payoff condition (P) in
Section 4E(4) fails.

Here too is bounded by an existential quantifier, taking account of the fact that s is
hopeful with respect to Z.
Let a and b respectively witness the outer existential statements in conditions (iii)
and (iv). Following the notation of Sections 4D (1) and 4E (4) let be the first Woodin
cardinal greater than , let w = s() + $n, %, and let = + 1. Using the fact that
is a limit of Woodin cardinals fix some W which is greater than max{a , b },
and greater than . Using condition (N4) in Definition 5A.1 fix some n < which
is distinct from n and not used in s(). Since w = s() + $n, % it follows that n is
also not used in w.
Condition (iii), applied with the objects and n fixed above, states the existence
of an ordinal < a which makes condition (P) in Section 4D (1) fail. Let a be some
such ordinal. Then:

(v) suc (s, w, C ) fails with the assignment C = Back I ( , )(C ), where C =
C( , , n , a ).

Condition (iv), similarly applied with the objects and n fixed above, produces some
ordinal b < b so that:
5D Woodin limits of Woodin cardinals 199

(vi) suc (s, w, D ) fails with the assignment D = Back II ( , )(D ), where D =
D( , , n , b ).

Using conditions (v) and (vi), Lemma 5B.1 produces a -sequence s which extends
s and avoids C and D , with s ( + 1) = w. Hypothesis 5C.2 then produces a -
sequence s which extends s and avoids C and D . It is easy to check that , n , s ,
a , and b satisfy the requirements in the conclusion of Claim 5D.7. #

The reader may wish to carefully compare the proof of Claim 5D.7 to that of
Claim 5D.4. In the proof of Claim 5D.7 we have the ordinals a and b given to
us by conditions (iii) and (iv). In the proof of Claim 5D.4 on the other hand it is up
to us to pick when applying conditions (iii) and (iv). This difference between the
two cases traces back to the definitions in Sections 4D (1) and 4E (4), specifically to
Remarks 4D.4 and 4E.12.

Corollary 5D.9. There are no objects which satisfy assumptions (I1)(I4).

Proof. Fix , , n, s, and a , b ON. Suppose for contradiction that these objects
satisfy assumptions (I1)(I4). Iterated applications of Claim 5D.7 produce, among other
things, two sequences of ordinals $ka | k < % and $kb | k < %. By condition (7) of
the claim these sequences are descending, giving the desired contradiction. #

Claim 5D.10. Let belong to [max{, }, ) W . Let s be a nice, saturated, -


sequence. Suppose that s avoids C(0, , 0, L ) and D(0, , 0, L ). Then it cannot be
that s is hopeful with respect to both Y and Z.

Proof. This is just the special case of Corollary 5D.9 corresponding to = 0, n = 0,


and a = b = L . #

5D (3) Another impossibility. Again let us work with [max{, }, ) W ,


L, n < , and an annotated position s of relative domain + 1. Conditions (W1)
and (I1), which we considered above, covered the cases that s is either hopeful with
respect to neither one of Y and Z, or hopeful with respect to both. Let us next consider
the case that s is hopeful with respect to one but not the other. Let us assume that:

(J1) s is hopeful with respect to Y , but not hopeful with respect to Z.

(J2) s is a nice, saturated -sequence; and

(J3) s() does not use n.

Let a ON be given. Assume further that:

(J4) s avoids C(, , n, a ) and D(, , n, L ).

Note how we use the given ordinal a in our reference to C(. . . ), and the local indis-
cernible L in our reference to D(. . . ).
200 5 When both players lose

Claim 5D.11 (under the assumptions (J1)(J4)). There is a W , an n , an


annotated position s , and an ordinal a so that:

(1) s extends s strictly;

(2) s ( + 1) = s() + $n, %;

(3) s is hopeful with respect to Y but not hopeful with respect to Z;

(4) s is a nice, saturated -sequence;

(5) n is not used in s ( + 1);

(6) s avoids C( + 1, , n , a ) and D( + 1, , n , L ); and

(7) a < a strictly. #

We omit the proof of Claim 5D.11. It is simply a matter of crossing the proof of
Claim 5D.4 with the proof of Claim 5D.7.

Corollary 5D.12. There are no objects which satisfy assumptions (J1)(J4).

Proof. This is similar to Corollary 5D.9. The proof of Corollary 5D.9 produced two
infinite descending sequences of ordinals. Here repeated use of Claim 5D.11 produces
only one infinite descending sequence of ordinals. But this is still enough for a contra-
diction. #

Claim 5D.13. Let belong to [max{, }, ) W . Let s be a nice, saturated, -


sequence. Suppose that s avoids C(0, , 0, L ) and D(0, , 0, L ). Then it cannot be
that s is hopeful with respect to Y and not hopeful with respect to Z.

Proof. This is the special case of Corollary 5D.12 corresponding to = 0, n = 0, and


a = L . #

The next claim is a dual to Claim 5D.13. Its proof, which we omit, involves phrasing
a dual to Claim 5D.11 to deal with the case that s is hopeful with respect to Z but not Y .

Claim 5D.14. Let belong to [max{, }, ) W . Let s be a nice, saturated, -


sequence. Suppose that s avoids C(0, , 0, L ) and D(0, , 0, L ). Then it cannot be
that s is hopeful with respect to Z and not hopeful with respect to Y . #

Remark 5D.15. Claims 5D.10, 5D.13, and 5D.14 all use Hypothesis 5C.2. The hy-
pothesis comes in through Claim 5D.7 and its parallels. Note that the hypothesis is only
needed for < .
5D Woodin limits of Woodin cardinals 201

5D (4) Construction. We now combine the steps of Sections 5D (1) through 5D (3)
above to prove Lemma 5D.19, a parallel of Lemma 5C.9 for the case of a Woodin limit
of Woodin cardinals.
Definition 5D.16. A -sequence t is supernice if it is nice and if in addition there exists
a witness w for rdm(t) so that:
(1) w is amenable to t; and
(2) there are infinitely many numbers which are not used in w.
Remark 5D.17. If t is supernice then a witness w satisfying the conditions of Defini-
tion 5D.16 can be found in any generic extension of M which has t as an element and
collapses rdm(t) to be countable.
Remark 5D.18. If rdm(t) is a successor ordinal, + 1 say, then a witness w satisfying
the conditions of Definition 5D.16 exists trivially; simply take w = {$0, %}. Nice
-sequences are therefore trivially supernice when W . The two notions only differ
when is a Woodin limit of Woodin cardinals.
Recall that we are working with a fixed Woodin limit of Woodin cardinals , and
fixed -names Y and Z. Lemma 5D.19 is phrased for these objects.
Lemma 5D.19 (using Hypothesis 5C.2, but only for < ). Let < be an element
of W . Let r be a nice, saturated -sequence. Suppose that r avoids Back I ( , )(Y ) and
Back II ( , )(Z). Then there exists a supernice, saturated -sequence t which extends
r and avoids Y and Z.
Proof. Using the assumption that r avoids Back I ( , )(Y ) and Back II ( , )(Z) fix
some q so that:
q is col(, )-generic/M;
r belongs to M[q]; and
r belongs to neither Back I (, )(Y )[q] nor Back II ( , )(Z)[q].
Let be the first Woodin cardinal above . Recall that < is the least ordinal
which seals Y and < is the least ordinal which seals Z. Let < be the first
element of W which is greater than and greater than max{, }. Let w = {$0, %}.
This is a witness for + 1.
Let C = C(0, , 0, L ) and let A = Back I ( , )(C). We know that r does not
belong to Back I (, )(Y )[q]. Looking at Definition 4D.6 we see that this certainly
implies:
(i) suc (r, w, A) fails.
Let D = D(0, , 0, L ) and let B = Back I ( , )(D). Working as we did above
but with Definition 4E.14 and the fact that r  Back II ( , )(Z)[q] we get:
202 5 When both players lose

(ii) suc (r, w, B) fails.


Conditions (i) and (ii) put us in a position to apply Lemma 5B.1. Using that lemma
we obtain a nice, saturated -sequence r which extends r and avoids A and B. Now
A = Back I ( , )(C) and B = Back II ( , )(D). Using Hypothesis 5C.2 on r we
obtain next a nice, saturated -sequence s which extends r and avoids C and D. (The
hypothesis is used with = . The use is valid since is smaller than .)
So far we worked under the assumptions of Lemma 5D.19 to construct an s satis-
fying:
(iii) s is a nice, saturated -sequence;
(iv) s extends r; and
(v) s avoids C and D.
s is a -sequence, and is larger than and . We can thus ask whether or not
s is hopeful with respect to each of Y and Z. Using condition (v) we may apply
Claims 5D.10, 5D.13, and 5D.14. These claims together rule out the possibility that s
is hopeful with respect to either Y , or Z, or both. Thus:
(vi) s is not hopeful, not with respect to Y and not with respect to Z.
Fix, in some generic extension of M, a sequence $k | k < % of ordinals so that
k < for each k, and supk< k = . We work to construct for each k < the
following objects:
(A) k W and k L with k < k ;
(B) a nice, saturated k -sequence sk ; and
(C) a natural number nk which is not used in sk (k ).
We make sure that:
(1) sk is not hopeful, not with respect to Y and not with respect to Z; and
(2) sk avoids C(k , k , nk , L ) and D(k , k , nk , L ).
Set 0 = , 0 = 0, and n0 = 0 to begin with. Conditions (1) and (2) for k = 0 follow
from conditions (vi) and (v).
We continue by induction on k, using repeated applications of Claim 5D.4. In stage
k we use the claim with = k and m = 2k so as to ensure:
(vii) k+1 > k ; and
(viii) nk+1 > 2k.
Having used the claim to obtain k+1 , nk+1 , and sk+1 we set k+1 = k + 1. Conditions
(A)(C), (1), and (2) for k+1 follow from the conditions in the conclusion of Claim 5D.4.
The conditions in Claim 5D.4 also state that:
5D Woodin limits of Woodin cardinals 203

(ix) sk+1 extends sk strictly; and

(x) sk+1 (k+1 ) = sk (k ) + $nk , k %.



Let t = k< sk . t is an annotated position of relative domain supk< k . supk< k
is precisely equal to by conditions (A) and (vii). Thus t is an annotated position of
relative domain . t is nice, saturated, and M-clear because it is the increasing union
of positions which are nice, saturated, and M-clear. So:
Claim 5D.20. t is a nice, saturated -sequence. #
It remains to show that t is supernice, and avoids Y and Z.
Claim 5D.21. t is supernice.

Proof. Let w = k< sk (k ). We claim that w witnesses the conditions of Defini-
tion 5D.16. It is easy to check, directly with Definition 4A.19, that w is amenable to t.
Working inductively with condition (x) one can check that w is in fact precisely equal to
{$nk , k % | k < }. So the numbers used in w are precisely the numbers nk for k < .
By condition (viii) the set {nk | k < } is thin in the sense that it has at most k + 1
elements below 2k + 1. From this is follows that there are infinitely many numbers in
the complement of {nk | k < }. In other words there are infinitely many numbers
which are not used in w. #

Recall that G(t) is the filter given by Definition 4B.23. It is equal to {[ ] |


is an identity in M& and t |= }. We already established that t is M-clear. Since
is a Woodin limit of Woodin cardinals it follows from Corollary 4B.30 that G(t) is
W -generic/M. It follows from Claim 4B.41 that t belongs to M[G(t)].
Claim 5D.22. t does not belong to Y [G(t)].

Proof. Suppose for contradiction that t = t[G(t)] belongs to Y [G(t)]. Recall that
is the least ordinal which seals Y . Looking at Definition 4B.43 we see that there must
exist some condition [ ] in G(t) (M&) which forces t Y . From this it follows
that there must exist some good identity in M& so that t |= and [ ] W t Y .
t extends s0 = s which has relative domain + 1 > . Since belongs to M& it has
height smaller than . So t |= iff s |= . In conclusion we found a good identity
M& so that s |= and [ ] W t Y . But then s is hopeful with respect to Y
(see Definition 4D.2), contradicting condition (vi). #

Claim 5D.23. t does not belong to Z[G(t)].

Proof. Similar to the last claim, using this time the fact that s is not hopeful with respect
to Z, given again by condition (vi). #

We established already that t is a supernice, saturated -sequence. t extends s


which in turn extends r. The last two claims demonstrate that t avoids Y and Z. This
completes the proof of Lemma 5D.19. # (Lemma 5D.19)
204 5 When both players lose

5E Relative successors
Fix a relative successor W . Let C and D be -names. We work to prove the
following Lemma:

Lemma 5E.1. Let be the largest Woodin cardinal below . Let t be a nice (supernice
if is a Woodin limit of Woodin cardinals), saturated -sequence. Suppose that t
avoids Back I (, )(C ) and Back II (, )(D ). Then there exists a nice, saturated
-sequence t which extends t and avoids C and D .

Proof. We divide the proof of Lemma 5E.1 into two cases, depending on whether is
an element of W or a Woodin limit of Woodin cardinals. In both cases we follow the
notation of Sections 4D(2) and 4E(5).

Case 1. Suppose first that is an element of W . Since t avoids Back I (, )(C ) and
Back II (, )(D ), there exists some g which is col(, )-generic/M, and such that:

(i) t belongs to M[g];

(ii) t  Back I (, )(C )[g]; and

(iii) t  Back II (, )(D )[g].

Let w = {$0, %}. This is a (rather trivial) witness for + 1, the relative domain of t.
The definitions in Section 4D(2) are such that from conditions (i) and (ii) it follows
that:

(iv) suc (t, w, C ) fails.

(We are using here the negation of condition (A2) in Section 4D (2), applied with the par-
ticular witness w listed above.) Working similarly with the definitions in Section 4E (5)
it follows from condition (iii) that:

(v) suc (t, w, D ) fails.

It is now easy to complete the proof of Lemma 5E.1 using an application of


Lemma 5B.1. # (Case 1)

Case 2. Suppose next that is a Woodin limit of Woodin cardinals. Let G denote
the filter associated to t by Definition 4B.23. More precisely let G equal {[ ] |
is an identity in M& and t |= }. t is assumed to be a -sequence, hence in particular
M-clear. is a Woodin limit of Woodin cardinals. By Corollary 4B.30, G is W -
generic/M. Moreover by Claim 4B.41:

(i) t belongs to M[G].

By Definition 5D.1 the fact that t avoids Back I (, )(C ) and Back II (, )(D ) says
that:
5F Compositions 205

(ii) t  Back I (, )(C )[G]; and


(iii) t  Back II (, )(D )[G].
Let g be col(, )-generic/M[G]. Using the assumption that t is supernice, and
using Remark 5D.17, fix some w so that:
(iv) w is a witness for ; w is amenable to t; there are infinitely many numbers not
used in w; and
(v) w belongs to M[G][g].
The definitions in Section 4D(2) are such that from conditions (i), (ii), and (v) it
follows that:
(v) suc (t, w, C ) fails.
(We are using here the negation of condition (B2) in Section 4D (2), applied with the par-
ticular witness w found above.) Working similarly with the definitions in Section 4E (5)
it follows from condition (iii) that:
(vi) suc (t, w, D ) fails.
It is now easy to complete the proof of Lemma 5E.1 using an application of
Lemma 5B.1. We leave the exact details of this to the reader and only note that both
W and col(, ) are absorbed into col(, ), so t and w can both be absorbed into an
extension M[g ] where g is col(, )-generic/M. # (Case 2)
# (Lemma 5E.1)

5F Compositions
Recall that the pullback and mirrored pullback operations are defined using five cases.
The cases are listed in Section 4D(3) and mirrored in Section 4E (6). Our work so
far handles situations which correspond to the three basic cases. Let us now handle
situations which correspond to the remaining two cases, listed as cases (IV) and (V) in
Section 4D (3).
Lemma 5F.1. Let be a relative successor in W . Let C and D be -names.
Suppose that Hypothesis 5C.2 holds for < . Let be the largest Woodin cardinal
below .
Let < be an element of W . Let r be a nice, saturated -sequence. Suppose that
r avoids Back I (, )(C ) and Back II (, )(D ). Then there exists a nice, saturated
-sequence t which extends r and avoids C and D .
Proof. Let C = Back I (, )(C ) and let D = Back II (, )(D ). The definition in
case (IV) of Section 4D(3) is such that Back I ( , )(C ) is equal to Back I ( , )(C).
Similarly Back II (, )(D ) is equal to Back II ( , )(D). Thus:
206 5 When both players lose

(i) r avoids Back I (, )(C) and Back II (, )(D).

Applying Hypothesis 5C.2 or Lemma 5D.19, depending on whether is an element of


W or a Woodin limit of Woodin cardinals, we obtain a nice (supernice if is a Woodin
limit of Woodin cardinals), saturated -sequence t which extends r and avoids C and
D. Then using Lemma 5E.1 we obtain a nice, saturated -sequence t which extends
t and avoids C and D . #

Remark 5F.2. Let us briefly take note of the use of Hypothesis 5C.2 in the proof of
Lemma 5F.1. If is an element of W then the proof makes a direct use of the hypothesis,
with = , to obtain t. If is a Woodin limit of Woodin cardinals then the proof makes
an indirect use of the hypothesis through Lemma 5D.19. In that case the hypothesis is
only needed for < , since this is all that Lemma 5D.19 requires.

Lemma 5F.3. Let be a Woodin limit of Woodin cardinals. Let be the first Woodin
cardinal above . Let be a Woodin cardinal greater than . Let C and D be
-names. Suppose that Hypothesis 5C.2 holds for < .
Let t be a supernice, saturated -sequence. Suppose t avoids Back I (, )(C ) and
Back II (, )(D ). Then there exists a nice (supernice if is a Woodin limit of Woodin
cardinals), saturated -sequence t which extends t and avoids C and D .

Proof. Let C = Back I ( , )(C ) and let D = Back II ( , )(D ). The definition
in case (V) of Section 4D(3) is such that Back I (, )(C ) is equal to Back I (, )(C ).
Similarly Back II (, )(D ) is equal to Back II (, )(D ). Thus:

(i) t avoids Back I (, )(C ) and Back II (, )(D ).

Applying Lemma 5E.1 we obtain a nice, saturated -sequence t which extends t and
avoids C and D . Rewriting the last part we see that:

(ii) t avoids Back I ( , )(C ) and Back II ( , )(D ).

From this we can obtain a nice (supernice if is a Woodin limit of Woodin cardinal),
saturated -sequence which extends t and avoids C and D . We obtain t through
an application of: Lemma 5C.9 if is a relative limit in W ; Lemma 5D.19 if is a
Woodin limit of Woodin cardinals; Lemma 5E.1 if is a relative successor and equal
to the first Woodin cardinal above ; and Lemma 5F.1 if is a relative successor and
greater than the first Woodin cardinal above . #

Remark 5F.4. Let us take note of the use of Hypothesis 5C.2 in the proof of Lemma 5F.3.
The use of the hypothesis is always indirect. If is a relative limit in W then the
hypothesis comes in through its use in Lemma 5C.9. If is a Woodin limit of Woodin
cardinals then the hypothesis comes in through its use in Lemma 5D.19. If is a relative
successor and greater than the first Woodin cardinal above then the hypothesis comes
in through its use in Lemma 5F.1. In all cases the hypothesis is only needed for < .
5G Conclusion 207

5G Conclusion
The lemmas of the previous sections combine to yield the following theorem:

Theorem 5G.1. Let be a Woodin cardinal. Let C and D be -names. Let be a


Woodin cardinal smaller than .
Let t be a supernice, saturated -sequence. Suppose t avoids Back I (, )(C )
and Back II (, )(D ). Then there exists a supernice, saturated -sequence t which
extends t and avoids C and D .

Proof. We work by induction on . Note that Theorem 5G.1 subsumes Hypothe-


sis 5C.2. Indeed the hypothesis is simply the special case of the theorem where both
and belong to W . (By Remark 5D.18, nice and supernice are the same in this special
case.) Thus when proving Theorem 5G.1 for a particular we may inductively assume
that:

(H) Hypothesis 5C.2 holds for < .

Let us now fix , C , D , and t satisfying the assumptions of the theorem, and
work to obtain t .

Case (I). If is a relative limit in W and belongs to W . In this case we obtain t


through an application of Lemma 5C.9, with = . By condition (H) we have enough
of Hypothesis 5C.2 to allow this use of Lemma 5C.9. The lemma produces a nice,
saturated -sequence t which extends t and avoids C and D . The reader may be
slightly worried that Theorem 5G.1 asks for a supernice -sequence, and Lemma 5C.9
only provides a nice t . But for W the two notions are the same, by Remark 5D.18.
(Similar reasoning on the move from nice to supernice applies also in cases (III)(V)
below, but we shall not comment on it anymore.) #

Case (II). If is a Woodin limit of Woodin cardinals and belongs to W . In this case
we obtain t through an application of Lemma 5D.19. Condition (H) provides enough
of Hypothesis 5C.2 to allow this use of the lemma. #

Case (III). If is the first Woodin cardinal above . In this case we obtain t through
an application of Lemma 5E.1. #

Case (IV). If is a relative successor in W , is an element of W , and is strictly


larger than the first Woodin cardinal above . In this case we obtain t through an
application of Lemma 5F.1. (Our current plays the role of in that lemma.) As usual
condition (H) provides enough of Hypothesis 5C.2 to allow this use of the lemma. #

Case (V). If is a Woodin limit of Woodin cardinals, and is strictly larger than
the first Woodin cardinal above . In this case we obtain t through an application of
Lemma 5F.3. Again condition (H) provides enough of Hypothesis 5C.2 to allow this
use of the lemma. #
208 5 When both players lose

The cases listed above correspond to the cases listed in Section 4D (3) and mirrored
in Section 4E (6). It is easy to check that they are mutually exclusive, and exhaust all
possible layouts for and . # (Theorem 5G.1)

We have been working in M and in generic extensions of M throughout this chapter.


It should be noted that Theorem 5G.1 only asserts the existence of t in some generic
extension of M. No assertion is made about its existence in V. A similar comment
applies to Corollary 5G.3 below.

Definition 5G.2. Let be a Woodin cardinal. Let C be a -name. Define ini (, C) to


be the formula suc (t0 , w0 , C0 ) holds true with:
t0 equal to the empty annotated position of relative domain 0;
w0 equal to the empty witness for 0; and
C0 equal to Back(0 , )(C) where 0 is the first Woodin cardinal of M.
For a -name D let ini (, D) be defined similarly, but using suc and the mirrored
pullback.

Corollary 5G.3. Let be a Woodin cardinal. Let C and D be -names. Suppose


that the formulae ini (, C) and ini (, D) both fail. Then there exists a supernice,
saturated -sequence t which avoids C and D.

Proof. Let 0 , t0 , and w0 be as in Definition 5G.2. Let C0 = Back I (0 , )(C) and let
D0 = Back II (0 , )(D). The assumption that ini (, C) and ini (, D) both fail tells
us that:

(i) suc (t0 , w0 , C0 ) and suc (t0 , w0 , D0 ) both fail.


Using this condition, an application of Lemma 5B.1 produces a real y0 so that:
(ii) t0 , w0 , y0 is a nice (supernice by Remark 5D.18) 0 -sequence; and
(iii) t0 , w0 , y0 avoids C0 = Back I (0 , )(C) and D0 = Back II (0 , )(D).
t0 , w0 , y0 is trivially saturated, since there are no relative limits below 0 + 1. If
is equal to 0 we are done; we can simply take t = t0 , w0 , y0 . Otherwise an appli-
cation of Theorem 5G.1 produces a supernice, saturated -sequence t which extends
t0 , w0 , y0 and avoids C and D. #

Corollary 5G.3 at last achieves the goal stated at the start of this chapter. We will
apply it later on with C and D naming sets whose union includes all -sequences in the
appropriate generic extension of M. Using the theorem we will be able to argue that at
least one of ini (, C) and ini (, D) holds.
Chapter 6
Along a single branch

In Chapter 4 we defined a pullback operation on names. We worked in a transitive


model M of ZFC . Given Woodin cardinals < in M, and given a -name C , we
defined a -name Back I (, )(C ), called the (, )-pullback of C . Intuitively we
thought of Back I (, )(C ) as naming the set of -sequences from which player I can
win to either enter an interpretation of a shift of C , or reach a I-acceptable obstruction
along the way.
We now work to make this intuition precise.
In Section 6A we define precisely a game through which players I and II collaborate
to create an embedding j and a j ( )-sequence t . Is goal in the game, roughly
speaking, is to either make t enter an interpretation of j (C ), or reach a I-acceptable
obstruction during the construction of t . For the rest of the chapter we work on
constructing a winning strategy for I in this game, starting from t in an interpretation of
the (, )-pullback of C . The end products of our work are the results in Section 6G.
Theorem 6G.1 in particular is the precise meaning of our intuitive expectations from
the pullback operation.

6A The game
Let M be a transitive model of ZFC . Let < be Woodin cardinals of M. Let C
be a -name in M. Let t be a -sequence over M. We take this section to describe a
game we denote G branch = G branch (M, t, )(C ). The game, which is played in V,
sets rules for the construction of the following objects through a collaboration between
players I and II:
a wellfounded model M ;
an elementary embedding j : M M ; and
a j ( )-sequence t over M , extending t.
The format of the collaboration is for the most part standard: I plays iteration trees
of length ; II picks branches through those trees; and the two players work together
to produce the real part of t . But every once in a while we allow exceptions to this
standard format. The exceptions come in through rules (L5) and (L6) below. These
rules let player II insert intervals of her choice into t , and insert extenders of her choice
into the iteration which creates M and j .
The construction of M , j , and t may hit some snags along the way. For example
it may reach an illfounded model, or it may reach an obstruction it cannot remove,
210 6 Along a single branch

or it may reach a limit which is not countable. Conditions (I1)(I7) below spell out
the possible snags and decide the winner in the case of a construction which is left
incomplete because of a snag.
A complete construction is one which ends through the payoff condition (P1) or
the payoff condition (P2) below. It produces a model M = M+1 , an embedding
j = j0,+1 : M M+1 , and a j0,+1 ( )-sequence t = t+1 over M+1 . Player I
wins a complete construction just in case that t belongs to an interpretation of j (C ).
Let us now begin the precise description of the rules of G branch (M, t, )(C ). We
set M0 = M and t0 = t to begin with. The game starts with mega-round 0, which is
played according to the rules of the successor case below.
Successor case. At the start of a (zero or) successor mega-round we have, either
through the initial settings above or through the work of the players in mega-rounds
prior to , the following objects:
a wellfounded model M ;
an elementary embedding j0, : M0 M ; and
an annotated M -position t .
These objects are such that:
(i) t is M -clear; and
(ii) rdm(t ) < j0, ( ).
It follows from condition (ii) that there are Woodin cardinals of M above rdm(t ). Let
denote the least such. Then j0, ( ).
Mega-round proceeds according to the following rules:
(S1) I picks a witness w for rdm(t ) over M ;
(S2) I and II collaborate in the usual fashion to produce a real y ;
(S3) I plays a length iteration tree T on M , with critical points above rdm(t ) and
using only extenders which are countable in V (see Appendix A); and
(S4) II plays a cofinal branch b through T .
Let Q be the direct limit along the branch b of T and let k : M Q be the
direct limit embedding.
branch ends and player I wins.
(I1) If Q is illfounded then G
From now on assume that Q is wellfounded. The restriction in rule (S3) is such
that crit(k ) > rdm(t ). From this it follows that t may be regarded as an annotated
position over Q . Similarly w may be regarded as a witness over Q . Working over
Q set t = t , w , y . This is an annotated position, of relative domain k ( ) + 1,
over Q .
6A The game 211

kkk
/ M kSk
SkSSSS k / Q = M+1 /

T

Diagram 6.1. The successor case.

branch ends. Player I wins just in case that there


(I2) If t is obstructed over Q then G
exists a I-acceptable obstruction for t over Q .

From now on assume that t is obstruction free over Q . Set M+1 = Q , j,+1 =
k , and t+1 = t . Combining condition (i) above with the fact that t is obstruction
free over Q we see that t+1 is M+1 -clear. So condition (i) above holds for +1. Let
+1 = k ( ). The relative domain of t+1 is +1 + 1. Note that +1 j0,+1 ( )
since j0, ( ).

(P1) If +1 = j0,+1 ( ) then G branch = G


branch (M, t, )(C ) ends. Player I wins
just in case that there exists some g so that:

(1) g is col(, j0,+1 ( ))-generic/M+1 ; and


(2) t+1 j0,+1 (C )[g].

If +1 < j0,+1 ( ) then we pass to mega-round + 1. # (Successor case)

At the start of a limit mega-round we have, through the work of the players in
previous mega-rounds, the following objects:

wellfounded models M for < ;

elementary embeddings j, : M M for < < ; and

annotated M -positions t for < .

The objects constructed in mega-rounds prior to are such that:

(a) t is M -clear for each < ;

(b) rdm(t ) < j0, ( ) for each < ;

(c) the sequence $t | < % is strictly increasing; and

(d) crit(j, ) rdm(t ) for each < and all (, ).

Let M be the direct limit of the system $M , j, | < %, and let j, for
< be the direct limit embeddings.
branch ends and player I wins.
(I3) If M is illfounded then G
212 6 Along a single branch

branch ends and player I loses.


(I4) If = 1V then G
There is no need to specify priority between conditions (I3) and (I4), since they cannot
both hold: if = 1V then M is the direct limit of 1V wellfounded models, and so by
necessity wellfounded itself.
Suppose from now on that M is wellfounded, and that is smaller than 1V . From
condition (d) it follows that:
(e) crit(j, ) rdm(t ) for each < .
Fromthis it follows that t may be regarded as an annotated position over M . Let
t = < t . The union makes sense by condition (c). t is an annotated M -position.
Using conditions (a) and (b) we see that:
(f) t is M -clear; and
(g) rdm(t ) j0, ( ).
Using condition (c) and the fact that is a limit ordinal we get:
(h) rdm(t ) is a limit of Woodin cardinals in M .
The rules for mega-round divide into two cases, depending on whether or not rdm(t )
itself is Woodin.
Phantom limit case. If rdm(t ) is a Woodin cardinal in M . There are no moves in
mega-round in this case. We simply set M+1 = M , j,+1 = id, and t+1 = t .
Let and +1 equal rdm(t ). +1 is a Woodin limit of Woodin cardinals in M+1 . We
have +1 j0,+1 ( ) by condition (g). (Note that j0,+1 = j0, , since j,+1 = id.)
(P2) If +1 = j0,+1 ( ) then G branch = G
branch (M, t, )(C ) ends. Player I wins
just in case that there exists some G so that:
(1) G is j0,+1 (W )-generic/M+1 ; and
(2) t+1 j0,+1 (C )[G].

If +1 < j0,+1 ( ) then we pass to mega-round + 1. # (Phantom limit case)


Standard limit case. If rdm(t ) is not a Woodin cardinal in M . Let = rdm(t ). We
have j0, ( ) by condition (g). Equality is impossible since j0, ( ) is Woodin in
M , and is not. So < j0, ( ) strictly. It follows that there are Woodin cardinals
of M above . Let denote the least such. We have j0, ( ).
Mega-round , in the case that rdm(t ) is not Woodin in M , starts according to
the following rules:
(L1) I picks a witness w for over M ;
(L2) I plays a length iteration tree T on M , with critical points larger than and
using only extenders which are countable in V; and
6A The game 213

(L3) II plays a cofinal branch b through T .


Let Q be the direct limit along the branch b of T and let k : M Q be the
direct limit embedding.
branch ends and player I wins.
(I5) If Q is illfounded then G
Assume from now on that Q is wellfounded. Mega-round continues according
to the following rule:
(L4) I picks y in (Q & + 1) .
The restriction in rule (L2) is such that crit(k ) > = rdm(t ). t may thus be
regarded as an annotated position over Q . Working over Q set s = t , w , y .
s is an annotated position, of relative domain k ( ) + 1, over Q .
branch ends
(I6) If s is obstructed, but not I-acceptably obstructed over Q , then G
and player I loses.

Suppose from now on that s is either obstruction free or I-acceptably obstructed


over Q . Mega-round proceeds in one of two ways.
If s is obstruction free over Q then player II may elect an early end to mega-
round . In this case we set M+1 = Q , j,+1 = k , and t+1 = s . We let
+1 = k ( ). The relative domain of t+1 is +1 + 1. We have +1 j0,+1 ( )
since j0, ( ). If +1 = j0,+1 ( ) then the game ends with the payoff condition
(P1) stated above. If +1 < j0,+1 ( ) then we pass to mega-round + 1.
If s is obstructed, or if player II does not elect an early end, then mega-round
continues with what we call a leap. The leap consists of moves by player II alone,
subject to rules (L5) and (L6) below:
(L5) II plays E subject to the following conditions:
(1) E is a k ( ) + 1-strong extender in some model Q which agrees with
Q to k ( ) + 1;
(2) E is countable in V; and
(3) crit(E ) = .

M and Q agree past by rule (L2). Folding this into the conditions in rule (L5)
it follows that M and Q agree past the critical point of E . So E can be applied
to M . Let M+1 = Ult(M , E ) and let j,+1 be the ultrapower embedding. The
situation is illustrated in Diagram 6.2.
branch ends and player I wins.
(I7) If M+1 is illfounded then G
Suppose from now on that M+1 is wellfounded. Note that the agreement between
M+1 and M is sufficient that s may be regarded as an annotated position over M+1 .
Let W+1 denote the class W of Section 4A, computed in M+1 . Let u be the
even half of y . More precisely let u = $y (2i) | i < %.
214 6 Along a single branch

j,+1

kkk E &
/ M kSk
SkSSSS k / Q _ _ _ _ _ _ Q

M+1 /

T

Diagram 6.2. Standard limit case with a leap according to rules (L5) and (L6).

(L6) II plays +1 j,+1 ( ) W+1 , and an annotated position t+1 of relative


domain +1 + 1 over M+1 , subject to the following conditions:

(1) t+1 extends s (perhaps not strictly);


(2) t+1 is M+1 -clear; and
(3) there exists some n < so that u (n ) is a -functor in the sense of M ,
and so that t+1 c+1 ( , +1 )[] where c+1 = j,+1 (u (n )).

Remark 6A.1. Rule (L6) lets player II choose t+1 . But player I gets to regulate this
choice using the even half of her move y in rule (L4). The even half of y affects the
possible choices of t+1 through condition (3) in rule (L6).
Note that the condition makes sense: u (n ) is a -functor in the sense of M . It
is thus a function with domain equal to the set I = ${, % ( L ) ( W ) |
< }. Using the fact that +1 belongs to j,+1 ( ) W one can check that
$ , +1 % belongs to j,+1 (I ), namely to the domain of c+1 . c+1 is a functor over
M+1 . It therefore assigns a col(, )-name to each pair $, % in its domain. So the
reference to c+1 ( , +1 )[] in condition (3) makes sense.
Rule (L6) ends mega-round in the case of a leap. We pass to mega-round + 1
equipped with the annotated M+1 -position t+1 chosen by II subject to this rule. Note
that rdm(t+1 ) must be smaller than j0,+1 ( ). This follows from the requirement
+1 j,+1 ( ) W+1 in rule (L6) and the fact that is smaller than j0, ( ).
# (Standard limit case)

The successor case and two limit cases complete the description of the game
branch (M, t, )(C ). A winning strategy for I in G
G branch (M, t, )(C ) should be
viewed as a construction mechanism. Our goal in this chapter is to create the mecha-
nism. We intend to prove:

Theorem (6G.1). Suppose that M& + 1 is countable in V. Suppose that t belongs


to an interpretation of Back I (, )(C ), where the pullback is computed in M. Then
player I has a winning strategy in G branch (M, t, )(C ).

We will use the construction mechanism given by this theorem as part of our work
in the next chapter.
6A The game 215

Remark 6A.2. We think of positions in G branch (M, t, )(C ) as being given by se-
quences of the form P = $T , b , E , t +1 | < %. Strictly speaking not all the
objects in this sequence are defined for each . For example E is only defined if is a
standard limit, and only if II elected a leap in mega-round . T and b are only defined
if is a successor or a standard limit.
Let us for uniformity adopt the convention that T is the tree which consists entirely
of padding if is a phantom limit, and b in this case is the unique branch through T .
That way E is the only object which need not be defined for all . Note that the
convention fits with the settings in phantom limit cases. j, +1 is the identity in such
cases, so we may think of it as the direct limit embedding through a tree which consists
entirely of padding.
There are moves in mega-round which are not listed among the objects T , b ,
E , and t +1 . But all the moves in mega-round can be recovered from (t and) these
objects. Thus a sequence P = $T , b , E , t +1 | < % gives a complete account of
a position of length in G branch (M, t, )(C ).
branch (M, t, )(C ) gives rise to:
Remark 6A.3. A position P of length in G
models M for ;
embeddings j, : M M for < ; and
annotated positions t over M .
We refer to the sequence $M , j, , t | < %, which consists of the entire array
of models, embeddings, and annotated positions, as the history of P . We refer to the
triple $M , j0, , t %, which consists only of the final model, embedding, and annotated
position, as the outcome of P .
Remark 6A.4. Let $M , j0, , t % be the outcome of a position of length in
branch (M, t, )(C ). Then:
G
(1) t is M -clear; and
(2) if < 1V then j0, preserves countability (see the section on extenders in
Appendix A).
branch and, in the case of condi-
Both conditions are easily implied by the structure of G
tion (2), Fact 3 and Claim 4 in Appendix A. Note that a position of length = 1V is
lost by I subject to the snag (I4). So condition (2) holds for all outcomes of positions
which are not lost by I.

6A (1) How to leap. We describe here a scenario through which player II may obtain
legal moves for rules (L5) and (L6) in the standard limit case above. The point of the
description is to record the scenario for future reference, and also to shed some light
on these two rules. We shall see that they fit tightly with the notion of I-acceptable
obstructions.
216 6 Along a single branch

Let P be a non-terminal position of limit length in G branch (M, t, )(C ). Let


$M , j0, , t % be the outcome of P . Let = rdm(t ) and let be the first Woodin
cardinal of M above .
Suppose is not Woodin in M , so that mega-round of G branch (M, t, )(C )
following P is played subject to the rules of the standard limit case. Let w , T , b ,
and y be legal moves corresponding to rules (L1)(L4). Suppose that these moves are
non-terminal. More precisely suppose that they do not bring about the end of G branch
through one of the snags (I5) and (I6).
Following the notation of the standard limit case let Q be the direct limit along
the branch b of T , and let k : M Q be the direct limit embedding. Let
s = t , w , y . This is an annotated position of relative domain k ( )+1 over Q .

Lemma 6A.5. Let Q be a transitive model which agrees with Q to k ( ) + 1. Let


W denote the class W of Section 4A, computed in Q .
Let be an element of W , and let t be an annotated position of relative domain
+ 1 over Q . Suppose that t is I-acceptably obstructed over Q . Let $E ,  % be

an obstruction witnessing this. Suppose that:

(1) Q & + is countable in V;

(2) t extends s (perhaps not strictly);

(3) every strict initial segment of t is Q -clear (t itself of course is obstructed by


the initial assumptions);

(4) crit(E ) is equal to ; and

(5) Ult(M , E ) is wellfounded.

Let +1 = and let t+1 = t . Then E , +1 , and t+1 satisfy the demands of
rules (L5) and (L6).

Proof. We work through a series of claims to establish that the conditions in rules (L5)
and (L6) hold for the objects of the lemma. We repeatedly use the fact that $E ,  % is
a I-acceptable obstruction for t over Q .
Claim 6A.6. is greater than or equal to k ( ).

Proof. This follows from the fact that t extends s . #

Claim 6A.7. E is a + 1-strong extender in Q , and belongs to Q & + .

Proof. $E ,  % is by assumption an obstruction for t over Q . It therefore satisfies


the conditions of Definition 4B.26. The current claim follows from condition (1) in that
Definition. #

Corollary 6A.8. E is countable in V, and at least k ( ) + 1-strong in Q .


6A The game 217

Proof. This follows immediately from the last two claims, and the assumption in
Lemma 6A.5 that Q & + is countable in V. #

Claim 6A.9. Q and M agree past .

Proof. M agrees with Q past , since all extenders in T have critical points above
by the conditions in rule (L2). By assumption Q agrees with Q up to k ( ) + 1,
which is greater than . So Q and M agree past . #

Let M+1 = Ult(M , E ). The ultrapower makes sense by the previous claim. Let
j,+1 : M M+1 be the ultrapower embedding. Let W+1 denote the class W of
Section 4A, computed in M+1 .
Claim 6A.10. belongs to j,+1 ( ) W+1 .

Proof. The strength of E given by Claim 6A.7 is such that M+1 and Q agree to
+ 1. It follows from this that W and W+1 are the same up to + 1. So belongs
to W+1 .
The ultrapower embedding by E sends its critical point past the strength of E
(see Fact 2 in Appendix A). So j,+1 ( ) > + 1. It follows from this and from the
conclusion of the previous paragraph that belongs to j,+1 ( ) W+1 . #

Claim 6A.11. t is an annotated position over M+1 , and all strict initial segments of
t are M+1 -clear.

Proof. This follows from the initial assumptions about t over Q , and the fact that
M+1 agrees with Q to + 1 = rdm(t ). #

Claim 6A.12. t is obstruction free over M+1 .

Proof. $E ,  % is a I-acceptable obstruction for t over Q . It therefore satisfies the


conditions in Definition 4C.6. In particular it is a minimal obstruction for t over Q .
So t is obstruction free over Ult(Q , E ). The claim follows from this and from the
standard agreement between Ult(Q , E ) and M+1 = Ult(M , E ). #

Corollary 6A.13. t is M+1 -clear.

Proof. This is simply the conjunction of the last two claims. #

Let u be the even half of y . Precisely, let u = $y (2i) | i < %.


Claim 6A.14. For each n < , u (n) is a -functor in the sense of M .

Proof. Since Q and M agree past it is enough to check that each u (n) is a
-functor in the sense of Q . This property of u follows from the fact that $E ,  % is a
I-acceptable obstruction for t over Q , using the suitability requirement in condition (2)
of Definition 4C.6. #
218 6 Along a single branch

Let : Q Ult(Q , E ) be the ultrapower embedding of Q by E . Recall that


j,+1 is the ultrapower embedding of M by E . Since both and j,+1 are created
through an ultrapower by E , they agree on subsets of crit(E ) = . In particular
they agree on -functors.
Claim 6A.15. There exists n < so that t belongs to j,+1 (c )( , )[], where
c stands for u (n ).

Proof. One last time we use the fact that $E ,  % is a I-acceptable obstruction for t
over Q , and therefore satisfies the conditions in Definition 4C.6. The current claim
follows directly from condition (3) in that definition, using the observation above that
and j,+1 agree on -functors. #

Equipped with the claims and corollaries above it is easy to go over the conditions
in rules (L5) and (L6), and verify that they hold true for E , +1 = , and t+1 = t .
# (Lemma 6A.5)

Lemma 6A.5 provides a scenario through which player II may obtain legal moves
for a leap in G branch . The lemma shows a connection between leaps and I-acceptable
obstructions. It was the crucial assumption that E is part of a I-acceptable obstruction
for t that carried us through the proof.

6A (2) The skipping game. We noted earlier that the main goal of this chapter is to
construct winning strategies for player I in various instances of G branch . As a main
step toward this goal we will first construct winning strategies in a related game which
allows player I to skip through rounds of G branch . We now define this skipping game.
Work as before with a transitive model M; Woodin cardinals < in M; a -
sequence t over M; and a -name C in M. Define G skip (M, t, )(C ) to be the
game which is played according to the rules of G branch (M, t, )(C ), except that the
successor mega-round is played as follows:

Successor case for skipping games. Let be a successor (or zero). Mega-round
begins with rules (S1)(S4), listed above in the successor case of G branch . This produces
w , y , T , and b . Following the definitions in the successor case let Q be the direct
limit along the branch b of T and let k be the direct limit embedding. If Q is
illfounded then following condition (I1) in the successor case the game ends and I wins.
Assuming that Q is wellfounded, let t = t , w , y . If t is obstructed over Q
then the game ends in the manner of condition (I2) in the successor case. Suppose now
that t is obstruction free over Q .
Player I has two options at this point. She can declare an early end to mega-
round . In this case we simply continue to follow the successor case in the definition
of G branch . We set M+1 = Q , j,+1 = k , t+1 = t , and +1 = k ( ). If

+1 = j0,+1 ( ) then the game ends subject to condition (P1) in the successor case.
If +1 < j0,+1 ( ) then the game continues and we pass to mega-round + 1 (of the
skipping game).
6A The game 219

Alternatively player I can call for a skip. In this case mega-round continues with
rules additional to the ones of the successor case of G branch . Let W denote the class
W of Section 4A, computed in M . The additional rules are:

(S5) I plays and C subject to the following conditions:

(1) belongs to W ;
(2) is greater than and smaller than j0, ( );
(3) C is a -name in M .

Let C = Back I ( , )(C ), where the pullback is computed inside M . I must


play C in such a way that:

(4) t belongs to an interpretation of k (C ) over Q .

(S6) II plays h , M+1 , and t+1 so that:

(1) M+1 is a wellfounded model of ZFC , h : Q M+1 (displayed


in Diagram 6.3) is elementary, crit(h ) > k ( ), and h is preserves
countability;
(2) t+1 is an (h k )( )-sequence over M+1 ;
(3) t+1 extends t ; and
(4) t+1 belongs to an interpretation of (h k )(C ) over M+1 .

These two rules should be viewed as an allow and choose arrangement. Player I
through her choice of and C determines which extensions of t are allowedthese
are the ones which belong to interpretations of shifts of C and player II chooses
among them.

kkk '
/ M kSk
SkSSSS k / Q M+1 /

T

Diagram 6.3. Successor case with a skip.

Rules (S5) and (S6) complete mega-round in the case of a skip. We set j,+1 =
h k , set +1 = j,+1 ( ) and pass to mega-round + 1 using the objects M+1
and t+1 chosen by player II subject to rule (S6). # (Skip case)
branch (M, t, )(C ) sets rules for the construction of:
Recall that G
220 6 Along a single branch

a wellfounded model M ;
an elementary embedding j : M M ; and
a j ( )-sequence t over M , extending t.
skip (M, t, )(C ) sets similar rules for the construction of the same objects. The
G
only difference between the two games is that G skip allows player I to skip along the
way to . I can in mega-round choose some , possibly much larger than , and

catapult the game directly to the level of a -sequence without having to actually play
through all the Woodin cardinals between and .

Remark 6A.16. Let j0,+1 be an embedding created through a position in G skip . Then
j0,+1 preserves countability. This follows from the restrictions to extenders which are
countable in V in rules (S3), (L2), and (L5), the restriction to an h which preserves
countability in rule (S6), and the restriction in condition (I4) which implies that is
countable.

Our goal next is to construct winning strategies for I in skipping games. This is
done in Sections 6B through 6E, with the central cases being Section 6C which handles
pullbacks from relative limits, and Section 6D which handles pullbacks from Woodin
limits of Woodin cardinals. In Section 6F we return to G branch and construct winning
strategies for I in that game by sewing together winning strategies in various skipping
games. The end results are listed in Section 6G. Chapter 7 will appeal to these end
results about G skip .
branch , but not to any of the intermediary results on G

6B Successors, basic step


Let M be a model of ZFC . Let be a relative successor in W (computed in M) and let
C M be a -name. Let = e( ) and let t be an annotated M-position of relative
domain . Let w be a witness for over M. We work to isolate the successor case of
branch , and see how I can play in that case.
G
suc (M, t, w, C ) to be the following game:
Define G
players I and II collaborate as usual to produce a real y;
I plays a length iteration tree T on M, with critical points above rdm(t) and
using only extenders which are countable in V; and
II plays a cofinal branch b through T .
This completes the game. We let Q be the direct limit along the branch b of T and let
k : M Q be the direct limit embedding. We let t = t, w, y. This is an annotated
position of relative domain k( ) + 1 over Q. Player I wins the run consisting of y, T ,
and b just in case that one of the following conditions holds:
6C Relative limits 221

(B1) Q is illfounded;

(B2) t is I-acceptably obstructed over Q; or

(B3) Q is wellfounded, t is obstruction free over Q, and there exists some g so that:

(1) g is col(, k() )-generic/Q, and


(2) t k(C )[g].

Remark 6B.1. It t belongs to k(C )[g] then it has to be obstruction free over Q.
(Even more than that: t has to be a k( )-sequence over Q and therefore Q-clear.) So
the statement that t is obstruction free in condition (B3) is redundant. We include it to
emphasize the contrast with condition (B2).
suc (M, t, w, C ) is closely related to the successor case of G
G branch (. . . ). The rules

of Gsuc (M, t, w, C ) correspond precisely to rules (S2)(S4) in Section 6A. Moreover
the payoff conditions (B1)(B3) correspond very closely to conditions (I1), (I2), and
(P1) in Section 6A.

Lemma 6B.2. Suppose that:

(1) M& + 1 is countable in V;

(2) t and w exist in a generic extension of M by a forcing of size or less (in M if


= 0); and

(3) suc (t, w, C ) (defined in Section 4C(2)) holds in that generic extension.
suc (M, t, w, C ).
Then I has a winning strategy in G

Proof. This is a straightforward application of the techniques of Chapter 1. The main


point is the correspondence between the payoff conditions (B2) and (B3), and the
definition of A in Section 4C(2). We leave the precise details to the reader. Let us only
note that the assumption on the size of the forcing in condition (2) is used in conjunction
with Remark 4C.9 and Lemma 1C.7, to see that the relevant iteration embeddings extend
to act on the generic extension which contains {t, w}. #

6C Relative limits
Let M be a transitive model of ZFC . Let be a relative limit in W (where W is
computed in M). Let Y be a -name in M. We work with these fixed objects through-
out Section 6C. Later in the section we will work with some t which belongs to an
interpretation of a pullback of Y , and aim to construct a winning strategy for player I
skip (M, t, )(Y ). But first we work to establish some auxiliary claims.
in G
222 6 Along a single branch

We follow the notation of Section 4C(3). We use C to denote the function which
assigns, to each ordinal < jumpM ( ), each P (M& )< , each W , and
each L, the name C(, , P , ) given by Definition 4C.14. The function C
belongs to M. It can be shifted via elementary embeddings which act on M.
We use R to denote the function which assigns to each ordinal < jumpM ( ) the
name R[ ] of Definition 4C.18. This function too belongs to M, and can be shifted via
elementary embeddings which act on M. Let = e( ). Recall that R[ ] is a name in
col(, ) for a subset of (M& ) (M&) .

Remark 6C.1. The restriction to < jumpM ( ) is in line with the definitions of
Section 1F. Its point is to make C and R set functions which literally belong to M,
rather than class functions definable over M. It is easy to check that all the ordinals
which come up in the constructions later in the section are smaller than the appropriate
shift of jumpM ( ). (Note especially that L in Definition 6C.3 is smaller than jumpM ( ),
by Claim 1A.16.) So the restriction to < jumpM ( ) in the domain of C does not
pose any problem.

Let Rmix denote the mixed pivot games map associated to R, , and X = M&.
This is a Lipschitz continuous map which assigns to each x (M&) the mixed pivot
game Rmix [x]. Round n of this game is displayed in Diagram 6.4, which we copy from
Chapter 1. Precise details can be found in Section 1F.

I f (n), T f (n) + 1 n ln , pn , un
II Ef (n) , Ef (n)+1 , wn

Diagram 6.4. Round n of Rmix [x].

We use the letter P to range over both finite positions in Rmix [x] and infinite runs
of Rmix [x]. We say that P has length n if it covers rounds 0 through n 1. We say
that P has length n + 0.2 if in addition it covers the moves f (n) and T f (n) + 1 in
round n. This is illustrated by the vertical dotted line in Diagram 6.4. A position of
length n + 0.2 covers moves precisely up to this dotted line. We refer to the part of
round n left of the dotted line as the first fifth of round n.

6C (1) Finite. Let x be some element of (M&) . Let n be a natural number. Let P
be a position of length n + 0.2 in Rmix [x]. P gives rise to:

a list of ordinals 0 , . . . , n1 ;

an increasing list of natural numbers f (0), . . . , f (n);

an iteration tree T f (n) + 1 on M, with a final model Mf (n) and an embedding


j0,f (n) : M Mf (n) ; and

a sequence P = $a0 , . . . , an1 % in (Mf (n) &j0,f (n) ( ))n .


6C Relative limits 223

We follow the notation of Section 1F(1) when discussing these objects. We work to
develop some terminology concerning P, leading to Claim 6C.7 which we will need
later on.
Claim 6C.2. The critical point of j0,f (n) is greater than .
Proof. The rules of the mixed pivot game, stated in Section 1F (1), force all extenders
used in j0,f (n) to have critical points above rank(X). X in our case is M&, see
Remark 4C.12. So all the extenders used in T f (n) + 1 must have critical points above
rank(M&), and the claim follows. #
Definition 6C.3. If n = 0 define (P) to be j0,f (0) (L ) where $L , H % is the least
pair of local indiscernibles of M relative to . If n > 0 define (P) to be equal to
jf (m),f (n) (m ) where m < n is largest so that f (m) belongs to [0, f (n))T .
If n > 0 then f (0) belongs to [0, f (n))T . This matter is discussed in Section 1F (1).
The reference to the largest m < n so that f (m) belongs to [0, f (n))T in Definition 6C.3
therefore makes sense.
Definition 6C.4. Let belong to W , and let belong to L. Define K(, , P)
to be equal to j0,f (n) (C)(, , P , (P)).
P in Definition 6C.4 is the sequence $a0 , . . . , an1 % given by P. The name
j0,f (n) (C)(. . . ) is the one given by Definition 4C.14 shifted to Mf (n) . We saw above
that the critical point of j0,f (n) is larger than . So W and L in the sense of M are the
same as W and L in the sense of Mf (n) , up to . The reference to j0,f (n) (C)(, , . . . ) in
Definition 6C.4 therefore makes sense. j0,f (n) (C)(, , . . . ) is a -name in the sense of
Mf (n) . Since the critical point of j0,f (n) is greater than which in turn is greater than ,
j0,f (n) (C)(, , . . . ) is also a -name in the sense of M. It follows that K(, , P) is a
-name in M.
Remark 6C.5. If n = 0 and f (0) = 0 then K(, , P) is simply equal to C(, , , L ).
Note that the conditions n = 0 and f (0) = 0 determine P completely: T f (0) + 1
must be the trivial tree on M with first and last model equal to M, and there are no
subsequent moves in P. We refer to the unique P determined by the conditions n = 0
and f (0) = 0 as the trivial position of length 0.2.
Fix W and L for the rest of Section 6C (1). Let g be col(, )-
generic/M. Fix a -sequence t over M. Suppose that:
(C1) t belongs to K(, , P)[g].
Claim 6C.6 (under assumption (C1) above). n is not used in t ().
Proof. Recall that n is equal to the length of the sequence P given as part of the position
P. Assumption (C1) says that t belongs to an interpretation of j0,f (n) (C)(, , P , (P)).
The fact that n is not used in t () now follows directly from condition (2) in the definition
of C, Definition 4C.14. #
224 6 Along a single branch

Let w = t () + $n, %. This is a witness for + 1 = rdm(t). Let x = t, w.


Suppose that:

(C2) x agrees with x to n.

Recall that P was fixed a position of length n + 0.2 in Rmix [x]. Since x and x agree
to n we may regard P also as a position in Rmix [x]. Let mix be the mixed pivot
strategies map associated to R, , and X = M&. (See Section 1F (2) for details
regarding this map.) Fix a surjection  : M& + 1. Suppose that:

(C3) P is played according to mix [, x].

Suppose finally that:

(C4) P is useful (see Definition 1F.8).

Let denote + 1. Let denote the first Woodin cardinal of M above .

Claim 6C.7 (under assumptions (C1)(C4) above). There exist n , , and P so that:

(1) n is a natural number greater than n;

(2) is an element of W , greater than ;

(3) P is a position of length n + 0.2 in Rmix [x], extending P;

(4) P is played according to mix [, x];

(5) P is useful; and

(6) suc (t, w, C ) holds true with the assignment C = Back I ( , )(C ), where
C = K( , , P ).

Proof. We use Glim (. . . ) to refer to the games defined in Section 4C(3), and played ac-
cording to Diagram 4.2. The assumption that t belongs to an interpretation of K(, , P)
tells us that I wins the game j0,f (n) (Glim )(t, , , P , (P)). Let be a winning strat-
egy for I in this game. Since the game is clopen and belongs to Mf (n) [g] we may
fix inside Mf (n) [g].
and mix [, x] together give rise to n , , and P :

plays and n ;

mix [, x] plays the moves corresponding to II in P ;

we play f (m) = f (m 1) + 2 for I in the first fifth of each round m > n in P ,


so that T f (m) + 1 is simply the tree constructed up to round m, unextended;
and

and its shifts to the models Mf (m) play Is moves outside the first fifth in each
round m n of P .
6C Relative limits 225

We leave it to the reader to verify that the objects , n , and P constructed through
this use of and mix [, x] satisfy the conditions of the claim. Let us just note that:
condition (6) follows from the shift of the payoff condition (P) in Section 4C (3) to the
model Mf (n ) ; and condition (5) follows from the fact that P is useful, the definition of
(P), and the demands in Section 4C(3) which force the ordinals i to descend. #

6C (2) Infinite. Let t be an annotated M-position of relative domain or possibly


less. Let w be a witness for rdm(t). Set x = t, w. Let  be a surjection of onto
M& + 1. Let P be an infinite run of Rmix [x]. Suppose that:

(D1) P is useful and played according to mix [, x].

P gives rise to:

a sequence of ordinals  = $n | n < %;

an infinite increasing list of natural numbers f = $f (n) | n < %;

a length iteration tree T on M, with models Mn and embeddings jm,n : Mm


Mn for m < n < .

Let b be a cofinal branch through T . Let Q be the direct limit along b and let jk,b : Mk
Q for k b be the direct limit embeddings. The run P also gives rise to:

a sequence a = $an | n < % in (Q&j0,b ( )) .

We follow the notation of Section 1F(1) when discussing the objects  , f , T , and a
given by P.
Suppose that:

(D2) Q is wellfounded.

We work under the assumptions (D1) and (D2) for the rest of Section 6C (2).

Claim 6C.8. b is an odd branch.

Proof. Recall that a cofinal branch of T is even if it contains arbitrarily large nodes in
{f (n) | n < }. Otherwise the branch is odd. The fact that P is useful implies that all
cofinal even branches of T lead to illfounded direct limits, see Definition 1F.8. Since
Q is wellfounded b must be odd. #

Claim 6C.9. The embeddings jk,b have critical points greater than .

Proof. The critical point of jk,b is greater than rank(X) by Claim 1F.12, and X in our
context is equal to M& by Remark 4C.12. #

Let I = {$, % ( L) ( W ) | < }. Recall that (P, b) denotes


jf (m),b (m ) where m < is largest so that f (m) b, see Definition 1F.6.
226 6 Along a single branch

Definition 6C.10. For each n < let uP,b (n) : I Q& be the function defined by
uP,b (n)(, ) = j0,b (C)(, , a n, (P, b)). Let uP,b denote the sequence $uP,b (n) |
n < %.

uP,b is simply the sequence ua, given by the shift of Definition 4C.17 to Q, applied
with = (P, b).
Recall that t, fixed as the outset of Section 6C (2), is assumed to have relative domain
or possibly less. The next claim shows among other things that t must in fact have
relative domain equal to . Like all claims here it is made under assumptions (D1)
and (D2).

Claim 6C.11. The relative domain of t is precisely equal to . Moreover there exists
some v (Q& + 1) so that one of the following conditions holds with the settings
y = uP,b v and s = t, w, y:

(1) s is I-acceptably obstructed over Q; or

(2) s is obstruction free over Q and there exists some h so that:

(a) h is col(, j0,b ( ))-generic/Q, and


(b) s j0,b (Y )[h].

Proof. The proof of Claim 6C.11 is a straightforward application of the methods of


Section 1F. The key point is the correspondence between the conclusions of the claim
and the definition of R in Section 4C(3). Let us go over this quickly.
Lemma 1F.11 tells us that P is a mixed R-pivot for x. Since b is an odd branch
of T it follows that there exists some h so that:

(i) h is col(, j0,b ( ))-generic/Q; and

a , x% j0,b (R)[b ][h], where b stands for (P, b).


(ii) $

Combining the second condition with the shift of Definition 4C.18 to Q we see that:

(iii) x is a -code; and

(iv) there exists some v (Q& + 1) so that x  (uP,b v) is either I-acceptably


obstructed over Q, or an element of j0,b (Y )[h].

The conclusion of the current claim follows immediately from conditions (iii) and (iv).
It is in passing from condition (ii) to conditions (iii) and (iv) that we make use of the
definition of R in Section 4C(3). #

Fix some n < . We work with this fixed n for the rest of Section 6C (2).
Let m < be largest so that f (m) belongs to b. (P, b) by definition is equal to
jf (m),b (m ). Recall that e(n), in the notation of Section 1F, is equal to 0 if n = 0 and
to f (n 1) + 2 if n > 0.
6C Relative limits 227

Let k be the first element of the odd branch b which is larger than e(n), and larger
than f (m). Let P be the position of length n + 0.2 in Rmix [x] which follows P for
rounds 0 through n1, and contains the moves f (n) = k and T f (n)+1 = T k+1
for the first fifth of round n. We use f (i), i , Mi , etc. when discussing the objects
which form P . Diagram 6.5 displays the connection between P and round n of P . We
leave it to the reader to check that the moves f (n) = k and T f (n) + 1 = T k + 1
satisfy the requirements of rule (1) in Section 1F (1).

II
#
P Me(n) _ _I _ Mf (n) Mf (n)+1 Mf (n)+2 _ I _ Mf (n+1) Mk

P Me(n) _ _ _ _ _ _ _ _ _ _I _ _ _ _ _ _ _ _ _ _ Mk
(Me (n) ) (Mf (n) )

Diagram 6.5. Round n (the first fifth) in P .

Definition 6C.12. We refer to the position P defined above as trunc(P, b, n).


Let c = uP,b (n). c is a function which assigns a col(, )-name c(, ) to each
pair $, % I .
Claim 6C.13. Let $, % belong to I . Let g be col(, )-generic/M. Then c(, )[g]
K(, , P )[g].
Proof. c(, ) is by definition equal to j0,b (C)(, , a n, jf (m),b (m )). Pulling this
equality back to Mk via jk,b we see that:
(i) c(, ) = j0,k (C)(, , a n, jf (m),k (m )).
Note that a n is not affected by jk,b , because of Claim 1F.13 and because k is larger
than e(n). , , and c(, ) are also not affected by jk,b , since they lie below .
Let m < n be largest so that f (m) belongs to [0, k)T . Then:
(ii) K(, , P ) = j0,k (C)(, , a n, jf (m),k (m )).
This can be seen directly by applying the definition of K to P .
Both m and m belong to [0, k)T . m is largest so that f (m) belongs to [0, k)T . m
on the other hand is just the largest number below n so that f (m) belongs to [0, k)T .
So m m. Using the fact that P is useful (see Definition 1F.8) it follows from this that
m jf (m),f (m) (m ) (with strict inequality iff m > m strictly). Using jf (m),k to shift
the inequality to Mk we get:
(iii) jf (m),k (m ) jf (m),k (m ).
The dependence of j0,k (C)(, , a n, )[g] on is monotone increasing by Re-
mark 4C.15. It follows from this and from conditions (i)(iii) that c(, )[g]
K(, , P )[g]. #
228 6 Along a single branch

6C (3) Construction. We can now phrase and prove the main result of Section 6C.
We demonstrate the existence of winning strategies for player I in G skip (M, t, )(Y ),
for annotated positions t which belong to interpretations of pullbacks of Y .
Lemma 6C.14 (for a relative limit and a -name Y in M). Let < be an element
of W . Let t be a -sequence over M.
Suppose that M& +1 is countable in V. Suppose that t belongs to an interpretation
of Back I (, )(Y ), where the pullback is computed in M. Then player I has a winning
skip (M, t, )(Y ).
strategy in G
Proof. Let g witness that t belongs to an interpretation of Back I (, )(Y ). g is col(, )-
generic/M, and t belongs to Back I (, )(Y )[g]. Let $L , H % be the least pair of local
indiscernibles of M relative to . Definition 4C.20 tells us that Back I (, )(Y ) is equal
to C(0, , , L ). So:
(i) t C(0, , , L )[g].
Remember that we aim to demonstrate that player I has a winning strategy in
skip (M, t, )(Y ). Fix an imaginary opponent willing to play for II in this game.
G
We describe how to play for I, and win. We work in mega-rounds subject to the rules
in Section 6A. At the start of mega-round for a successor or zero we will have the
following objects:
(A) a wellfounded model M ;
(B) an elementary embedding j0, ;
(C) W and a -sequence t over M ;
(D) a generic g for col(, ) over M ;
(E) L and n < ;
(F) a map : a M &j0, ( ) + 1 where a is the set of numbers used
in the witness t ( ); and
(G) a position P of length n + 0.2 in j0, (Rmix )[x ], where x equals
t  , t ( ).
W and L above are the classes W and L of Section 4A, computed in M .
Fix for the entirety of the proof a bijection r : , with the following
property:
(ii) For each n < , r  n n .
Using and the bijection r define a map  : M &j0, ( ) + 1 by:

(r(i)) if r(i) a ; and
 (i) =
otherwise.
6C Relative limits 229

We will make sure that the objects listed in conditions (A)(G) satisfy the following
conditions at the start of mega-round , for a successor or zero:

(1) P is played according to j0, (mix )[x ,  ].

(2) P is useful.

(3) t belongs to j0, (K)( , , P )[g ].

(4) t ( ) is precisely equal to {$n , % | < and is a successor or zero}.

(5) extends j, for each < .

(6) Let < be a successor or zero. Suppose that n n for each successor
(, ]. Then P extends j, (P ).

Only the first three conditions will actually be used in the successor (or zero) case of
the construction. The remaining three conditions are maintained as preparation for the
limit case.
Set to begin with: M0 = M, 0 = , t0 = t, g0 = g, 0 = 0, and n0 = 0.
Let 0 : M& + 1 be the empty function. Let P0 be the trivial position of
length 0.2, see Remark 6C.5. Conditions (1) and (2) for = 0 hold trivially with
these assignments. Condition (3) follows from condition (i) above, and Remark 6C.5.
Condition (4) holds trivially, since is the only witness for 0 = 0. Conditions (5) and
(6) are vacuous for = 0.
Let us now describe mega-round . We divide into three main cases: successor (or
zero); limit with an early end; and limit with a leap.

Successor (or zero). We start with the case that is a successor or zero. Using condi-
tion (3) and Claim 6C.6 we see that:

(iii) n is not used in t ( ). (In other words n  a .)

Let w = t ( )+$n , %. This is a witness for +1 = rdm(t ). Let x = t , w .


Using Claim 4A.16 we see that:

(iv) x and x = t  , t ( ) agree to n .

Note that j0, preserves countability. This is trivial if = 0, and follows from
Remark 6A.16 if is a successor. We assume in Lemma 6C.14 that M& + 1 is
countable in V. Combining this with the fact that j0, preserves countability it follows
that M &j0, ( )+1 is countable in V. Fix a surjection : {n } M &j0, ( )+1.
Define  : M &j0, ( ) + 1 by:


(r(i)) if r(i) a ;
 (i) = (r(i)) if r(i) {n } ; and


otherwise.
230 6 Along a single branch

There are no conflicts between the first two clauses in the definition of  , since n  a .
 : M&j0, ( ) + 1 is surjective, since its range contains the range of . Using
condition (ii) above it is easy to check that  and  agree to n . We already saw that
x agrees with x to n . Combining these agreements with condition (1) we get:
(v) P is played according to j0, (mix )[ , x ].
Conditions (3), (iv), (v), and (2) allow for an application of Claim 6C.7 over M . Let
n , , and P be the objects given by that claim. These objects satisfy the conditions
listed in Claim 6C.7. In particular:
(vi) is greater than the first Woodin cardinal of M above , and smaller than
j0, ().
Let = + 1 and let be the first Woodin cardinal of M above . Let C denote
j0, (K)( , , P ). Let C = Back I ( , )(C ). Condition (6) of Claim 6C.7
tells us that suc (t , w , C ) holds true. suc (t , w , C ) is absolute between generic
extensions of M which have t and w as elements. So we dont have to specify exactly
in which extension it holds. We will use the fact that it holds in M [g ].
The fact that suc (t , w , C ) holds in M [g ] puts us in a position to apply Lemma
6B.2. The lemma states that player I has a winning strategy in G suc (M , t , w , C ).

Let  be such a strategy.
Remember that we are working with an imaginary opponent to construct a run of
skip (M, t, )(Y ). We are currently working on mega-round , where is a successor
G
or zero. The mega-round is played according to rules (S1)(S6) in Section 6A.
The assignment of w above covers the move corresponding to rule (S1).
Let  (playing for I) and the imaginary opponent (playing for II) cover the moves
corresponding to rules (S2)(S4). They produce a real y , an iteration tree T , and a
cofinal branch b through T .
Following the notation of Section 6A let Q be the direct limit along the branch b
of T . Let k : M Q be the direct limit embedding. Let t = t , w , y .
We constructed Q , k , and t using the strategy  , which is winning for I in
suc (M , t , w , C ). Copying from the payoff conditions (B1)(B3) in Section 6B
G
we see that at least one of the following conditions must hold:
Q is illfounded;

t is I-acceptably obstructed over Q ; or

Q is wellfounded, t is obstruction free over Q , and t belongs to an interpre-


tation of k (C ) over Q .
skip (M, t, )(Y ) ends and player I wins through condition (I1)
If Q is illfounded then G

in Section 6A. If t is I-acceptably obstructed over Q then again G skip (M, t, )(Y )
ends, and player I wins through condition (I2) in Section 6A. So suppose that:
6C Relative limits 231

(vii) Q is wellfounded, t is obstruction free over Q , and t belongs to an interpre-


tation of k (C ) over Q .
So far we constructed moves corresponding to rules (S1)(S4) in Section 6A. We
continue mega-round with a skip subject to rules (S5) and (S6).
The assignments of and C above cover the moves described in rule (S5). We
leave it to the reader to verify that these assignments satisfy the conditions of the rule. Let
us just note that condition (4) in rule (S5) follows from the final clause in condition (vii)
above.
Let the imaginary opponent play the moves h , M+1 , and t+1 corresponding to
rule (S6). Let j,+1 = h k . The demands placed on the imaginary opponent
through rule (S6) are such that:

(viii) t+1 extends t ; and

(ix) t+1 belongs to an interpretation of j,+1 (C ).

Let +1 = + 1. It follows from condition (viii) and from the definition of t


that t+1 (+1 ) equals w . Combining this with the definition of w we get:
(x) t+1 (+1 ) equals t ( ) + $n , %.
It follows from this that a+1 , the set of numbers used in t+1 (+1 ), is precisely equal
to a {n }. Define +1 : a+1 M+1 &j0,+1 ( ) + 1 by setting:
(xi) +1 = (j,+1 ) (j,+1 ).
The union is a function since the domains of and are disjoint. The former has
domain a , the latter has domain {n } , and n  a by condition (iii).
Let n+1 = n , let P+1 = j,+1 (P ), and let +1 = j,+1 ( ). (Recall that
n , P , and are the objects obtained above through an application of Claim 6C.7.)

Combining condition (ix) with the definition of C we see that t+1 belongs to an
interpretation of j0,+1 (K)(+1 , +1 , P+1 ). Let g+1 be a generic object which
witnesses this.
We have now defined all the objects listed in conditions (A)(G) for + 1. It is easy
to check that conditions (1)(6) hold for + 1 with the definitions we made. Let us
only make the following comments: Condition (1) for + 1 follows from the properties
of P obtained through the application of Claim 6C.7, specifically the fact that P is
played according to j0, (mix )[ , x ]. Condition (6) for + 1 follows from the same
condition for and the fact that P extends P . # (Successor case)
Recall that we are working with the imaginary opponent to construct a run of
skip (M, t, )(Y ). So far we described the construction for mega-round in the case
G
that is a successor or zero. Let us now handle limit mega-rounds. Let be a limit ordi-
nal. Suppose the construction reached mega-round . Suppose that conditions (1)(6)
hold for = 0 and for all successor < . We describe how to play in mega-round .
232 6 Along a single branch

Let M be the direct limit of the system $M , j, | < %. Let j, be the


direct limit embeddings. If M is illfounded then G skip (M, t, )(Y ) ends and player I
wins through condition (I3) in Section 6A. So suppose that M is wellfounded.
Let t = < t . This is an annotated M -position. Let S be the set consisting
of zero and all successor ordinals smaller than . The relative domain of t is equal to
sup{ + 1 | S }. From condition (vi) it follows that:

(a) rdm(t ) j0, ().

Using conditions (4) and (iii) it is easy to see that the numbers n for S are
distinct. Since is a limit it follows that n as . For each n < let
n < be a successor ordinal large enough that:

(b) n n for all successor [n , ).

Let w = {$n , % | S }. This is a witness for rdm(t ). Let x = t , w .


Using condition (4) we see that w extends tn (n ) for each n < . Adding
condition (b) we see that w is in fact an n-extension of tn (n ). By Claim 4A.18 it
follows that:

(c) xn and x agree to n.



Let = S (j, ). The union makes sense and yields a function by
condition (5). Let a = {n | S }. This is the set of numbers used in w . Define
 : M &j0, ( ) + 1 by:

(r(i)) if r(i) a ; and
 (i) =
otherwise.

Using our definitions, condition (b), and condition (ii) it is easy to see that:

(d) jn ,  n and  agree to n.

By condition (xi), the range of contains the range of j, for each successor
< . Recall that was picked a surjection on M &j0, ( ) + 1. So the range of
contains the image under j, of M &j, ( ) + 1 for each successor < . It follows
that : a M &j0, ( ) + 1 is a surjection. This in turn implies that:

(e)  : M &j0, ( ) + 1 is a surjection.

Given condition (c) we may regard jn , (Pn ) as a position in j0, (Rmix )[x ]. The
position is useful and played according to j0, (mix )[ , x ]. Using conditions (6)
and (b)it is easy to see that jn+1 , (Pn+1 ) extends jn , (Pn ) for each n < . Set
P = n< jn , (Pn ). Then:

(f) P is an infinite run of j0, (Rmix )[x ]. P is useful and played according to
j0, (mix )[ , x ].
6C Relative limits 233

Let T be the iteration tree given by P . P and T will be the driving forces behind
our moves for player I in mega-round .
Remark 6C.15. The reader should compare the limit settings here with the limit con-
struction in Section 3B(3). In both constructions we use the agreement secured in
mega-rounds prior to to argue for the convergence which produces the mixed pivot
needed at . In both constructions the argument hinges on the fact that the numbers n
are distinct, and therefore converge to .
We already saw that the numbers n for S are distinct. It follows from this that
skip (M, t, )(Y ) does not end with a loss for I through condition (I4)
is countable. So G
in Section 6A. We now describe how to play in mega-round . We divide into three
cases, but one of them is degenerate.

Degenerate limit. If all cofinal branches through T lead to illfounded direct limits.
skip (M, t, )(Y ) by playing T at the first opportunity.
In this case player I can win G
The opportunity comes already at mega-round if falls under the conditions of the
standard limit case in Section 6A. Otherwise the opportunity comes at mega-round
+ 1. Once T is played, II is forced into a loss through one of the conditions (I1) and
(I3) in Section 6A. # (Degenerate limit)

Suppose now that some cofinal branch through T leads to a wellfounded direct
limit.
Claim 6C.16. rdm(t ) is equal to j0, ().

Proof. Immediate from Claim 6C.11, applied over M . To apply the claim we must
work with a cofinal wellfounded branch through T . Which particular wellfounded
branch we take does not matter here. #

It follows from the last claim that rdm(t ) is a limit of Woodin cardinals in M , but
not itself a Woodin cardinal. Mega-round therefore falls under the standard limit case
in Section 6A, played subject to rules (L1)(L6).
The assignments of w and T above cover the moves corresponding to rules (L1)
and (L2). Let the imaginary opponent play b subject to rule (L3).
Let Q be the direct limit along the branch b of T . Let k : M Q be the
direct limit embedding. If Q is illfounded then G skip (M, t, )(Y ) ends and player I
wins through condition (I5). So suppose that Q is wellfounded.
Let u = uP ,b , where uP ,b is given by Definition 6C.10 applied over M . Let
v be given by Claim 6C.11, applied with P and b over M . Let y = u v .
This last assignment covers the move corresponding to rule (L4) in Section 6A. Set
s = t , w , y . Translating the conclusion of Claim 6C.11 to the current context
we get:

(g) s is either I-acceptably obstructed over Q ; or else it is obstruction free and


belongs to an interpretation of (k j0, )(Y ).
234 6 Along a single branch

skip (M, t, )(Y ) does not end with a loss for I through
In particular it follows that G
condition (I6) in Section 6A.
Our work so far covers the moves corresponding to rules (L1)(L4) in Section 6A.
The imaginary opponent may now elect an early end to mega-round , or elect to
continue with a leap subject to rules (L5) and (L6).

Early end. If the imaginary opponent elects an early end. Following the notation of
Section 6A set M+1 = Q , j,+1 = k , and t+1 = s .
skip do not allow
s must be obstruction free over Q , since otherwise the rules of G
II to elect an early end in mega-round . Using condition (g) it follows that s belongs
to an interpretation of (k j0, )(Y ). In other words:

t+1 belongs to an interpretation of j0,+1 (Y ).

This means that G skip (M, t, )(Y ) ends, and player I wins through the payoff condi-
tion (P1) in Section 6A. # (Early end)

Leap. If the imaginary opponent elects a leap in mega-round . Let E , +1 , and t+1
be the moves made by the imaginary opponent subject to rules (L5) and (L6).
Following the notation of Section 6A let M+1 = Ult(M , E ) and let j,+1 be the
ultrapower embedding. If M+1 is illfounded then G skip (M, t, )(Y ) ends and player I
wins through condition (I7) in Section 6A. So suppose that M+1 is wellfounded.
The moves +1 and t+1 played by the imaginary opponent must satisfy the con-
ditions of rule (L6) in Section 6A. Let n witness that they satisfy condition (3) in that
rule. Let c = u (n ) and let c+1 = j,+1 (c ). Let = j0, (). Condition (3)
in rule (L6) tells us that t+1 belongs to an interpretation of c+1 ( , +1 ). Let g+1
witness this. Then:

(xii) g+1 is col(, +1 )-generic/M+1 and t+1 c+1 ( , +1 )[g+1 ].

Let P be the position trunc(P , b , n ) given by Definition 6C.12 applied over


M . Using Claim 6C.13 over M we see that:

(xiii) For every pair $, % j0, (I ), and every g which is col(, )-generic/M ,
c (, )[g] is contained in j0, (K)(, , P )[g].

Let P+1 = j,+1 (P ). Shifting the last condition to M+1 and applying it with
the pair $ , +1 % we see that:

(xiv) c+1 ( , +1 )[g+1 ] j0,+1 (K)( , +1 , P+1 )[g+1 ].

Combining condition (xiv) and the final clause in condition (xii) we get:

(xv) t+1 belongs to j0,+1 (K)( , +1 , P+1 )[g+1 ].

Let +1 = and let n+1 = n . Let +1 equal j,+1 . We have defined


all the objects of conditions (A)(G), listed above in the successor case, for + 1. It is
easy to check that conditions (1)(6) hold for + 1 with our definitions. Let us only
6D Woodin limits of Woodin cardinals 235

make the following comments: Condition (3) for + 1 follows from condition (xv)
above. Condition (4) for + 1 follows from the definition of w and the fact that
t+1 ( ) = w .
We may now proceed to mega-round + 1 of the construction. # (Leap)

The various cases above provide a complete inductive description of a construction


of a run of G skip (M, t, )(Y ), joint with an imaginary opponent who plays for II. In
the early end case the construction terminates with a victory for player I through the
payoff condition (P1) of Section 6A. In the other cases the construction may terminate
through one of the conditions (I1)(I7), but it only terminates with victory for player I.
The construction can therefore be formalized into a winning strategy for player I in
skip (M, t, )(Y ).
G # (Lemma 6C.14)

6D Woodin limits of Woodin cardinals


Let M be a transitive model of ZFC . Let be a Woodin limit of Woodin cardinals
in M. Let Y be a -name in M. We work with these fixed objects throughout Section 6D.
Later in the section we will fix some r which belongs to an interpretation of a pullback
skip (M, r, )(Y ).
of Y , and aim to construct a winning strategy for I in G
We follow the notation of Section 4D(1). < is the least ordinal which seals Y .
We use C to denote the function which associates to each [, ) W , each
L, each n , and each ordinal < jumpM ( ), the name C(, , n, ) given by
Definition 4D.5. There is a dependence on the name Y in that definition. But Y is fixed
and we suppress it in the notation. The function C belongs to M. It can (and will) be
shifted via elementary embeddings which act on M.

Remark 6D.1. Again the point of the restriction to < jumpM ( ) is to make C a set
function which literally belongs to M, rather than a class function definable over M.
The ordinals which come up in the constructions later in the section are all smaller
than jumpM ( ), so this restriction does not pose any problems.

6D (1) Hopeful. Fix in addition to the objects listed above:

[, ) W ; and

a -sequence s over M.

Let $L , H % be the least pair of local indiscernibles of M relative to . By Claim 1A.16,


L and H are smaller than jumpM (). So we may talk about C(. . . , L ) and C(. . . , H ).
Suppose that:

(H1) is countable in V;

(H2) s is hopeful with respect to Y over M; and


236 6 Along a single branch

(H3) s belongs to an interpretation of C(0, , 0, L ).

Note that assumption (H2) makes sense, since s is a -sequence and belongs to the
interval [, ).

Lemma 6D.2 (under the assumptions (H1)(H3) listed above). Player I has a winning
skip (M, s, )(Y ).
strategy in G
skip (M, s, )(Y ). We work
Proof. Fix an imaginary opponent willing to play for II in G
with the imaginary opponent to construct a run of mega-rounds in G skip (M, s, )(Y ).
We shall verify after mega-rounds that the run is won by player I. At the start of
mega-round , for < , we will have:

(A) a wellfounded model M ;

(B) an elementary embedding j0, : M M ;

(C) W and a -sequence t over M ;

(D) a generic g for col(, ) over M ;

(E) a function  : j0, ();

(F) L , and n < .

W and L above are the classes W and L of Section 4A, computed in M .


We intend to secure the following inductive conditions:

(1)  is cofinal in j0, ();

(2) t is hopeful with respect to j0, (Y ) over M ; and

(3) t belongs to j0, (C)( , , n , j0, (L ))[g ].

We set to begin with M0 = M, 0 = , t0 = s, 0 = 0, and n0 = 0. Condition (2)


follows from assumption (H2). Assumption (H3) allows picking g0 so as to satisfy
condition (3). Assumption (H1) allows picking 0 so as to satisfy condition (1).
Let us start mega-round of the construction. Using Claim 5D.2 inside M we
may switch from j0, (L ) in condition (3) to the ordinal j0, (H ). We get:

(i) t belongs to j0, (C)( , , n , j0, (H ))[g ].

By Definition 4D.5 this means in particular that:

(ii) n is not used in t ( ); and

(iii) player I wins the game j0, (Gwl )(t , , , n , j0, (H )).
6D Woodin limits of Woodin cardinals 237

Remark 6D.3. The game j0, (Gwl )(t , . . . ) consists of moves , , n , and subject
to the rules listed in Section 4D(1), shifted to M . The precise format of this game is
displayed in Diagram 4.3. Note particularly that the move is the responsibility of
player II. This is because t is hopeful with respect to j0, (Y ) over M , by condition (2).
Keeping the last remark in mind we see that condition (iii) expands to:

(iv) For every < j0, (), there is a and an n satisfying the shift to M of rule (2)
in Section 4D(1), so that for every < j0, (H ), the shift to M of the payoff
condition (P) in Section 4D(1) holds.

Following notation similar to that of Section 4D (1) let w = t ( ) + $n , %, let


= + 1, and let be the first Woodin cardinal of M above .
Let =  (). Let and n be given by condition (iv), applied with = .
These objects satisfy the shift to M of rule (2) in Section 4D (1). In particular this
means that > and < j0, (). Folding into this the value of we have:

(v) >  () and < j0, ().

Let C = j0, (C)( , , n , j0, (L )) and let C = Back I ( , )(C ). Apply-


ing condition (iv) with = j0, (L ) we see that:

(vi) suc (t , w , C ) holds.

(This is just the shift to M of the payoff condition (P) in Section 4D (1).) From
condition (vi) and Lemma 6B.2 it follows that I wins G suc (M , t , w , C ). Let 

be a winning strategy for I in this game.
Remember that we are working with an imaginary opponent to construct a run of
skip (M, s, ). We are currently working on mega-round , which is played according
G
to rules (S1)(S6) in Section 6A.
The assignment of w above covers the move described in rule (S1).
Let  (playing for I) and the imaginary opponent (playing for II) cover the moves
described in rules (S2)(S4). They produce a real y , an iteration tree T , and a cofinal
branch b through T .
Following the notation in Section 6A let Q be the direct limit along the branch b
of T . Let k be the direct limit embedding. Let t = t , w , y .
suc (M , t , w , C ), guarantees that at
The use of  , a winning strategy for I in G
least one of the following conditions must hold:

Q is illfounded;

t is I-acceptably obstructed over Q ; or

Q is wellfounded, t is obstruction free over Q , and t belongs to an interpre-


tation of k (C ) over Q .
238 6 Along a single branch

(These conditions are simply the payoff conditions (B1)(B3) in Section 6B, translated
skip (M, s, )(Y ) through
to the current context.) If Q is illfounded then player I wins G

condition (I1) in Section 6A, and our job is done. If t is I-acceptably obstructed over
Q then player I wins G skip (M, s, )(Y ) through condition (I2) in Section 6A, and
again our job is done. So suppose that:

(vii) Q is wellfounded, t is obstruction free over Q , and t belongs to an interpre-


tation of k (C ) over Q .
We continue mega-round with a skip following rules (S5) and (S6) in Section 6A.
The assignments of and C above cover the moves described in rule (S5). The
reader may easily verify that these assignments satisfy the conditions in that rule. Note
specifically how condition (4) in rule (S5) follows from the last clause in condition (vii)
above.
Let the imaginary opponent cover the moves h , M+1 , and t+1 described in
rule (S6). Let j,+1 = h k . Let +1 = j,+1 ( ). The demands placed on the
imaginary opponent through rule (S6) are such that:

(viii) t+1 extends t ; and

(ix) t+1 belongs to an interpretation of j,+1 (C ).


j0,+1 preserves countability by Remark 6A.16. Using the initial assumption that
is countable in V it follows that j0,+1 () is countable in V, so certainly j0,+1 ( )
has cofinality in V. Pick in V a cofinal map +1 : j0,+1 ( ). By increasing
the values taken by +1 as needed, make sure that it also satisfies:
(x) +1 (i) (j,+1  )(i) for all i < .
The choice of +1 above secures the inductive condition (1) for + 1. The
inductive condition (2) for + 1 follows from the same condition for (but shifted to
M+1 ), the fact that t+1 extends t , and Claim 4D.3 applied in M+1 . In shifting the
hopefulness of t from M to M+1 we use the fact that j,+1 has critical point greater
than rdm(t ). This fact follows from the restrictions on critical points in rules (S3) and
(S6) in Section 6A.
Let +1 = and let n+1 = n . Using condition (ix), pick g+1 so as to satisfy
the inductive condition (3) for + 1.
Our work above completes mega-round of the construction (where is smaller
than ), and puts us in the position necessary to start mega-round + 1. Let us now
fastforward to the start of mega-round .
Let M be the direct limit of the system $M , j, | < %. Let j, : M
M be the direct limit maps. If M is illfounded then player I wins G skip (M, s, )(Y )
through condition (I3) in Section 6A, and our job is done. So suppose that M is
wellfounded.
Translating the limit case condition (e) in Section 6A to our context we get:
6D Woodin limits of Woodin cardinals 239

(xi) crit(j, ) + 1 for each < .



Let t = < t . t is an annotated position over M . Translating the limit case
condition (f) in Section 6A to our context we get:

(xii) t is M -clear.

Claim 6D.4. rdm(t ) = j0, ().

Proof. Condition (v) shifted to M+1 states that +1 > (j,+1  )() and +1 <
j0,+1 ( ) for each < . Applying the embedding j+1, to these inequalities we see
that +1 > (j,  )() and +1 < j0, () for each < . (Note that +1 is not
moved by j+1, , because of condition (xi).)
For each < let  () = (j,  )(). By the argument in the previous
paragraph +1 is trapped between  () and j0, ( ) for each < . Using con-
ditions (1) and (x) it is easy to see that  : j0, ( ) is cofinal in j0, ( ). So
sup< +1 = j0, (). The claim follows. #

Corollary 6D.5. rdm(t ) is Woodin in M .

Proof. This follows from the previous claim, since is Woodin in M. #

Let W M denote the forcing W of Section 4B, defined in M. Let t M be the


name of Definition 4B.39. Let G(t ) be the filter associated to t by Definition 4B.23,
executed over M .
Claim 6D.6. G(t ) is j0, (W)-generic/M and j0, (t)[G(t )] = t .

Proof. Claim 6D.4 and condition (xii) combine to state that t is an M -clear annotated
position of relative domain j0, () over M . The current claim follows through an
application of Corollary 4B.30 and an application of Claim 4B.41. #

Claim 6D.7. t belongs to j0, (Y )[G(t )].

Proof. Recall that one of the initial assumptions in Lemma 6D.2, assumption (H2) to be
exact, states that s is hopeful with respect to Y over M. By Definition 4D.2 this means
that there exists some good identity M& so that s |= and [ ] W t Y .
( here is the least ordinal which seals Y .)
Condition (xi) implies that j0, : M M has critical point greater than 0 . 0
was set equal to at the start of the construction. , which was fixed at the start of
Section 6D (1), is greater than . So , which belongs to M&, is not moved by j0, .
We know that [ ] forces t Y over M. Applying j0, , and using the fact that is
not moved, we see that [ ] j0, (W) j0, (t) j0, (Y ) over M .
t extends t0 , which was set equal to s at the start of the construction. Using the fact
that s |= over M it follows that t |= over M . So [ ] belongs to G(t ). Since [ ]
forces j0, (t) j0, (Y ) over M , it follows that j0, (t)[G(t )] j0, (Y )[G(t )].
But j0, (t)[G(t )] is equal to t by Claim 6D.6. So t j0, (Y )[G(t )]. #
240 6 Along a single branch

Remember that we are now working on mega-round of G skip (M, s, )(Y ). Corol-
lary 6D.5 tells us that this mega-round follows the rules of the phantom limit case in Sec-
tion 6A. Following the notation of the phantom limit case let +1 = = rdm(t ), and
let j,+1 = id. By Claim 6D.4, +1 = j0,+1 ( ). This means that G skip (M, s, )(Y )
ends with the payoff condition (P2) in Section 6A. By Claim 6D.7 player I wins.
Working with an imaginary opponent who plays for II we described how to construct
skip (M, s, )(Y ). The mechanism we described always leads to runs which
a run of G
are won by player I. It can therefore be formalized into a winning strategy for I in
skip (M, s, )(Y ).
G # (Lemma 6D.2)

6D (2) Not hopeful. Recall that we are working with a transitive model M of ZFC , a
Woodin limit of Woodin cardinals in M, and a -name Y over M. Recall that <
is the least ordinal which seals Y . Recall that C M is the function which associates
to each [, ) W , each L, each n < , and each ordinal < jumpM ( ),
the name C(, , n, ) given by Definition 4D.5.
As in Section 6D(1) let us fix:

[, ) W ; and

a -sequence s over M.

Let $L , H % be the least pair of local indiscernibles of M relative to . Suppose that:

(N1) is countable in V;

(N2) s is not hopeful with respect to Y over M; and

(N3) s belongs to an interpretation of C(0, , 0, L ).

These assumptions are similar to the assumptions (H1)(H3) in Section 6D (1), except
that here we assume that s is not hopeful.

Lemma 6D.8 (under the assumptions (N1)(N3) listed above). Player I has a winning
skip (M, s, )(Y ).
strategy in G
skip (M, s, )(Y ). We work
Proof. Fix an imaginary opponent willing to play for II in G
with the imaginary opponent to construct a run of mega-rounds in G skip (M, s, )(Y ).
At the start of mega-round , for < , we will have:

(A) a wellfounded model M ;

(B) an elementary embedding j0, : M M ;

(C) W and a -sequence t over M ;

(D) a generic g for col(, ) over M ;

(E) a function  : j0, ();


6D Woodin limits of Woodin cardinals 241

(F) L , and n < ; and

(G) an ordinal in M .

This list of objects is similar to the one at the start of the proof of Lemma 6D.2. But
note the addition here of the ordinal listed in condition (G).
We intend to secure the following inductive conditions:

(1)  is cofinal in j0, ();

(2) t is not hopeful with respect to j0, (Y ) over M ; and

(3) t belongs to j0, (C)( , , n , )[g ].

This list too is similar to the one in the proof of Lemma 6D.2. But there are a couple
of differences. In condition (2) we state that t is not hopeful. In condition (3) we use
the ordinal of condition (G).
Set to begin with M0 = M, 0 = , t0 = s, 0 = 0, n0 = 0, and 0 = L .
Condition (2) follows from assumption (N2). Assumption (N3) allows picking g0 so
as to satisfy condition (3). Assumption (N1) allows picking 0 so as to satisfy condi-
tion (1).
Let us start mega-round of the construction. By Definition 4D.5, condition (3)
implies that:

(i) n is not used in t ( ); and

(ii) player I wins the game j0, (Gwl )(t , , , n , ).

The game j0, (Gwl )(t , . . . ) consists of moves , , n , and subject to the
rules listed in Section 4D(1), shifted to M . We refer the reader to Diagram 4.3 for the
precise format of the game. Note that the move is the responsibility of player I here.
This is because t is not hopeful with respect to j0, (Y ) over M , by condition (2).
Remark 6D.9. The fact that the move here falls on player I should be contrasted
with the situation in Lemma 6D.2. There was the responsibility of player II, see
Remark 6D.3.
Keeping in mind that is the responsibility of player I in j0, (Gwl )(t , . . . ) we
see that condition (ii) expands to:

(iii) For every < j0, (), there is a and an n satisfying the shift to M of rule (2)
in Section 4D(1), and there is an ordinal < , so that the shift to M of the
payoff condition (P) in Section 4D(1) holds.

Note that is bounded by an existential quantifier here. This is in line with the fact
that is the responsibility of player I in j0, (Gwl )(t , . . . ).
Let w = t ( ) + $n , %, let = + 1, and let be the first Woodin cardinal
of M above .
242 6 Along a single branch

Let =  (). Let , n , and be given by condition (iii), applied with = .


Note that the ordinal is given to us by condition (iii). The constraints placed by that
condition are such that:

(iv) < .

Let C = j0, (C)( , , n , ) and let C = Back I ( , )(C ).


Construct M+1 , j,+1 , +1 , t+1 , +1 , +1 , and n+1 by copying precisely
the construction of these objects in the proof of Lemma 6D.2, only using the names C
and C defined above. We omit the actual copying, but note that among other things
we get:

(v) t+1 belongs to an interpretation of j,+1 (C ).

Let +1 = j,+1 ( ). Since C was set equal to j0, (C)( , , n , ), con-


dition (v) allows picking g+1 so as to satisfy the inductive condition (3) of the current
construction. The inductive condition (2) carries over to +1 since t+1 extends t , and
since crit(j,+1 ) > . The argument is similar to the one in the proof of Lemma 6D.2.
The inductive condition (1) holds for + 1 through the choice of +1 .
Our work above completes mega-round of the construction (where is smaller
than ) and puts us in the position necessary to start mega-round + 1. Suppose now
that we reached mega-round . Let M be the direct limit of the system $M , j, |
< %. By condition (iv) we have:

+1 < j,+1 ( ) for each < .


skip (M, s, )(Y ) ends through condi-
If follows from this that M is illfounded. So G
tion (I3) in Section 6A, and player I wins. # (Lemma 6D.8)

The reader may wish to compare closely the proofs of Lemmas 6D.2 and 6D.8. Note
how here we argue for victory not through the payoff condition (P2) in Section 6A, but
through the snag of illfoundedness in condition (I3). The initial assumption here
that s is not hopeful eliminates any hope for victory through condition (P2). But the
same assumption gives rise to the ordinals which end up witnessing victory through
condition (I3).

6D (3) Combined. We continue to work with a transitive model M of ZFC , a Woodin


limit of Woodin cardinals in M, and a -name Y in M. L is the lower ordinal in
the least pair of local indiscernibles of M relative to . < is the least ordinal
which seals Y . C M is the function which associates to each [, ) W , each
L, each n < , and each ordinal < jumpM ( ), the -name C(, , n, )
given by Definition 4D.5.
For each [, ) W let C denote the name C(0, , 0, L ).
6D Woodin limits of Woodin cardinals 243

Claim 6D.10. Suppose that is countable in V. Let belong to [, ) W . Let s be


a -sequence over M. Suppose that s belongs to an interpretation of C . Then player I
skip (M, s, )(Y ).
has a winning strategy in G
Proof. s is a -sequence and is larger than . So we may ask whether or not s is
hopeful with respect to Y . If s is hopeful with respect to Y then the claim follows
from Lemma 6D.2. If s is not hopeful with respect to Y then the claim follows from
Lemma 6D.8. #
Lemma 6D.11 (for a Woodin limit of Woodin cardinals and a -name Y in M). Let
< be an element of W . Let r be a -sequence over M.
Suppose that is countable in V. Suppose that r belongs to an interpretation of
Back I ( , )(Y ), where the pullback is computed in M. Then player I has a winning
strategy in G skip (M, r, )(Y ).
Proof. Let be the first Woodin cardinal of M above . Remember that C , defined
for each [, ) W , denotes the name C(0, , 0, L ). For each > in [, )
W let A = Back I ( , )(C ). Let q witness that r belongs to an interpretation of
Back I ( , )(Y ). q is col(, )-generic/M, and r belongs to Back I ( , )(Y )[q]. The
fact that r Back I (, )(Y )[q] allows us, directly by Definition 4D.6, to fix some
> in [, ) W , and to fix in M[q] some witness w for + 1, so that:
(i) suc (r, w, A ) holds.
From this and Lemma 6B.2 it follows that player I has a winning strategy in
suc (M, r, w, A ). Let  be a winning strategy for I in this game.
G
Fix now an imaginary opponent willing to play for II in G skip (M, r, )(Y ). We
describe how to play for I, and win.
Let M0 = M and let t0 = t. We start with mega-round 0. We play according to the
rules (S1)(S6) in Section 6A.
Let I play the witness w of condition (i) above for the move w0 described in rule (S1).
Then let I follow  for the moves described in rules (S2)(S4). Player I and the
imaginary opponent together create y0 , T0 , and b0 . Following the notation of Section 6A
let Q0 be the direct limit along the branch b0 of T0 , and let k0 be the direct limit
embedding. Let t0 = r, w, y0 .
The use of  guarantees that at least one of the following conditions holds:
Q0 is illfounded;
t0 is I-acceptably obstructed over Q0 ; or

Q0 is wellfounded, t0 is obstruction free over Q0 , and t0 belongs to an interpre-


tation of k0 (A ) over Q0 .
(These conditions are simply the winning conditions listed in Section 6B, translated
to the current context.) If Q0 is illfounded then G skip (M, r, )(Y ) ends and I wins

through condition (I1) in Section 6A. If t0 is I-acceptably obstructed over Q0 then
skip (M, r, )(Y ) ends and I wins through condition (I2). So suppose that:
G
244 6 Along a single branch

(ii) Q0 is wellfounded, t0 is obstruction free over Q0 , and t0 belongs to an interpre-


tation of k0 (A ) over Q0 .
We continue mega-round 0 with a skip according to rules (S5) and (S6) in Section 6A.
Let I play 0 = and C0 = C for the moves described in rule (S5). It is easy
to check that these moves satisfy the conditions of rule (S5). Let us just note that
condition (4) in rule (S5) follows from the last clause in condition (ii) above.
The imaginary opponent responds with moves h0 , M1 , and t1 subject to the condi-
tions in rule (S6). Let j0,1 = h0 k0 . The conditions of rule (S6) are such that:
(iii) t1 belongs to an interpretation of j0,1 (C ) over M1 .
j0,1 preserves countability by Remark 6A.16. Since is assumed countable in V
it follows that j0,1 () is countable in V. Using this and condition (iii) we may apply
Claim 6D.10 over M1 , with s = t1 . The claim states that player I has a winning strategy
in G skip (M1 , t1 , j0,1 ())(j0,1 (Y )). Continue to play for I by following this strategy.
This is possible since the rules for mega-round in G skip (M1 , t1 , j0,1 ( )) are precisely
the same as the rules for mega-round 1 + in our run of G skip (M, r, ). Following
the strategy given by Claim 6D.10 may lead to victory for I through one of the snags
described in conditions (I1)(I7) in Section 6A, and in this case we are done. Otherwise
it leads to:
(iv) a wellfounded model M ;
(v) an elementary embedding j1, : M1 M ; and
(vi) a j1, (j0,1 ())-sequence t over M , which belongs to an interpretation of
j1, (j0,1 (Y )).
(These conditions are simply the end conditions in the construction for Lemma 6D.2,
translated to apply over M1 .) The position described by conditions (iv)(vi) is a victory
skip (M, r, )(Y ), so again we are done.
for player I in G
Working with an imaginary opponent who plays for II in G skip (M, r, )(Y ) we
described how to play for I, and win. The description can be formalized to give a
winning strategy for I in G skip (M, r, )(Y ). #

6E Relative successors and compositions


Let M be a transitive model of ZFC . Let be a relative successor in W (computed
in M). Let be the largest Woodin cardinal of M below . We work with these fixed
objects throughout Section 6E.
Recall that the pullback operation is defined in five cases, listed in Section 4D (3).
Our work so far handled situations which correspond to cases (I) and (II). We handle
here the situations which correspond to cases (III)(V).
6E Relative successors and compositions 245

Lemma 6E.1. Let t be a -sequence over M. Let C be a -name in M.


Suppose that M& +1 is countable in V. Suppose that t belongs to an interpretation
of Back I (, )(C ), where the pullback is computed in M. Then player I has a winning
skip (M, t, )(C ).
strategy in G
Proof. We divide the proof into two cases, depending on whether is a Woodin limit
of Woodin cardinals or an element of W . In both cases we follow the notation of
Section 4D (2).
Case 1. Suppose first that is an element of W . Using the assumption that t belongs
to an interpretation of Back I (, )(C ) fix some g so that:
(i) g is col(, )-generic/M; and
(ii) t belongs to Back I (, )(C )[g].
Using the last condition and looking at the definitions in Section 4D (2) we see that:
(iii) In M[g] there exists a witness w for + 1 so that suc (t, w, C ) holds.
Fix such a witness w. Since M& +1 is countable in V we may appeal to Lemma 6B.2.
Applying the lemma to the objects of condition (iii) we get:
suc (M, t, w, C ).
(iv) Player I has a winning strategy in G
It is easy from this to see that I has a winning strategy in G skip (M, t, )(C ). The
game ends after mega-round 0, and I can win by playing w for rule (S1), following her
winning strategy in G suc (M, t, w, C ) for rules (S2)(S4), and declaring an early end
so as to avoid rules (S5) and (S6). # (Case 1)
Case 2. Suppose next that is a Woodin limit of Woodin cardinals in M. In this case
Back I (, )(C ) is a name in the forcing W . Using the assumption that t belongs to
an interpretation of Back I (, )(C ) fix some G so that:
(i) G is W -generic/M; and
(ii) t belongs to Back I (, )(C )[G].
Using the last condition and looking at the definitions in Section 4D (2) we see that:
(iii) The statement there exists some witness w for so that suc (t, w, C ) holds is
forced by some condition in col(, ) over M[G].
Were assuming in Lemma 6E.1 that M& + 1 is countable in V. Since < this
assumption certainly allows finding in V generics for col(, ) over M[G]. Using
condition (iii) fix some g and some w so that:
(iv) g is col(, )-generic/M[G];
(v) w M[G][g] is a witness for ; and
246 6 Along a single branch

(vi) suc (t, w, C ) holds true (in M[G][g]).

Applying Lemma 6B.2 we see that:


suc (M, t, w, C ).
(vii) Player I has a winning strategy in G

As in case 1 it is now easy to check that player I has a winning strategy in


skip (M, t, )(C ).
G # (Case 2)

The two cases above complete the proof of Lemma 6E.1. #

Lemma 6E.2 (for a relative successor W ). Let < be an element of W . Let


r be a -sequence over M. Let C be a -name in M.
Suppose that M& +1 is countable in V. Suppose that r belongs to an interpretation
of Back I ( , )(C ), where the pullback is computed in M. Then player I has a winning
skip (M, r, )(C ).
strategy in G

Proof. We work by induction on . (The use of the inductive assumption is made in


Claim 6E.3 below.)
The assumption that is an element of W smaller than implies that is smaller
than or equal to . If is equal to then the current lemma follows from Lemma 6E.1.
So suppose that < . Let C = Back I (, )(C ). The definition of the pullback in
case (IV) in Section 4D(3) is such that Back I ( , )(C ) = Back I ( , )(C). So:

(i) r belongs to an interpretation of Back I ( , )(C).


skip (M, r, )(C).
Claim 6E.3. Player I has a winning strategy in G

Proof. The claim follows from condition (i) using: Lemma 6D.11 if is a Woodin limit
of Woodin cardinals in M; Lemma 6C.14 if is a relative limit in W ; and an inductive
application of Lemma 6E.2 if is a relative successor in W . #

Remember that our goal is to show that player I has a winning strategy in
skip (M, r, )(C ). One can obtain such a strategy by composing Claim 6E.3 with
G
an application of Lemma 6E.1. Let us quickly describe this composition. We describe
how to play for I in G skip (M, r, )(C ), and win.
Start by following a winning strategy for I in G skip (M, r, )(C), given by Claim 6E.3.
This may lead to a victory for I through one of the snags described in conditions (I1)
(I7) in Section 6A, and in this case we are done. Otherwise it leads to a position which
is won by player I in G skip (M, r, )(C), through one of the payoff conditions (P1) and
(P2). In other words it leads to:

(ii) a wellfounded model M+1 ;

(iii) an elementary embedding j0,+1 : M M+1 ; and

(iv) a j0,+1 ()-sequence t+1 over M+1 , which belongs to an interpretation of


j0,+1 (C).
6E Relative successors and compositions 247

Note that M+1 &j0,+1 ( ) + 1 is countable in V. This follows from Remark 6A.16
and the initial assumption that M& + 1 is countable in V. Using condition (iv)
an application of Lemma 6E.1 over M+1 shows that I has a winning strategy in
skip (M+1 , t+1 , j0,+1 ( ))(j0,+1 (C )). Continue by following this strategy. This
G
may lead to a victory for I through one of the snags described in conditions (I1) and
(I2) in Section 6A, and in this case we are done. Otherwise it leads to:

(v) a wellfounded model M+2 ;

(vi) an elementary embedding j+1,+2 : M+1 M+2 ; and

(vii) a j0,+2 ( )-sequence t+2 over M+2 , which belongs to an interpretation of


j0,+2 (C ).
skip (M, r, )(C ).
This is a victory for I in G # (Lemma 6E.2)

Let us continue to work under the assumption that is a relative successor in


W , and is the largest Woodin cardinal of M below . The next lemma handles a
situation which corresponds to case (V) in the definition of the pullback operation in
Section 4D (3).

Lemma 6E.4. Suppose that is a Woodin limit of Woodin cardinals in M. Let t be


a -sequence over M. Let > be a Woodin cardinal in M. Let C be a -name
in M.
Suppose that M& +1 is countable in V. Suppose that t belongs to an interpretation
of Back I (, )(C ), where the pullback is computed in M. Then player I has a winning
skip (M, t, )(C ).
strategy in G

Proof. The assumption that > implies that is greater than or equal to . If
= then the current lemma follows from Lemma 6E.1. So suppose that > .
Let C = Back I ( , )(C ). The definition of the pullback operation in case (V) in
Section 4D (3) is such that Back I (, )(C ) = Back I (, )(C ). So:

(i) t belongs to an interpretation of Back I (, )(C ).

Using Lemma 6E.1 it follows that:


skip (M, t, )(C ).
(ii) Player I has a winning strategy in G

Our goal is to show that player I has a winning strategy in G skip (M, t, )(C ). One
can obtain such a strategy by composing condition (ii) above with a follow-up strategy
given by Claim 6E.5 below. Let us quickly describe this composition. We describe how
skip (M, t, )(C ), and win.
to play for I in G
skip (M, t, )(C ), given
We start by letting player I follow her winning strategy in G
by condition (ii). This may lead to a victory for I through one of the snags described
in conditions (I1) and (I2) in Section 6A, and in this case we are done. Otherwise it
leads to:
248 6 Along a single branch

(iii) a wellfounded model M1 ;

(iv) an elementary embedding j0,1 : M M1 ; and

(v) a j0,1 ( )-sequence t1 over M1 , which belongs to an interpretation of j0,1 (C ).


skip (M1 , t1 , j0,1 ( ))(j0,1 (C )).
Claim 6E.5. I has a winning strategy in G

Proof. Note that M1 &j0,1 ( ) + 1 is countable in V. This can be seen using Re-
mark 6A.16 and the initial assumption that M& + 1 is countable in V. The claim now
follows from condition (v) through an application over M1 of the appropriate lemma in
the following list: Lemma 6C.14 if is a relative limit in W ; Lemma 6D.11 if is a
Woodin limit of Woodin cardinals in M; and Lemma 6E.2 if is a relative successor
in W . #

Remember that we are working to describe a winning strategy for player I in



Gskip (M, t, )(C ). So far we played mega-round 0 and reached the position de-
scribed in conditions (iii)(v). We continue by letting player I follow her winning
strategy in G skip (M1 , t1 , j0,1 ( ))(j0,1 (C )), given by Claim 6E.5. This is possible
since the rules for mega-round in G skip (M1 , t1 , j0,1 ( )) are precisely the same as the
rules for mega-round 1 + in the current run of G skip (M, t, ). Following the strategy
given by Claim 6E.5 may lead to a victory for I through one of the snags described in
conditions (I1)(I7) in Section 6A, and in this case we are done. Otherwise it leads to
a position which is won by player I in G skip (M1 , t1 , j0,1 ( ))(j0,1 (C )), through one
of the payoff conditions (P1) and (P2). In other words it leads to:

(vii) a wellfounded model M+1 ;

(viii) an elementary embedding j1,+1 ; and

(ix) a j1,+1 (j0,1 ( ))-sequence t+1 over M+1 , which belongs to an interpretation
of j1,+1 (j0,1 (C )).
skip (M, t, )(C ), so again we are done.
This position is a victory for I in G
# (Lemma 6E.4)

6F Skips
We work in this section to combine various strategies for I in skipping games into a
branch . Let us start by consolidating our knowledge of strategies in
strategy for I in G
skipping games.

Definition 6F.1. Let M be a transitive model of ZFC , let < be Woodin cardinals
of M, let C be a -name in M, and let t be a -sequence over M.
The tuple $M, t, , C % is called promising just in case that:
6F Skips 249

(1) M& + 1 is countable in V; and


(2) t belongs to an interpretation of Back I (, )(C ) (computed in M).
We omit in the notation since it can be recovered from rdm(t).
Lemma 6F.2. Let $M, t, , C % be promising. Then player I has a winning strategy
skip (M, t, )(C ).
in G
Proof. Let be the unique ordinal so that t is a -sequence. The current lemma is
simply: Lemma 6C.14 if is a relative limit in W and W ; Lemma 6D.11 if is
a Woodin limit of Woodin cardinals in M and W ; Lemma 6E.2 if is a relative
successor and W ; and Lemma 6E.4 if is a Woodin limit of Woodin cardinals in M.
#
Let  skip be a function which associates to each promising tuple $M, t, , C % a
skip (M, t, )(C ). Such
skip (M, t, , C ) which is winning for player I in G
strategy 
a strategy exists by Lemma 6F.2.
We intend to combine instances of  skip into a strategy in G branch , with the goal

of showing that I wins Gbranch (M, t, )(C ) whenever $M, t, , C % is promising.

Sections 6F (1) through 6F(3) establish terminology which connects positions in G branch

with positions in Gskip . Section 6F(4) gives an informal idea of the argument which
sews together instances of  skip to produce a strategy in G branch . Sections 6F (5) and
6F (6) give the actual argument.

6F (1) Routes. Let $M, t, , C % be promising. We work here to connect positions in


branch (M, t, )(C ) with positions in G
G skip (M, t, )(C ).
We think of positions of mega-rounds in G skip (M, t, )(C ) as being given by

sequences $T , b , C , h , E , t+1 | < %. Not all the objects listed are defined
for each . For example C and h are only defined if is a successor, and only if I
called for a skip in mega-round . E is only defined if is a standard limit, and only
if II elected a leap in mega-round .
There are moves in mega-round of G skip (M, t, )(C ) which are not listed among
the objects T , b , C , h , E , and t+1 . But all the objects played can be recovered
from the objects listed. For example w and y can be recovered from (t and) t+1 ,
and M+1 in the case of a skip can be recovered from (M and) h . So the se-
quence $T , b , C , h , E , t+1 | < % gives a complete account of a position in
skip (M, t, )(C ).
G
skip (M, t, )(C ) is a sequence
Definition 6F.3. A reduced position in G
R = $T , b , h , E , t+1 | < %

which can be expanded, via the addition of objects $C | < %, to a position in


skip (M, t, )(C ). The combined sequence $T , b , C , h , E , t+1 | < % is
G
called an expansion of R.
250 6 Along a single branch

Definition 6F.4. A reduced position R = $T , b , h , E , t+1 | < % is said to


skip (M, t, , C ) just in case that it can be expanded to a position
be consistent with 
which is played according to  skip (M, t, , C ).

Let R = $T , b , h , E , t+1 | < % be a reduced position consistent with


skip (M, t, , C ). Note that the moves C correspond to player I in G
 skip . It follows

from this that there is a unique expansion of R to a position which is played accord-
ing to  skip (M, t, , C ). We refer to this unique expansion as the position given
by R. (There is some abuse of notation here, since the notion depends on the tuple
$M, t, , C % through the reference to  skip (M, t, , C ). But this tuple is always
clear from the context.)
We think of positions in G branch (M, t, )(C ) as being given by sequences

$T , b , E , t +1 | < %. (This notation was already introduced in Section 6A.)
Here again not all objects are defined for each . E is only defined if is a standard
limit, and only if II elected a leap in mega-round . Again there are moves in mega-
round which are not among the objects listed, the objects T , b , E , and t +1 that is.
But all the moves in mega-round can be recovered from these objects.
Let be an ordinal. Let P = $T , b , E , t +1 | < % be a position of length
in G branch (M, t, )(C ). We work with this fixed P for the rest of Section 6F (1).
P gives rise to:
models M for ; and
embeddings j, : M M for < .
The definitions and claims below are made with reference to these models and embed-
dings.
Definition 6F.5. A skip frame for P is a function f , from some ordinal + 1 into + 1,
satisfying the following conditions:
(1) f : + 1 + 1 is increasing and continuous at limits;
(2) f (0) = 0 and f ( ) = ;
(3) if < is a limit then f ( + 1) = f () + 1;
(4) if < is a successor or zero then f ( + 1) is a successor, either equal to or
greater than f () + 1.
is called the length of f , denoted lh(f ).
Claim 6F.6. Let f be a skip frame for P . Then f maps limit ordinals to limit ordinals,
successor ordinals to successor ordinals, and zero to zero.
Proof. That successors are mapped to successors follows from conditions (3) and (4)
in Definition 6F.5. That limits are mapped to limits follows from condition (1). That
zero is mapped to zero follows directly from condition (2). #
6F Skips 251

Remark 6F.7. As a special case of Claim 6F.6 we see that lh(f ) is a limit (respectively
successor, zero) iff lh(P ) is a limit (respectively successor, zero).

Definition 6F.8. Let f be a skip frame for P . For each < lh(f ) let:

jf ()+1,f (+1) if f ( + 1) > f () + 1; and
h =
undefined if f ( + 1) = f () + 1.

The definition of h is made with reference to both f and P . (P is needed to give


rise to the embeddings j, .) But we suppress this dependence in the notation.
Define red(P , f ) to be the sequence $Tf () , bf () , h , Ef () , tf (+1) | < lh(f )%.
Notice that this sequence has the format of a reduced position in G skip (M, t, )(C ).
The notation red stands for reduced. We are using the skip frame f to convert
P into a reduced position R = red(P , f ) in the skipping game. If is such that
f ( + 1) = f () + 1 then we take mega-round f () in P and pass it without change
to be mega-round in R. If is such that f ( + 1) > f () + 1 then we take mega-
rounds in P for all [f (), f ( + 1)) and lump them together into mega-round
in R. The lumping, which is illustrated in Diagram 6.6, is done by calling for a skip
in mega-round of R and using h = jf ()+1,f (+1) for that skip.

kkk
Mf () kSk
SkSSSS / Qf () = Mf ()+1 / Mf (+1)
jf ()+1,f (+1)
Tf ()

kkk +
Mf () kSk
SkSSSS / Qf () Mf (+1)
Tf ()

Diagram 6.6. Mega-rounds f () to f ( + 1) in P (upper line) lumped together into a single


mega-round in R (lower line).

Definition 6F.9. Let P be a position in G branch (M, t, )(C ). Let f be a skip


frame for P . f is a route to P just in case that red(P , f ) is a reduced position in
skip (M, t, )(C ), consistent with 
G skip (M, t, , C ).

Intuitively a route to P is a way to generate P through a play of the skipping game,


following Is winning strategy  skip (M, t, , C ). The definition of a route depends

on M, t, , and C through the reference to  skip (M, t, , C ). But we suppress this
dependence in the notation.

Claim 6F.10. Let f be a route to P . Let be a successor ordinal smaller than lh(f ).
Then mega-round in red(P , f ) contains a skip just in case that f ( + 1) > f () + 1.
252 6 Along a single branch

Proof. This is immediate from the definitions. Mega-round in red(P , f ) contains a


skip iff h is defined, and this by Definition 6F.8 happens precisely when f ( + 1) is
greater than f () + 1. #
By a P -route we mean a route either to P or to a proper initial segment of P .
Lemma 6F.11. Let f be a P -route. Let = lh(f ). Then there is at most one P -route
of length + 1 extending f .
Proof. If f ( ) = lh(P ) then there are no P -routes which extend f at all. So suppose
that f ( ) < lh(P ). Let denote f ( ).
Let f and g be P -routes of length + 1, both extending f . Let = f ( + 1)
and let = g ( + 1). We work to show that = . Since both f and g extend
f this is enough to establish that f = g .
If is a limit then both f ( +1) and g ( +1) must equal f ( )+1 by condition (3)
in Definition 6F.5, and in particular f ( + 1) = g ( + 1). So suppose that is a
successor (or zero). Using Claim 6F.6 it follows that too is a successor (or zero).
Since is a successor (or zero), mega-round in P is played according to rules
(S1)(S4) in Section 6A. Let w , y , T , and b be the moves played in mega-round
of P . Let k be the direct limit embedding along the branch b of T . The settings in
branch are such that:
G
(i) j, +1 = k .
Let R denote red(P  , f ) and let S denote red(P  , g ). Let R denote
red(P , f ). Then:
(ii) R and S are reduced positions of length + 1;
(iii) both R and S extend R;
skip (M, t, , C ); and
(iv) both R and S are consistent with 
(v) in both R and S the moves corresponding to rules (S1)(S4) in mega-round
are precisely the moves w , y , T , and b from mega-round of P .
The last condition follows from the fact that both f ( ) and g ( ) are equal to f ( ),
which is equal to .
Let K denote the position in G skip (M, t, )(C ) which consists of the mega-
rounds given by R, followed by the moves w , y , T , and b for rules (S1)(S4) in
mega-round . Both R and S extend K, and do so in a manner consistent with
skip (M, t, , C ). We shall use this to argue that in fact they are equal. We divide

the proof into two cases, depending on whether  skip (M, t, , C ) calls for a skip or
for an early end following K.
Case 1. If  skip (M, t, , C ) elects an early end following the position given by K. In
this case both R and S , which extend K and follow the dictates of  skip (M, t, , C ),
must take an early end in mega-round . Using the appropriate instances of Claim 6F.10
6F Skips 253

it follows that f ( + 1) = f ( ) + 1 and that g ( + 1) = g ( ) + 1. Since f ( )


and g ( ) are both equal to = f ( ) it follows from this that f ( + 1) = g ( + 1),
as required. # (Case 1)

Case 2. If  skip (M, t, , C ) calls for a skip following the position given by K. In
this case both R and S must contain a skip in mega-round . The skips are made
according to the format set in rules (S5) and (S6) in Section 6A.
Let a , Ca , ha , Ma +1 , and ta +1 be the moves corresponding to the skip in mega-
round of R . Let b , Cb , hb , Mb +1 , and tb +1 be the moves corresponding to the
skip in mega-round of S .
Since both R and S extend K in a manner consistent with the strategy

skip (M, t, , C ), the moves corresponding to player I are the same in R and in S .
In particular this means that:

(vi) a = b .

This ultimately will allow us to show that = .


Since R = red(P  , f ), the moves ha and ta +1 are the ones induced by P 
and f through the definition of red(. . . ) above. So:

(vii) ha is equal to jf ( )+1,f ( +1) and ta +1 = tf ( +1) .

Recall that denotes f ( + 1), and denotes f ( ) which is the same as f ( ).


Folding this into the last condition we get:

(viii) ha is equal to j +1, and ta +1 = t .

The skip in mega-round of R is subject to rules (S5) and (S6) in Section 6A.
Using condition (2) in rule (S6) and condition (1) in rule (S5) we see that:

(ix) The relative domain of ta +1 is equal to (ha k )(a ) + 1.

Combining this with conditions (viii) and (i) above we get:

(x) rdm(t ) = j, (a ) + 1.

Recall that denotes g ( + 1). Working as we did above but with g and S we
get the following condition, which is a parallel for g of condition (x).

(xi) rdm(t ) = j, (b ) + 1.

Remember that we are trying to prove that = . Suppose this is not the case.
Assume for definitiveness that < .
branch are such that:
The rules of G

(xii) the critical point of j , is greater than rdm(t ); and

(xiii) t extends t strictly.


254 6 Along a single branch

Combining condition (x), the shift of condition (xi) via j , , and most importantly
condition (vi) we get:

(xiv) j , (rdm(t )) is equal to rdm(t ).

Conditions (xii) and (xiv) imply that rdm(t ) = rdm(t ), and this contradicts condi-
tion (xiii). # (Case 2)

# (Lemma 6F.11)
branch (M, t, )(C ). Let f and f
Corollary 6F.12. Let P and P be positions in G
be routes to P and P respectively. Suppose that P extends P ( perhaps not strictly).

Then f extends f ( perhaps not strictly).

Proof. Suppose for contradiction that f does not extend f . It is easy to check directly
from Definition 6F.5 that f cannot be a strict initial segment of f . So there must be
some ordinal on which f and f take different values. Let be the least such. Because
of the continuity requirement in Definition 6F.5, must be a successor ordinal. Let
= 1.
Let f = f  + 1. f is a P -route of length . f  + 2 and f  + 2 are both
P -routes of length + 1, extending f. Moreover f  + 2 and f  + 2 are distinct

since f and f take different values on = + 1. This contradicts Lemma 6F.11.


#
branch (M, t, )(C ) just in case that there
Definition 6F.13. P is said to be secure in G
is a route to P .
branch (M, t, )(C ). Then there exists exactly
Corollary 6F.14. Let P be secure in G
one route to P .

Proof. Suppose that f and g are two routes to P . Corollary 6F.12 implies that f
extends g, and that g extends f . So f = g. #

Corollary 6F.15. Let be an ordinal. Let $P | <  % be a strictly increasing


sequence of positions in G branch (M, t, )(C ). Let P =
< P . Suppose that each

P is secure in Gbranch (M, t, )(C ). Then P too is secure in G branch (M, t, )(C ).

Proof. For each < letf be the unique route to P . By Corollary 6F.12, $f | < %
is increasing. Let g = < f . Let be the domain of g. It is easy to see that
is a limit ordinal. Let be the length of P . It is easy to see that g is cofinal in .
Define f : + 1 + 1 by:

g() if < ; and
f () =
if = .

It is easy to see that f is a route to P . #


6F Skips 255

branch (M, t, )(C ), and secure. Then P is won


Claim 6F.16. Let P be terminal in G
by player I.
Proof. Let f be the unique route to P . Let R = red(P , f ). Using the fact that P is ter-
branch (M, t, )(C ) it is easy to see that R is terminal in G
minal in G skip (M, t, )(C ).
R is consistent with skip (M, t, , C ) by Definition 6F.9, and skip (M, t, , C ) is
a winning strategy for player I. So R = red(P , f ) is won by I in G skip (M, t, )(C ).
It follows from this that P is won by player I in G branch (M, t, )(C ).
#

6F (2) Extensions. Let $M, t, , C % be promising. Let P be secure in the game


branch (M, t, )(C ), and non-terminal. We work on ways to extend P by one mega-
G
round, in such a way that the extension is either secure, or else it translates to a position
skip in which player I calls for a skip. The former case of course sets the grounds
in G
for an iteration. In the latter case we show that Is moves give rise to a new promising
tuple.
Let f be the unique route to P . Let R = red(P , f ). Observe that R is non-terminal
skip (M, t, )(C ). This can be seen easily using the assumption that P is non-
in G
terminal in G branch (M, t, )(C ). Let = lh(P ) and let = lh(R). We follow
the notation established in Section 6F(1). P is given by a sequence $T , b , E , t +1 |
< %. The sequence gives rise to models M and embeddings j, for < .
We use G branch (M, t, , C )[P ] to denote mega-round of G branch (M, t, )(C )
following the position P . This is a game of just one mega-round, played according to
the rules of either the successor case or the limit cases in Section 6A.
We use G skip (M, t, , C )[R] to denote mega-round of G skip (M, t, )(C ) fol-
lowing the position given by R. This too is a game of just one mega-round, played
according to the rules of either the successor case for skipping games or the limit cases
in Section 6A.
branch (M, t, , C )[P ] and G
If is a limit then G skip (M, t, , C )[R] are precisely
the same game. They both start from M and t , and they are both played according to
the rules of the limit cases in Section 6A.
If is a successor or zero then G branch (M, t, , C )[P ] is an initial segment of
skip (M, t, , C )[R]. Both games start from M and t . G
G branch (M, t, , C )[P ]
is played according to rules (S1)(S4) in Section 6A. G skip (M, t, , C )[R] may
continue with rules (S5) and (S6).
We use  skip (M, t, , C )[R] to denote the restriction of  skip (M, t, , C ) to
mega-round following the position given by R.  skip (M, t, , C )[R] is thus a
strategy for player I in G skip (M, t, , C )[R].
Definition 6F.17. Define  branch (M, t, , C )[P ] to be the restriction of the strategy
skip (M, t, , C )[R] to the game G
 branch (M, t, , C )[P ].
Definition 6F.17 is made under the assumption that P is non-terminal and secure in
branch (M, t, )(C ). R stands for red(P , f ) where f is the unique route to P . Note
G
that the restriction mentioned in Definition 6F.17 makes sense, because G branch (. . . )[P ]
is either exactly equal to G skip (. . . )[R] or else it is an initial segment of that game.
256 6 Along a single branch

Claim 6F.18. Let P + be a position of length +1 in G branch (M, t, )(C ), extending P .


Suppose that is a limit ordinal. Suppose further that mega-round in P + is played
branch (M, t, )(C ).
branch (M, t, , C )[P ]. Then P + is secure in G
according to 

Proof. This is a simple matter of checking that the natural extension of the route which
leads to P gives a route which leads to P + . Recall that f is the unique route to P and
= lh(f ). Let f + : + 2 + 2 be the function defined by:

+ f () if ; and
f () =
+ 1 if = + 1.

We claim that f + is a route to P + . To prove this we must check that R + = red(P + , f + )


is consistent with  skip (M, t, , C ).
R  is equal to red(R, f ) which is consistent with 
+ skip (M, t, , C ) since f
is a route to P . So it is enough to check that mega-round in R + is consistent with
skip (M, t, , C )[R].

is assumed to be a limit ordinal. It follows from this that the two games
branch (M, t, , C )[P ] and G
G skip (M, t, , C )[R] are precisely the same, and that

the strategies branch (M, t, , C )[P ] and 
skip (M, t, , C )[R] are precisely the
+
same. The fact that mega-round in R is consistent with  skip (M, t, , C )[R]
+
thus follows from the assumption that mega-round in P is played according to
branch (M, t, , C )[P ].
 #

Claim 6F.19. Let P + be a position of length +1 in G branch (M, t, )(C ), extending P .


Suppose that is a successor or zero. Suppose further that mega-round in P + is played
branch (M, t, , C )[P ]. Then one of the following conditions holds:
according to 
branch (M, t, )(C ); or
(1) P + is secure in G
skip (M, t, , C )[R] calls for a skip following the moves made in mega-round
(2) 
of P + .

Proof. Define f + and R + as in the proof of Claim 6F.18. Mega-round in R + follows


skip (M, t, , C )[R] for the moves corresponding to rules (S1)(S4) in Section 6A,

and then calls for an early end. If this call is consistent with  skip (M, t, , C )[R]
then an argument similar to the one in Claim 6F.18 shows that P + is secure and we
obtain condition (1) in Claim 6F.19. If on the other hand  skip (M, t, , C )[R] calls
for a skip then we obtain condition (2). #
branch (M, t, )(C ) if it falls under
Definition 6F.20. We say that P + opens a skip in G
the situation of condition (2) in Claim 6F.19.

Claims 6F.18 and 6F.19 together show that an extension P + of P obtained


through a use of G branch (M, t, , C )[P ]or in other words through a use of
skip (M, t, , C )[R]is either secure or else it opens a skip. In the former case
G

we can iterate our work, moving onward to extend P + using  skip (M, t, , C )[R + ].
6F Skips 257

But in the latter case  skip (M, t, , C )[R] calls for a skip following the moves in
mega-round of P + , and there are no moves in P + to answer that call. It is therefore
impossible in the latter case to use  skip (M, t, , C ) to continue and extend P + .
We work for the rest of Section 6F(2) in the situation of this latter case, the case that
P + opens a skip. We shall see that the moves in this case give rise to a new promising
tuple, allowing access to a new instance of  skip which can serve as a temporary
replacement for the instance  skip (M, t, , C ) above.
Let w , y , T , and b be the moves made in mega-round of P + . These moves
correspond to rules (S1)(S4) in Section 6A. Let Q be the direct limit along the
branch b of T . Let k : M Q be the direct limit embedding. The settings in the
successor case of G branch are such that:

(i) M+1 = Q ;
(ii) j,+1 = k ;
(iii) +1 = k ( ); and
(iv) t+1 = t w , y .
The moves w , y , T , and b are played according to  branch (M, t, , C )[P ],
which by definition is the restriction of  skip (M, t, , C )[R] to rules (S1)(S4). We

are working under the assumption that P + opens a skip.  skip (M, t, , C )[R] there-
fore calls for a skip after the moves w , y , T , and b .
Let and C be the moves played by  skip (M, t, , C )[R] subject to rule (S5),
in response to w , y , T , and b . Following the notation in Section 6A let be the
first Woodin cardinal of M above rdm(t ), and let W be the class W computed in M .
Translating the conditions in rule (S5) to the current context we see that:
(v) belongs to W ;

(vi) is greater than and smaller than j0, ( );

(vii) C is a -name in M ; and

(viii) t , w , y belongs to an interpretation of k (C ) where C is equal to


Back I ( , )(C ) computed in M .
(The indexing is a bit awkward here. Let us remind the reader that mega-round in
branch corresponds to mega-round in G
G skip .)
Let = k ( ) and let C = k (C ). With this notation the conditions above
+ +

combine to yield the following claims:


Claim 6F.21. + is smaller than j0,+1 ( ).
Proof. Immediate from the definition of + and conditions (ii) and (vi). #
Claim 6F.22. $M+1 , t+1 , + , C + % is promising.
258 6 Along a single branch

Proof. The fact that M& + 1 is countable in V, the fact that j0,+1 preserves count-
ability (see Remark 6A.4), and the last claim combine to imply that M+1 & + + 1 is
countable in V. This gives condition (1) in Definition 6F.1. Condition (2) in Defini-
tion 6F.1 translates in our context to the statement that t+1 belongs to an interpretation
of Back I (+1 , + )(C + ). This statement follows from condition (viii) above, using the
definitions of + and C + , and conditions (iii) and (iv). #

Now that we know that $M+1 , t+1 , + , C + % is promising we may attempt to


use  skip (M+1 , t+1 , + , C + ), instead of  skip (M, t, , C ), to create exten-
+
sions of P . Section 6F(3) below develops the terminology needed for such a use.
skip (M+1 , t+1 , + , C + ) can be used until finally it produces a position which is ter-

minal in the respective instance of G skip (and of course won by player I). We shall see
that at that point it is possible to return to using  skip (M, t, , C ).

6F (3) Closing skips. Let $M, t, , C % be promising. Let be an ordinal. Let


branch (M)(t, , C ). Q
Q = $T , b , E , t +1 | < % be a position of length in G
gives rise to:

models M for ; and

embeddings j, : M M for < .

Let be an ordinal smaller than . Let P = Q and let P + = Q + 1. Suppose that:


branch (M, t, )(C );
(C1) P = Q is secure in G
branch (M, t, , C )[P ]; and
(C2) mega-round in Q is played according to 
branch (M, t, , C ).
(C3) P + = Q + 1 opens a skip in G

Let + and C + be defined as in Section 6F(2). By Claims 6F.21 and 6F.22, + is smaller
than j0,+1 ( ) and, most importantly, $M+1 , t+1 , + , C + % is promising.

Definition 6F.23. tail(Q, ) denotes $T , b , E , t +1 | [ + 1, )%.

tail(Q, ) is the sequence consisting of mega-rounds + 1 and above in Q. It


branch (M+1 , t+1 , j0,+1 ( ))(j0,+1 (C )). If rdm(t +1 )
is literally a position in G
j+1, +1 ( ) + 1 for each [ + 1, ) then tail(Q, ) may also be regarded as a
+

position in G branch (M+1 , t+1 , + )(C + ).


Let us suppose that:

(C4) rdm(t +1 ) j+1, +1 ( + ) + 1 for each [ + 1, ), so that tail(Q, ) is a


branch (M+1 , t+1 , + )(C + );
position in G
branch (M+1 , t+1 , + )(C + ); and
(C5) tail(Q, ) is terminal in G
branch (M+1 , t+1 , + )(C + ).
(C6) tail(Q, ) is won by player I in G
6F Skips 259

Lemma 6F.24 (under assumptions (C1)(C6) above). Exactly one of the following
conditions holds:
branch (M, t, )(C ) and won by player I; or
(1) Q is terminal in G
branch (M, t, )(C ).
(2) Q is non-terminal and secure in G

Proof. By assumptions (C5) and (C6) the position tail(Q, ) is terminal and won by
branch (M+1 , t+1 , + )(C + ). A victory in G
player I in G branch may come either through
one of the snags described in conditions (I1)(I7) in Section 6A, or through one of the
payoff conditions (P1) and (P2) in that section. We divide the proof into two cases
depending on which of these two possibilities holds for tail(Q, ).

Case 1. If tail(Q, ) is won by player I through one of the snags described in con-
ditions (I1)(I7) of Section 6A. It is easy to see that in this case Q is terminal in
branch (M, t, )(C ) and won by player I through exactly the same snag. So we
G
obtain condition (1) in Lemma 6F.24. # (Case 1)

Case 2. If tail(Q, ) is won by player I through one of the payoff conditions (P1) and
(P2), in the game G branch (M+1 , t+1 , + )(C + ). Translating these conditions to the
current context we see that:

(i) is a successor ordinal;

(ii) t is a j+1, ( + )-sequence over M ; and

(iii) t belongs to an interpretation of j+1, (C + ).

By Claim 6F.21, + < j0,+1 ( ). So condition (ii) implies that the relative domain
of t is strictly smaller than j0, ( ). It follows from this that Q is non-terminal in
branch (M, t, )(C ). To obtain condition (2) in Lemma 6F.24 we must check that Q
G
is secure in G branch (M, t, )(C ).
Following the notation in Section 6F(2) let f be the unique route to P , let = lh(f ),
and let R = red(P , f ). Define f : + 2 + 1 by:


f () if < ;

f () = if = ; and


if = + 1.

The top two cases could be merged into one, since f ( ) = by condition (2) in
Definition 6F.5. But we wish to emphasize the fact that f sends to and + 1 to .
Using the fact that is a successor ordinal it is easy to check that f is a skip frame
for Q. f extends f . Let R = red(Q, f ). R then has the format of a reduced
position of length + 1 in G skip (M, t, )(C ), extending R. Mega-round of R is
given by T , b , h , and t , where:

(iv) h = j+1, .
260 6 Along a single branch

(We are simply copying from Definition 6F.8, using the fact that f ( ) = and
f ( + 1) = .)
skip (M, t, )(C ), consistent with
Claim 6F.25. R is a reduced position in G

skip (M, t, , C ).

Proof. Since R is a reduced position consistent with  skip (M, t, , C ) we only have

to worry about mega-round in R . Let w , y , T , and b be the moves in mega-
round of Q. These are the moves corresponding to rules (S1)(S4) in mega-round
of R . These moves are consistent with  skip (M, t, , C )[R] by condition (C2).
By condition (C3), P + = Q + 1 opens a skip in the game G branch (M, t, )(C ).

skip (M, t, , C )[R] therefore calls for a skip in response w , y , T , and b .
Let and C be the moves, for rule (S5) of the skip case in Section 6A, played by
skip (M, t, , C )[R] in response to w , y , T , and b .
the strategy 
Remember that our goal is to verify that mega-round in R is legal in

Gskip (M, t, )(C ) and consistent with  skip (M, t, , C )[R]. The moves corre-
sponding to rule (S6) in mega-round of R are h , M , and t . We have to check
that they can be regarded as a legal response to the moves and C made by
skip (M, t, , C )[R]. In other words we have to check that they satisfy the con-

ditions of rule (S6) in Section 6A, translated to the current context.
Most of the conditions in rule (S6) are easy to verify. Let us just comment on condi-
tion (4). Translated to the current context it says that t must belong to an interpretation
of (h k )(C ), or in other words to an interpretation of j+1, (C + ). (The move
from (h k )(C ) to j+1, (C + ) uses condition (iv) above and the definition of C + in
Section 6F (2).) The fact that t does indeed belong to an interpretation of j+1, (C + )
follows from condition (iii) above. It is through the appeal to condition (iii) that we
are using the crucial assumption in condition (C6), that tail(Q, ) is won by player I.
# (Claim 6F.25)

It follows from the last claim that f is a route to Q. Q is therefore secure in


branch (M, t, )(C ).
Gbranch (M, t, )(C ). We already saw that Q is non-terminal in G
So we get condition (2) in Lemma 6F.24. # (Case 2)

# (Lemma 6F.24)

6F (4) Discussion. We have now the necessary tools for sewing together instances of
skip into winning strategies for player I in G
 branch . The actual argument which does
this is given in Sections 6F(5) and 6F(6). Let us here try to view the argument as an
induction.
Fix a promising tuple $M, t, , C %. We aim to show that player I has a winning
branch (M, t, )(C ). Suppose inductively that:
strategy in G

() Player I has a winning strategy in the game G branch (N, s, + )(C + ) whenever
$N, s, + , C + % is promising and + < j () for some elementary j : M N .
6F Skips 261

We now describe how to play for I in G branch (M, t, )(C ), and win.
Start playing by following  skip (M, t, , C ). This can be done until, if ever, a
position is reached where that strategy calls for a skip. (Notice that G branch is the same
as G skip so long as skips are not called for.)
If no skip is ever called for then  skip (M, t, , C ) eventually leads to a victory
for player I. So suppose that at some point a position, P + of length + 1 say, is reached
where  skip (M, t, , C ) calls for a skip.
Notice that P + opens a skip in the sense of Definition 6F.20. Let + and C +
be defined as in Section 6F(2). By Claims 6F.22 and 6F.21, $M+1 , t+1 , + , C + % is
promising, and + < j0,+1 ( ). Using the inductive condition () above it follows
that I has a winning strategy in G branch (M+1 , t+1 , + )(C + ). Let  + be some such
strategy.
Continue now playing for I in G branch (M, t, )(C ), from the position P + onward,
by following  . More precisely, continue by creating extensions Q of P + subject to
+

the condition that tail(Q, ) is played according to  + . This will lead to a position Q
so that tail(Q, ) is won by I in G branch (M+1 , t+1 , + )(C + ).

If Q is terminal in Gbranch (M, t, )(C ) then by case 1 in Lemma 6F.24 it is
won by player I in this game, as required. So suppose that Q is not terminal in
branch (M, t, )(C ). Again by Lemma 6F.24, this time case 2, it must be that Q
G
allows closing the skip that was opened by P + . Thus, from Q onward it is possible to
return to using  skip (M, t, , C ).
Proceed by iterating this process. From Q follow  skip (M, t, , C ) until, if ever,
it calls for a skip. At that point make another use of the inductive condition () to
play for I until reaching a position where the skip can be closed, etc.
The sketch above shows how a use of  skip (M, t, , C ) can be combined with
uses of the inductive condition () to play for I in G branch (M, t, )(C ). The key
points are: (a) the way rule (S5) in Section 6A (2)the rule that opens a skipgives
rise to a new promising tuple; and (b) the way a victory for I in the instance of G branch
corresponding to this new promising tuple gives precisely the moves needed for rule (S6)
in Section 6A (2)the rule that closes a skip. These two points are given precisely in
Claim 6F.22 and Lemma 6F.24.
There is though a problem with the sketch. The condition () is not, strictly speaking,
inductive. It has an inductive nature, in that it provides for lower promising tuples
$N, s, + , C + % the property which we are trying to prove for the tuple $M, t, , C %.
If lower were to mean a tuple with + < then the condition would actually be
inductive. But lower, as stated and applied above, means a tuple with + < j ( ) for
some elementary j : N M.
Condition () may thus be viewed as inductive not over ordinals in M, but over
ordinals in some directed system of iterates of M. Unfortunately that system need not
be wellfounded. Still, any illfoundedness reached in a play of G branch is automatically
won by player I through one of the snags (I1), (I3), (I5), or (I7) in Section 6A. So
some hope remains that an argument whose spirit follows the sketch given above may
produce a winning strategy for I in G branch (M, t, )(C ).
262 6 Along a single branch

Sections 6F(5) and 6F(6) formulate such an argument. Induction is replaced


by nesting. When reaching a position P + which opens a skip we do not appeal to
condition () to obtain a winning strategy for I in G branch (M+1 , t+1 , + )(C + ), but
rather nest a construction of such a strategy into the main construction of a strategy
branch (M, t, )(C ). The nested construction may itself reach a skip, in which
in G
case we nest another construction inside the nested construction, etc. We can proceed
this way, and eventually reach a terminal position won by player I, for as long as the
nesting levels are finite. We shall see that infinite nesting, if it occurs, leads to a position
which gives rise to an illfounded iterate of Mthis is related to the nature of lower
explained above and follows precisely from Claim 6F.21and such positions too are
won by player I.

6F (5) Safe positions. Let $M, t, , C % be promising. Let be an ordinal. Let Q =


branch (M, t, )(C ).
$T , b , E , t +1 | < % be a position of length in the game G
Q gives rise to:
models M for ; and
embeddings j, : M M for < .
The definitions below are made with reference to these models and embeddings.
branch (M, t, )(C ) just in case that it is secure in
Definition 6F.26. Q is level 0 in G
branch (M, t, )(C ).
G

branch (M, t, )(C ). Let


Suppose for the next definition that Q is not secure in G
< be largest so that Q is secure in that game. The existence of a largest such
follows from Corollary 6F.15. Let P = Q, and let P + = Q + 1.
branch (M, t, )(C ) just in case that:
Definition 6F.27. Q is level n + 1 in G
branch (M, t, , C )[P ];
(1) mega-round in Q is played according to 
branch (M, t, )(C ); and
(2) P + opens a skip in G
branch (M+1 , t+1 , + )(C + ).
(3) tail(Q, ) is a level n position in G
Definition 6F.27 is made by induction on n, simultaneously for all Q and all promis-
ing tuples, where the base case is given by Definition 6F.26. Notice that condition (3)
in Definition 6F.27 makes sense inductively, since the tuple $M+1 , t+1 , + , C + % is
promising by Claim 6F.22.
branch (M, t, )(C ) if it is level n for some n < .
Definition 6F.28. Q is safe in G
If Q is safe then one can check that there is a unique n < witnessing this. We
branch (M, t, )(C ).
refer to this unique n as the level of Q in G
Claim 6F.29. Suppose that Q is safe in G branch (M, t, )(C ). If Q is terminal in
branch (M, t, )(C ) then it is won by player I.
G
6F Skips 263

Proof. Work by induction on the level of Q in G(M, t, )(C ), simultaneously for all
Q and all promising $M, t, , C %. The base case is given by Claim 6F.16. We leave
the inductive case to the reader, and only note that it uses case 1 in Lemma 6F.24. #

Eventually we intend to play for I in G branch (M, t, )(C ) in a way that always
sticks to safe positions. We will formulate a strategy that does this later on. Let us here
check that this tactic does not fail at limits.

Claim 6F.30. Suppose that is a limit ordinal. Suppose that every strict initial segment
branch (M, t, )(C ), but Q itself is not. (In particular Q is not secure.)
of Q is safe in G
Let < be largest so that Q is secure in G branch (M, t, )(C ). Let P = Q
and let P = Q + 1. Then:
+

branch (M, t, , C )[P ]; and


(1) mega-round in Q is played according to 
branch (M, t, )(C ).
(2) P + opens a skip in G

Proof. Q + 1 is a strict initial segment of Q since is a limit. By assumption it must


branch (M, t, )(C ). It cannot be secure, because of the maximality of .
be safe in G
So it must be level n + 1 for some n. The conditions in Definition 6F.27, applied on
Q + 1, yield the conclusion of the current claim. #

Claim 6F.31. Suppose that is a limit ordinal. Suppose that every strict initial segment
branch (M, t, )(C ), but Q itself is not.
of Q is safe in G
Let < be largest so that Q is secure in G branch (M, t, )(C ). Let P = Q,
let P = Q + 1, and let and C be defined as in Section 6F (2). Then every strict
+ + +

initial segment of tail(Q, ) is safe in G branch (M+1 , t+1 , + )(C + ), but tail(Q, ) itself
is not.

Proof. Let be larger than and smaller than or equal to . Using the previous
claim and working directly with Definition 6F.27 its clear that Q is level n + 1 in
branch (M, t, )(C ) iff tail(Q, ) is level n in G
G branch (M+1 , t+1 , + )(C + ). The
current claim follows immediately from this. #

Corollary 6F.32. Suppose that is a limit ordinal. Suppose that every strict initial
branch (M, t, )(C ), but Q itself is not. Then M is illfounded.
segment of Q is safe in G

Proof. Iterated applications of the previous claim produce sequences $k | k < %,


$k+ | k < %, and $Ck+ | k < % so that:

(i) $k | k < % is an increasing sequence of ordinals smaller than ;

(ii) for each k < , $Mk +1 , tk +1 , k+ , Ck+ % is promising;

(iii) for each k < , every strict initial segment of tail(Q, k ) is safe in the game
branch (M +1 , t +1 , + )(C + ) but tail(Q, k ) itself is not; and
G k k k k
264 6 Along a single branch

+
(iv) for each k < , k+1 is smaller than jk +1,k+1 +1 (k+ ).

In obtaining conditions (ii) and (iv) we make use of Claims 6F.22 and 6F.21 respectively.
It is the use of Claim 6F.21 that is most important here. Condition (iv), which traces
back to this claim, implies that M is illfounded. #

Corollary 6F.33. Suppose that Q is a position of limit length. Suppose that every strict
branch (M, t, )(C ). Then either:
initial segment of Q is safe in G
branch (M, t, )(C ); or
(1) Q itself is also safe in G
branch (M, t, )(C ) and won by player I.
(2) Q is a terminal position in G

Proof. If M is illfounded then condition (I3) in Section 6A states that Q is terminal


branch and won by player I. If M is wellfounded then by the previous corollary Q
in G
must be safe. #

Corollary 6F.33 formalizes the method of dealing with infinite nesting indicated at
the end of Section 6F(4). The crucial ingredient leading to the corollary is Claim 6F.21,
which states that + is smaller than j0,+1 ( ) in situations which open a skip. This
same claim was the key to the induction discussed in Section 6F (4).

6F (6) The strategy. For each promising $M, t, , C % and for each position Q which
branch (M, t, )(C ) define a strategy (M,
is non-terminal and safe in G t, , C )[Q]
for player I in the game G branch (M, t, , C )[Q] subject to the appropriate condition
below:

(T1) If Q is secure in G branch (M, t, )(C ) then define (M,


t, , C )[Q] to be
branch (M, t, , C )[Q] of Definition 6F.17.
equal to the strategy 

(T2) Otherwise set (M, +1 , t+1 , + , C + )[tail(Q, )]
t, , C )[Q] equal to (M
where is largest so that Q is secure in G branch (M, t, )(C ), and + and C +
are defined as in Section 6F(2).

The definition of (M, t, , C )[Q] is made by induction on the level of Q in


branch (M, t, )(C ). The base case of level 0 is handled in condition (T1). The
G

inductive case is given by condition (T2). To make sense of this inductive condition
one has to know that tail(Q, ) is non-terminal in G branch (M+1 , t+1 , + )(C + ); that
it is safe in this game; and that its level in this game is smaller than the level of Q
in G branch (M, t, )(C ). The last two properties are immediate from the definitions
in Section 6F (5). The first property follows from the fact that Q is non-terminal in
branch (M, t, )(C ) using Claim 6F.29 and Lemma 6F.24: If tail(Q, ) were terminal
G
in G branch (M+1 , t+1 , + )(C + ) then by Claim 6F.29 it would be won by player I in this
game. But then by Lemma 6F.24, Q would have to be terminal in G branch (M, t, )(C ).
(Q cannot be secure in this game because of the case assumption of condition (T2).)
6F Skips 265

Claim 6F.34. Let $M, t, , C % be promising. Let P be non-terminal and secure in


branch (M, t, )(C ). Let P + extend P by one mega-round, and suppose that this
G
mega-round is consistent with  branch (M, t, , C )[P ]. Then P + is either level 0 or

level 1 in Gbranch (M, t, )(C ).

Proof. Let = lh(P ). If is a limit then P + is secure in G branch (M, t, )(C )


by Claim 6F.18 and therefore has level 0 in this game. If is a successor then by
branch (M, t, )(C ) or else it opens a skip in this
Claim 6F.19 P + is either secure in G
game. In the former case P + is level 0, and in the latter case it is easy to check directly
from the definitions in Section 6F(5) that P + is level 1. #

Claim 6F.35. Let $M, t, , C % be promising. Let Q be non-terminal and safe


in G branch (M, t, )(C ). Let Q+ extend Q by one mega-round, and suppose that
this mega-round is consistent with (M, t, , C )[Q]. Then Q+ is safe in
branch (M, t, )(C ).
G

Proof. By induction on the level of Q in G branch (M, t, )(C ). The base case follows
from Claim 6F.34. The inductive case is straightforward using the nature of condi-
tion (T2) and the definitions in Section 6F(5). #

Corollary 6F.36. Let $M, t, , C % be promising. Then player I has a winning strategy
branch (M, t, )(C ).
in G
branch (M, t, )(C ) as follows:
a strategy for player I in G
Proof. Define ,
if Q is (non-terminal and) safe in G branch (M, t, )(C ) then let 
copy

(M, t, , C )[Q] for moves in mega-round lh(Q) of Gbranch (M, t, )(C )

following Q; and
play these moves in your favorite fashion.
otherwise let 
is just the natural join of the strategies (M,
 t, , C )[Q] on safe positions Q. We
do not care what  does on positions which are not safe.

Let Q be a terminal position in G branch (M, t, )(C ), and suppose that Q is

consistent with . Using Claims 6F.35 and 6F.33 it is easy to prove by induction on
that for every lh(Q ), Q  is either safe in G branch (M, t, )(C ), or terminal and
won by player I (or possibly both). The successor case of the induction uses Claim 6F.35,
and the limit case uses Corollary 6F.33. (It is the use of Corollary 6F.33 that forces us
to include the possibility that Q  is terminal and won by player I.) Now Claim 6F.29
states that a safe position, if terminal, is won by player I. So taking = lh(Q ) above
we see that Q itself is in either case won by player I.
The argument of the preceding paragraph shows that terminal positions played
according to  are won by player I in G branch (M, t, )(C ). So  is a winning
strategy for I in this game. #
266 6 Along a single branch

6G Conclusion
The work of the previous section culminates in the following result:

Theorem 6G.1. Let M be a model of ZFC . Let < be Woodin cardinals of M.


Let C be a -name in M, and let t be a -sequence over M. Suppose that:

(1) M& + 1 is countable in V; and

(2) t belongs to an interpretation of Back I (, )(C ), where the pullback is com-


puted in M.
branch (M, t, )(C ).
Then player I has a winning strategy in G

Proof. This is simply Corollary 6F.36. #

The theorem should be viewed as formalizing our intuition in Chapter 4 that


Back I (, )(C ) names the set of positions from which player I can win to either
enter an interpretation of a shift of C , or reach a I-acceptable obstruction along the
way. We proved Theorem 6G.1 through iterated appeals to Lemma 6F.2 which dealt
with skipping games. Lemma 6F.2 in turn was proved by cases, depending on the nature
of and . The two main cases were the one corresponding to relative limits in W ,
handled in Section 6C, and the one corresponding to Woodin limits of Woodin cardinals,
handled in Section 6D. For the former we used the methods of Chapters 1 and 3 on
the names defined in Chapter 4. For the latter the driving force was our adherence to
clear annotated positions, which accumulate to form a generic for Woodins extender
algebra.

Corollary 6G.2. Let M be a model of ZFC , let be a Woodin cardinal of M, and let
C be a -name in M. Suppose that:

(1) M& + 1 is countable in V; and

(2) ini (, C) holds in M.


branch (M, , )(C).
Then player I has a winning strategy in G

The formula ini in Corollary 6G.2 is the one given by Definition 5G.2. in
branch (M, , )(C) is the empty annotated M-position of relative domain 0.
G

Proof of Corollary 6G.2. Let C0 = Back I (0 , )(C), 0 being the first Woodin cardinal
in M. Using Lemma 6B.2 and the assumption that ini (, C) holds it can be seen that
player I has a winning strategy in G branch (M, , 0 )(C0 ). Let 
0 be such a strategy.

The corollary can be proved by composing 0 with an application of Theorem 6G.1.
0 leads into an interpretation of a shift of C0 , and from there Theorem 6G.1 provides

a strategy for I which leads into an interpretation of a shift of C. We leave the exact
details to the reader. #
6G Conclusion 267

Remark 6G.3. If is a Woodin limit of Woodin cardinals in M then the assumption of


condition (1) in Corollary 6G.2 can be weakened to the assumption that is countable
in V. Similarly in Theorem 6G.1, if is a Woodin limit of Woodin cardinals in M then
the assumption of condition (1) can be weakened to the assumption that is countable
in V. This can be seen by noticing that Lemma 6D.11 assumes that , not M& + 1, is
countable in V, and tracing this fact through subsequent lemmas.
Chapter 7
Games which reach local cardinals

We work here to assemble the results of Chapters 4 through 6 into a powerful theorem on
determinacy. The strength of the theorem stems from its reference to Woodins extender
algebra of Section 4B. It makes runs of long games generic, not over a collapse algebra,
but over Woodins extender algebra.
Genericity over Woodins extender algebra is reached through a construction which
involves a non-linear composition of length iteration trees. Along each branch of this
non-linear composition the construction develops a run of the game G branch , and it is a
strategy given by the results of the previous chapter that perpetuates the construction.
(This provides some late justification for the title of the previous chapter. The work of
that chapter handles each single branch of the construction here.) It is too early to go
into details, but let us indicate that most of the work which ties the current construction
to the previous chapter is done in Sections 7B and 7C. The main theorem itself is
described in Section 7A, and proved in Sections 7D and 7E.
We end the chapter with applications of the main theorem. First we prove determi-
nacy for games which run up to the first which is not collapsed to by a function
constructible from the play. Then we consider similar games with constructible from
replaced by  in where  is a pointclass. Finally we present an application due to
Woodin, demonstrating the consistency of the statement that all ordinal definable games
of length 1 on natural numbers are determined.

7A Shifted payoff
Let M be a transitive model of ZFC , let  be an iteration strategy for M, and let be a
Woodin limit of Woodin cardinals in M. We work with these fixed objects for the rest
of the section. Our discussion refers to Woodins algebra W , defined in Section 4B.
Let A M be a W -name for a set of sequences of reals of length . (Of course we
local (M, , , A)
mean that the sequences have length , not the reals.) Define the game G
to be played as follows: Players I and II alternate natural numbers as usual creating
reals z for < 1V . If ever an < 1V is reached so that condition (P) below holds
(in V) then the game ends and player I wins. If no such is reached prior to 1V then
player I loses.
(P) There exists a countable length iteration tree U on M, leading to a final model
M and to an iteration embedding j : M M , and there exists some H , so
that:
(1) U is consistent with the iteration strategy ;
7A Shifted payoff 269

(2) H is j (W )-generic/M ; and


(3) the sequence $z | < % belongs to j (A)[H ].

Claim 7A.1. Let $z | < % be a position in G local (M, , , A) which is won by


player I. Let U and H witness this. Then is precisely equal to j ( ), where j is the
iteration embedding given by U.

Proof. This follows from clause (3) in condition (P) using the fact that A names a set
of sequences of length . #

Claim 7A.2. Let $z | < % be a position in G local (M, , , A) which is won by


player I. Let U and H witness this. Let M be the final model of U. Then:

(1) $z | < % belongs to M [H ]; and

(2) is a cardinal in M [H ].

Proof. The fact that $z | < % belongs to M [H ] is a direct consequence of clause (3)
in condition (P). Let us check that is a cardinal of M [H ].
Lemma 4B.19 states that W has the chain condition in M. So j (W ) has the
j ( ) chain condition in M . It follows that j ( ) is a cardinal in M [H ]. By the

previous claim is precisely equal to j (). So is a cardinal in M [H ]. #

Claim 7A.2 is the source of strength of the game G local . It provides a lower bound

on the length of any run of Glocal (M, . . . ) which is won by I. Specifically the claim says
that such a run must at the very least reach 1 in the sense of M [H ], where M [H ] is
some generic extension of an iterate of M and contains the run in question.
Let B M be another W -name for a set of sequences of reals of length . Define
the game H local (M, , , B) to be played as follows: Players I and II again alternate
natural numbers as usual to create reals z for < 1V . If ever an < 1V is reached
so that condition (Q) below holds (in V) then the game ends and player II wins. If no
such is reached prior to 1V then player II loses.

(Q) There exists a countable length iteration tree U on M, leading to a final model
M and to an iteration embedding j : M M , and there exists some H , so
that:

(1) U is consistent with the iteration strategy ;


(2) H is j (W )-generic/M ; and
(3) the sequence $z | < % belongs to j (B)[H ].

The game H local precisely mirrors the game G local . Note particularly that in the
local the payoff is phrased for player II, rather than I. Claims similar to 7A.1
case of H
and 7A.2 hold also for H local , but we do not phrase them explicitly.
270 7 Games which reach local cardinals

The main result of this chapter is the following theorem. It is similar in format to the
earlier Theorems 2A.2 and 3E.1. But the use of Woodins extender algebra in G local and

Hlocal makes the current theorem apply to much longer games than those considered in
Chapters 2 and 3, as we shall see in Section 7F.

Theorem (7E.1). Suppose that is countable in V. Then at least one of the following
three cases holds:
local (M, , , A);
(1) player I has a winning strategy in G
local (M, , , B); or
(2) player II has a winning strategy in H

(3) there exists some G V which is W -generic/M, and there exists some sequence
of reals $z | < % M[G], so that $z | < % belongs to neither A[G] nor
B[G].

Moreover M can distinguish this. More precisely, there are formulae and so that: if
M |= [, A] then case (1) holds; if M |= [, B] then case (2) holds; and otherwise
case (3) holds.

Remark 7A.3. Both conditions (P) and (Q) above require the length of the iteration
tree U to be countable. So only countable length iteration trees come up through the
statement of Theorem 7E.1. Nonetheless for the proof of the theorem it is essential that
 acts not only on countable length iteration trees but also on iteration trees of length
1V . The reason for this is explained in Remark 7D.6.

7B Layout
Let be an ordinal. By a tree order on we mean an order U so that:

(1) U is a sub-order of <;

(2) for each ordinal < , the set { | U } is linearly ordered by U ;

(3) for each so that + 1 < , the ordinal + 1 is a successor in U ; and

(4) for each limit ordinal < , the set { | U } is cofinal in .

A tree order is precisely the kind of order given by an iteration tree. Condition (2) allows
talking about successors in U . An ordinal < is a successor in U just in case that
{ | U } has a maximal element. (Note that maximal in U is the same as maximal
in <, by condition (1).) The following claim relates successors in U to successors in <.

Claim 7B.1. Let U be a tree order on . Let be smaller than . Then is a successor
in U iff it is a successor ordinal.
7B Layout 271

Proof. Suppose first that is a successor in U . { | U 0} is empty by condition (1),


so cannot be equal to 0. cannot be a limit ordinal by condition (4). So it must be a
successor ordinal.
Suppose next that is a successor ordinal, + 1 say. Then by condition (3) it is
also a successor in U . #

Remark 7B.2. We use UEQ to denote the non-strict order associated to U . More
precisely we set UEQ just in case that ( U ) ( = ).

Definition 7B.3. A branch through U is a subset of which is linearly ordered by U


and downward closed under U . A branch is cofinal if it cofinal in .

Let M be a model of ZFC . A tree of trees (henceforth referred to as a tot) of


length on M is a structure U which consists of the following objects, and which
satisfies conditions (S), (U), and (L) below:

a tree order U on ;

models M for < ;

embeddings j, : M M for U , commuting in the natural way;

for each < , a length iteration tree T on M ;

for each < , a cofinal branch b through T ; and

for some (possibly not all) so that + 1 < , an extender E in Q .

Q in the last item is the direct limit along the branch b of T . We use k : M Q
to denote the direct limit embedding.
Note that E need not be defined for all so that + 1 < . Given so that
+ 1 < we use the phrase E = undefined to indicate that E is not defined.
We allow padding in iteration trees. In fact T may be a tree which consists entirely
of padding. In this case Q is equal to M and k is equal to the identity.
We demand the following conditions from the objects of a tot:

(S) If E = undefined then M +1 = Q , the U -predecessor of + 1 is equal to ,


and j, +1 is equal to k .

(U) If E is defined then M +1 = Ult(M , E ) where is the U -predecessor of +1,


and j, +1 is equal to the ultrapower embedding.

(L) M for limit is the direct limit of the models $M | U % under the em-
beddings $j, | U U %. j, : M M for U are the direct limit
embeddings.
272 7 Games which reach local cardinals

The labels (S), (U), and (L) stand for simple, ultrapower, and limit respec-
tively. The situations of conditions (S) and (U) are illustrated in Diagrams 7.1 and 7.2.
(Note that may be equal to in the case of condition (U). Diagram 7.2 is not meant
to imply that it is always strictly smaller.) Condition (L) is the natural limit condition
for any kind of iteration. Note that the set of so that U is linearly ordered by U ,
so the condition makes sense.
kkk
/ M SSk
kkSSSS k / Q = M +1 /

T

Diagram 7.1. When E = undefined.

j, +1

kkk / '
SkSSS
E
/ M Sk
_ _ _ _ _ _ _ M k Q M +1 /
S k
T

Diagram 7.2. When E is defined.

A tot U of length on M may be regarded as one big iteration tree on M, consisting


of all the models M , all the models along the trees T , and all the models Q . We use
merge(U) to denote this iteration tree, which we call the merge of U. If is a successor
then merge(U) is an iteration tree of length + 1 on M, leading to the model Q1 .
If is a limit then merge(U) is an iteration tree of length on M, with the models
M for < forming its backbone.

Definition 7B.4. Let  be an iteration strategy for M. A tot U is consistent with  just
in case that the iteration tree merge(U) is consistent with .

We say that a tot U of length on M is regular just in case that it satisfies conditions
(R1)(R4) below:

(R1) For each < , M has (in order type) at least + 1 Woodin cardinals.

Let be the -th Woodin cardinal of M . By this we mean that is the first Woodin
cardinal of M so that there are (in order type) Woodin cardinals of M below it. The
existence of such is given by condition (R1). Let +1 = k ( ).

(R2) For each < , T only uses extenders with critical points greater than sup{ +1 |
< }.

(R3) All extenders used in T are countable in V.


7B Layout 273

(R4) For each so that + 1 < , E (if defined) is +1 + 1-strong in Q .

We think of conditions (R2) and (R3) as being satisfied trivially if T is the tree which
consists entirely of padding.
For the rest of the section let us work with a regular tot U of length on M. The
claims below establish facts about this regular tot which will later allow associating to
it a sequence of annotated positions.

Claim 7B.5. Let be smaller than . Then:

(1) for each < , +1 is precisely the -th Woodin cardinal of M ;

(2) M and all subsequent models in U agree past sup{ +1 | < };

(3) +1 is precisely the -th Woodin cardinal of Q ;

(4) Q and all subsequent models in U agree past +1 .

Proof. Prove all conditions simultaneously by induction on . The main driving force
in the proof is the agreement between models of U given by conditions (R2) and (R4).
We leave the exact details to the reader. #

Conditions (3) and (4) in the last claim can informally be viewed as saying that U
leaves a trail of Woodin cardinals in its progress.
For each limit ordinal < let = sup{ +1 | < }. is a limit of Woodin
cardinals in M . We say that is a phantom limit (in U) if is itself Woodin in M .
Otherwise we say that is a standard limit (in U). This distinction corresponds to the
branch in Section 6A.
division of limit cases in the definition of G

Claim 7B.6. Let be an ordinal smaller than . Then the sequence $ | is a standard
limit and % precisely enumerates, in increasing order, the ordinals smaller than
+1 which are limits of Woodin cardinals in Q+1 but not themselves Woodin.

Proof. From conditions (1), (2), and (3) in the previous claim it follows that $ +1 |
< % precisely enumerates the Woodin cardinals of Q below +1 . The current
claim follows from this and from the definition of a standard limit. #

Let K U denote the set { < | is a successor, or zero, or a standard limit}. This
is simply the set of ordinals smaller than , except those which are phantom limits.

Claim 7B.7. Let be an ordinal smaller than . Let W denote the class W of
Section 4A, computed in M . Then the sequence $ +1 | K U % precisely
enumerates, in increasing order, the set W .

Proof. Immediate from condition (1) in Claim 7B.5 and from the definition of K U . Let
us just note that for each < , +1 is a Woodin limit of Woodin cardinals in M just
in case that is a phantom limit (and in fact +1 is equal to in this case). #
274 7 Games which reach local cardinals

Claim 7B.8. Let < be a limit ordinal. Then sup{ + 1 | W } is precisely


equal to .
Proof. The set K U is cofinal in , since it contains all successors below . Using
this observation and Claim 7B.7 it follows that sup{ + 1 | W } is equal to
sup{ +1 + 1 | < }, which in turn is easily seen to be equal to . #
Define 0 = 0. For each successor ordinal + 1 < define +1 by:

+1 if is a phantom limit; and
+1 =
+1 + 1 otherwise.

Claim 7B.9. Let + 1 be a successor ordinal smaller than . Then sup{ + 1 |


+1 W +1 } is precisely equal to +1 .
Proof. From condition (1) in Claim 7B.5 (for = + 1) and from the definition of
+1 it follows that +1 is the largest Woodin cardinal of M +1 below +1 . If is
not a phantom limit then +1 belongs to W +1 by Claim 7B.7, and so sup{ + 1 |
+1 W +1 } = +1 + 1. If is a phantom limit then +1 does not belong to
W +1 , but arbitrarily large Woodin cardinals of M +1 below it do, so sup{ + 1 |
+1 W +1 } = +1 . #

Claim 7B.10. Let be smaller than . Then the sequence $ +1 | K U %


precisely enumerates, in increasing order, the set W .
Proof. This follows from Claim 7B.7. Simply note that there are no elements of W in
the interval [ , ). #
For an ordinal < and for W let e () = sup{ + 1 | W }. This
definition of e (. . . ) is simply the shift of the definition of e(. . . ) in Section 4A to M .
Claim 7B.11. Let belong to K U . Then e ( ) is precisely equal to .
Proof. The limit case follows from Claim 7B.8, the successor case follows from
Claim 7B.9, and the case of = 0 can be verified directly. #
Claim 7B.12. Let belong to K U and let < be greater than . Then e ( +1 ) is
equal to .

Proof. Recall that +1 = k ( ). The previous claim tells us that e ( ) = . k has


critical point greater than by condition (R2). So shifting the fact that e ( ) =
via k we see that e( +1 ), in the sense of Q , is equal to . The current claim follows
from this and the fact that Q and M agree past +1 . #
Claim 7B.13. Let be an ordinal smaller than . Let L denote the class L of Sec-
tion 4A, computed in M . Then the sequence $ | K U % precisely enumerates,
in increasing order, the set L .
7B Layout 275

Proof. The set L is precisely equal to {e ( ) | W }, which by


Claim 7B.10 is equal to {e ( +1 ) | K U }. The current claim follows from this
and Claim 7B.12. #
Equipped with the above claims and definitions we can comfortably associate anno-
tated positions to the tot U. By a U-sequence let us mean a sequence $w , y | K U %
so that:
(C1) for each K U , w is a witness for over M ;
(C2) if K U is a successor ordinal or zero then y is a real; and
(C3) if K U is a standard limit then y belongs to (Q & + 1) .
Let us for the rest of the section work with a fixed sequence of this kind.
Definition 7B.14. For each < define t to be the function given by the following
clauses:
(1) the domain of t is equal to { +1 | K U } { | K U };
(2) for each K U , t ( ) is equal to w ; and
(3) for each K U , t ( +1 ) is equal to y .
Claim 7B.15. t is an annotated position of relative domain over M .
Proof. This follows easily from the definition of t , conditions (C1)(C3), and
claims 7B.10, 7B.12, and 7B.13 above. We leave the precise details to the reader. Let us
only note that by Claims 7B.10 and 7B.13 the domain of t is equal to (W L ),
as required. #
Claim 7B.16. z(t ) is precisely equal to $y1+ +1 | + 1 < 1 + %.
Proof. Recall that z(t ) denotes the real part of t , given by Definition 4A.21. It is equal
to the sequence $t () | dom(t ) and is a relative successor%. Using Claim 7B.10
it is easy to check that the relative successors in the domain of t are precisely the
ordinals +1 where is smaller than and is either zero or a successor. Since t ( +1 )
is by definition equal to y for such we see that:

z(t ) = $y | is smaller than and is either zero or a successor%.

The sequence on the right-hand-side can be presented more uniformly as the sequence
$y1+ +1 | + 1 < 1 + %. #
We continue to work with a regular tot U of length on M, and with a U-sequence
$w , y | K U %. The definitions below are made with reference to these objects,
and also with reference to the annotated positions of Definition 7B.14.
Fix an ordinal < . Let r denote the set { | U }. This is simply the branch
of U leading to M . r is linearly ordered by U , and U < for , r.
276 7 Games which reach local cardinals

(Both properties follow from the definition of a tree order at the start of this section.)
Let be the order type of r. Let f : + 1 r {} be the unique order preserving
isomorphism of {} and r {}. Then < f ( ) U f ( ) for , .
For each < , f ( + 1) is a successor ordinal. (This follows from Claim 7B.1.)
Let E equal Ef ( +1)1 . By this we also mean that E =undefined if Ef ( +1)1 =
undefined. E is thus the object (extender, or undefined) used to determine Mf ( +1)
and the embedding jf ( ),f ( +1) : Mf ( ) Mf ( +1) via conditions (S) and (U) above.
Definition 7B.17. Let P be the sequence $Tf ( ) , bf ( ) , E , tf ( +1) | < %, where
, f , and E for < are the objects defined above.
We refer to P as the strand of U and $w , y | K U % leading to .
Definition 7B.17 breaks U and $w , y | K U % into strands, taken along the
branches of U . To each < lh(U) the definition associates a sequence P . P is
generated through the part of U and $w , y | K U % which corresponds to the
branch of U leading to . Note that this part has the format of a position of length in
the game G branch . The notation for positions in that game is explained in Remark 6A.2,
and it is easy to check that P fits the same format. Indeed, conditions (R2)(R4) and
(C1)(C3) above are almost enough to literally make P a legal position in G branch , as
they fit exactly with the rules (S1)(S4) and (L1)(L4) of Section 6A. (We will see this
precisely in Claims 7C.3 and 7C.4 below.) But two elements are still missing. The first
is a demand of triviality for the part of U which corresponds to phantom limits, in line
with the triviality of the phantom limit case in Section 6A. The second element missing
is a connection between the moves made in P and the rules for leaps in Section 6A.
This connection will be made in Section 7C, with early ends used in Section 7C (1) and
leaps handled in Section 7C(2).
Claim 7B.18. lh(P ) is: a successor if is a successor; a limit if is a limit; and zero
if = 0.
Proof. The length of P is equal to the order type of r = { | U }. By Claim 7B.1
it is a successor iff is a successor. The current claim follows from this, the fact that
{ | U 0} is empty, and the fact that { | U } is not empty (in fact it is cofinal
in ) for limit . #

7B (1) Extensions. Let U be a regular tot of length . By an extension of U to a tot of


length + 0.2 we mean a structure U+ which consists of the following objects:
a tree order U + on + 1, extending U ;
the models, iteration trees, extenders, and embeddings of U;
(if is a successor) an additional extender E1 in Q1 , or the additional
assignment E1 =undefined;
an additional model M , determined subject to the appropriate condition (S), (U),
or (L) in the definition of a tot; and
7C Basic step 277

additional embeddings j, : M M for U , again determined subject to


the appropriate condition (S), (U), or (L) in the definition of a tot.

If is a successor and E1 =undefined then the U + -predecessor of must be equal


to 1, in line with condition (S) in the definition of a tot.
U+ is regular if U is regular and if in addition, when is a successor, E1 (if defined)
is + 1-strong in Q1 . The last clause simply places on the additional extender E1
the same requirements that are placed on prior extenders through condition (R4) in the
definition of regularity for a tot of length .

Remark 7B.19. If is a limit then an extension of U to a tot of length + 0.2 is fully


determined by the set { | U + }. This set is a cofinal branch through U , and any
such branch gives rise to an extension of U.
If is a successor then an extension of U to a tot of length +0.2 is fully determined
by the value of E1 and by the U + -predecessor of . (If E1 =undefined then
even this information already involves a redundancy, since the U + -predecessor of
must equal to 1.)

An extension of U to a regular tot of length + 0.2 allows extending many of the


claims and definitions made above for < , and applying them also to = . We
leave it to the reader to check precisely which claims and definitions extend in this way.
But let us note that Definitions 7B.14 and 7B.17 are among them. So an extension of
U to a regular tot U+ of length + 0.2 allows talking about t and P . t is in fact
independent of the particular extension used. But P depends on U+ . We say that P
is defined with reference to U+ to emphasize this dependence.

7C Basic step
Let M be a transitive model of ZFC . Let  be an iteration strategy for M. Let be
an ordinal and let U be a regular tot of length + 1 on M, consistent with the iteration
strategy . Let $w, y% = $w , y | K U % be a U-sequence. We work with these
fixed objects, referring freely the terminology, notation, and definitions of Section 7B.
Recall most importantly that Definition 7B.17 breaks U and $w , y | K U % into
strands along the branches of U . Each of these strands has the format of a position
branch . We work here under the assumption that each of these strands is in fact a
in G
branch , and played according to a fixed strategy for I. We show how
legal position in G
to extend U so as to obtain the same properties for the additional strand generated by
the extension.
Let us be more precise. Let be a Woodin limit of Woodin cardinals in M, and
suppose that is countable in V. Let Y be a -name in M. Suppose that:

branch (M, , )(Y ).


(A1) For each , P is a legal position in G
278 7 Games which reach local cardinals

We work under this assumption for the rest of the section. The assumption states that
branch . We will
each of the strands of U and $w , y | K U % is a legal position in G
add more assumptions later, to the effect that each strand is played according to a fixed
strategy for I.
Claim 7C.1. For each :
(1) the outcome of P is precisely equal to $M , j0, , t %; and
(2) the history of P is precisely equal to $M , j, , t | U UEQ %.
We refer the reader to Remark 6A.3 for the definitions of outcome and history.
Proof. This is easy to prove directly from the definition of P , using the similarities
between: (a) the definitions in Section 6A of the models and embeddings created
branch ; and (b) the way conditions (S), (U), and (L) in Section 7B
through a position in G
determine the models and embeddings along the branch { | UEQ }. #

Claim 7C.2. Suppose that P is non-terminal in G branch (M, , )(Y ). Let = lh(P ).

Then mega-round of Gbranch (M, , )(Y ) following P is played subject to the rules
of:
(1) the phantom limit case in Section 6A if is a phantom limit in U;
(2) the standard limit case in Section 6A if is a standard limit in U; and
(3) the successor case in Section 6A if is zero or a successor.
Proof. By Claim 7B.18, is a limit iff is a limit. Condition (3) follows immediately
from this. Conditions (1) and (2) follow using the fact that the model and annotated
position in the outcome of P are precisely M and t , and using the similarities between:
(a) the distinction between the phantom limit case and the standard limit case in the
Definition of G branch in Section 6A; and (b) the distinction between phantom limits and
standard limits in Section 7B. #
Recall that the phantom limit case in Section 6A contains no moves at all. So the
situation of phantom limits in U is rather trivial. We neglect it in the discussion below.
Claim 7C.3. Let be a successor or zero. Let = lh(P ). Suppose that P
branch (M, , )(Y ). Then the objects w , y , T , and b (given
is non-terminal in G
by the fixed U, w,
 and y) are legal moves in mega-round of G branch (M, , )(Y )
following P .
Proof. We have to show that w , y , T , and b satisfy the requirements of rules
(S1)(S4) in Section 6A, for mega-round following P . The starting point for this
mega-round is the outcome of P , which we know is equal to $M , j0, , t %. So the
requirements of rules (S1)(S4) translate in our context to the following conditions:
(1) w is a witness for rdm(t );
7C Basic step 279

(2) y is a real;
(3) T is a length iteration tree on M , with critical points above rdm(t ) and using
only extenders which are countable in V; and
(4) b is a cofinal branch through T .
Checking that these conditions hold true is a simple matter of matching them with the
properties of U, w,
 and y listed in Section 7B: condition (C1) in that section implies that
w is a witness for rdm(t ); condition (C2) states that y is a real; condition (R2) implies
that T uses only extenders with critical points above rdm(t ); and condition (R3) states
that T uses only extenders which are countable in V. #
Claim 7C.4. Let be a standard limit in U. Let = lh(P ). Suppose that P
is non-terminal in G branch (M, , )(Y ). Then the objects w , T , b , and y (given
by the fixed U, w, and y) are legal moves for rules (L1)(L4) in mega-round of
branch (M, , )(Y ) following P .
G
Proof. This is similar to the previous claim. Let us only note that condition (C3) in
Section 7B matches the requirement of rule (L4) in Section 6A. #
Suppose now that player I has a winning strategy in G branch (M, , )(Y ). Fix

such a strategy, branch , for the rest of the section. Given a non-terminal position P
branch (M, , )(Y ) let 
in G branch [P ] denote the restriction of 
branch to mega-round
= lh(P ) in G branch (M, , )(Y ) following P .
Equipped with  branch we add the following conditions to the standing assumptions
of the section:
branch (M, , )(Y ) and played according
(A2) For each , P is non-terminal in G

to branch .
(A3) Let be a successor or zero. Then w , y , T , and b are consistent with
branch [P ].

(A4) Let be a standard limit in U. Then w , T , b , and y are consistent with
branch [P ].

(A5) itself is not a phantom limit in U.
Assumption (A5) is simply part of the neglect of the trivial phantom limits. Assump-
tions (A2)(A4) are more serious. They say that each strand of U and $w , y | K U %
branch (M, , )(Y ) which is played according to Is winning strategy
is a position of G

branch , and moreover the moves at the endpoint of the strand are also consistent with
branch .

We now work to extend U to a tot U+ of length + 1.2, in a way that secures legality
and consistency with  branch also for the resulting new strand, namely the strand leading
to + 1. We do it in such a way that the new strand is not terminal through any of
the snags in Section 6A. Since is a Woodin limit of Woodin cardinals, a position
280 7 Games which reach local cardinals

of G branch (M, , )(Y ) which is terminal but not through a snag can only occur after a
phantom limit. is not a phantom limit, so altogether the extension we aim for is such
that the new strand is legal, consistent with  branch , and non-terminal. In other words
we aim to construct an extension U of U which secures condition (A2) for + 1.
+

The final model in U is the model Q . Working over that model let t = t , w , y .
t is an annotated position of relative domain k ( ) + 1 over Q .

Claim 7C.5. Every strict initial segment of t is Q -clear.

Proof. t is the largest strict initial segment of t . So it suffices to show that t is


Q -clear.
By assumption (A1), P is a legal position in G branch (M, , )(Y ). By Claim 7C.1,
the outcome of P is equal to $M , j0, , t %. Using condition (1) in Remark 6A.4 it
follows that t is M -clear. Using now the agreement between M and Q implied by
condition (R2) in Section 7B it follows that t is Q -clear. #

Claim 7C.6. t is either obstruction free over Q , or else it is I-acceptably obstructed


over Q .

Proof. The claim hinges on the fact that annotated positions which are obstructed but
not I-acceptably obstructed always result in a loss for I in G branch . Let us be more
precise.
Suppose first that is a successor or zero. Assumptions (A2) and (A3) imply that
the position in G branch (M, , )(Y ) which consists of P followed by the moves w ,
branch . So this position cannot be
y , T , and b , is according to Is winning strategy 
lost by I. In particular it cannot be lost by I through condition (I2) in Section 6A. The
claim follows from this.
The proof when is a standard limit is similar, using assumptions (A2) and (A4)
above, and condition (I6) in Section 6A. #

Remember that our goal is to extend U to a regular tot U+ of length + 1.2, in a way
that makes the strand P+1 legal in G branch (M, , )(Y ), non-terminal, and consistent

with branch . We divide the construction into two cases, depending on whether t is
obstruction free over Q , or I-acceptably obstructed. By the last claim these two cases
exhaust all possibilities.

7C (1) Obstruction free. Suppose first that t is obstruction free over Q . Let U+ be
the extension of U, to a regular tot of length + 1.2, determined by the assignment
E =undefined. This assignment determines the extension completely, see Remark
7B.19. The extension allows talking about t+1 and P+1 .

Lemma 7C.7 (assuming that t is obstruction free over Q ). Let P+1 be defined with
reference to the extension U+ of U determined by the assignment E =undefined.
Then P+1 is legal in G branch (M, , )(Y ), non-terminal, and played according to
branch .

7C Basic step 281

Proof. Suppose first that is a successor or zero. Let = lh(P ). Using the fact
that is the U + -predecessor of + 1 it follows, directly from Definition 7B.17, that
P+1 = P , T , b , E , t+1 . E is equal to undefined and t+1 is simply equal to
t , w , y . So P+1 is simply the position given by P followed by the moves w ,
y , T , and b for mega-round . Assumption (A1) and Claim 7C.3 imply that P+1 is a
legal position in G branch (M, , )(Y ). Assumptions (A2) and (A3) for = combine to
state that P+1 is according to  branch . P+1 cannot be terminal in G branch (M, , )(Y )
through the payoff condition (P1) in Section 6A. This is because is a Woodin limit
of Woodin cardinals in M and +1 is a relative successor in M+1 , so +1 couldnt
possibly be equal to j0,+1 (). P+1 cannot be terminal through condition (I2) in
Section 6A since t+1 = t is not obstructed over M+1 = Q . Finally P+1 cannot
be terminal through condition (I1) in Section 6A since Q , being a model on a tot
consistent with the iteration strategy , is wellfounded.
The proof when is a standard limit is similar. P+1 is the position given by P
followed by the moves w , T , b , and y for rules (L1)(L4) in Section 6A, followed
by an early end rather than a leap. (Formally this early end is indicated by the fact
that E =undefined.) The fact that P+1 is a legal position in G branch (M, , )(Y )

and played according to branch follows as before, only using Claim 7C.4 instead of
Claim 7C.3, and assumption (A4) instead of assumption (A3). P+1 cannot be terminal
in G branch (M, , )(Y ) through the payoff condition (P1) in Section 6A. This is because
is a Woodin limit of Woodin cardinals in M and +1 is a relative limit in M+1 , so +1
couldnt possibly be equal to j0,+1 (). P+1 cannot be terminal through condition (I6)
in Section 6A since t+1 = t is not obstructed over M+1 = Q , and it cannot be
terminal through condition (I5) since Q is wellfounded. #

7C (2) I-acceptably obstructed. Suppose now that t is I-acceptably obstructed over


Q . Let $E,  % be a I-acceptable obstruction for t over Q .
Following the notation in Section 7B let +1 = k ( ). Using the fact that $E,  %
is an obstruction for t , and copying from the conditions in Definition 4B.26, we see
that:

(i) E is +1 + 1-strong in Q ; and

(ii) crit(E) is smaller than +1 , and is not Woodin in Q .

It follows from conditions (i) and (ii) that crit(E) is a limit of Woodin cardinals in Q ,
but not itself Woodin. By Claim 7B.6 there must therefore exist some so that:

(iii) is a standard limit; and

(iv) crit(E) is equal to .

Q and M agree past by Claim 7B.5. So the extender E can be applied to the
model M . Let U+ be the extension of U which makes this application. In other words
let U+ be the extension of U, to a tot of length + 1.2, determined by the assignments:
282 7 Games which reach local cardinals

(a) E = E; and
(b) the U + -predecessor of + 1 is .
These assignments completely determine U+ , see Remark 7B.19.
Claim 7C.8. U+ is regular.
Proof. Immediate from the strength of E given by condition (i). #
Claim 7C.9. M+1 = Ult(M , E ) is wellfounded.
Proof. Immediate from the fact that U is consistent with . #
Claim 7C.10. Let s be equal to t w , y over Q . Then t extends s (perhaps
not strictly).
Proof. If < then s is simply equal to t +1 , and t extends it strictly. If =
then s and t are equal by their definitions. #

Let = lh(P ). By Claim 7C.2 mega-round of G branch (M, , )(Y ) following


P is played subject to the rules of the standard limit case in Section 6A. By Claim 7C.4,
w , T , b , and y are legal moves for rules (L1)(L4) in that mega-round.
Claim 7C.11. Let P denote the position in G branch (M, , )(Y ) which consists of P
followed by the moves w , T , b , and y for rules (L1)(L4) in mega-round . Then
P is:
branch ; and
(1) played according to 
branch (M, , )(Y ).
(2) non-terminal in G
Proof. Assumptions (A2) and (A4) for = combine to state that P is played
according to  branch . This establishes condition (1) of the claim. Since  branch is a
winning strategy for I it also establishes that P is not lost by player I. So P cannot

be terminal through condition (I6) in Section 6A. P cannot be terminal through


condition (I5) in Section 6A since Q , being a model along a tot consistent with the
iteration strategy , is wellfounded. #
Claim 7C.12. Q &+1 + is countable in V.
Proof. Assumption (A2) states that P , whose outcome equals $M , j0, , t %, is non-
terminal in G branch (M, , )(Y ). It follows that rdm(t ) is smaller than j0, ( ). This
in turn implies that , which is equal to the first Woodin cardinal of M above rdm(t ),
is smaller than j0, (). So +1 , which is equal to k ( ), is smaller than (k j0, )( ).
Since is inaccessible in M it follows that the cardinality of Q &+1 + in Q is
smaller than (k j0, )(). So it suffices to prove that (k j0, )( ) is countable in V.
By assumption is countable in V. (This is part of the initial settings of Section 7C.)
By condition (2) in Remark 6A.4 applied to P , j0, preserves countability. By condi-
tion (R3) in Section 7B and Fact 3 in Appendix A, k preserves countability. Combining
all these assertions it follows that (k j0, )() is countable in V. #
7C Basic step 283

Lemma 7C.13 (assuming that t is I-acceptably obstructed over Q ). Let P+1 be


defined with reference to the extension U+ of U determined by conditions (a) and (b)
branch (M, , )(Y ), non-terminal, and played
above. Then P+1 is a legal position in G

according to branch .
Proof. Recall that is the U + -predecessor of + 1, and that = lh(P ). is a
standard limit in U by condition (iii) above.
Definition 7B.17 is such that:
(v) P+1 = P , T , b , E , t+1 .
By Claim 7C.2, mega-round of G branch (M, , )(Y ) following P is played subject to
the rules of the standard limit case in Section 6A. It follows from this, from condition (v),
and from the fact that E is not undefined, that P+1 is the position which consists
of:
(K1) P ; followed by
(K2) the moves w , T , b , and y (given by the fixed U, w,
 and y) for rules (L1)(L4)
in mega-round ; followed by
(K3) a leap with the moves E , +1 , and t+1 for rules (L5) and (L6).
The part of P+1 described in conditions (K1) and (K2) is simply the position P of
Claim 7C.11. We already know that this part is legal, played according to  branch ,
and non-terminal. So the key to the lemma is verifying that the continuing moves
described in condition (K3) are legal and non-terminal. There is no need to worry about
branch for these moves since they are made by player II.
consistency with 

j ,+1

kkk &
Sk
E
/ M Sk
k SSSS k / Q _ _ _ _ _ _ Q M+1

T

Diagram 7.3. The models and embeddings connected with mega-round in P+1 .

The legality of the moves described in condition (K3) follows through an application
of Lemma 6A.5, with: Q = Q , E = E , = , and t = t . There are some
indexical difficulties in connecting Lemma 6A.5 to the current situation, since mega-
round in P+1 is spread over several stages of U+ , and chances are that none of
these stages is . We try to ease these difficulties by illustrating in Diagram 7.3 all the
models and embeddings related to mega-round in P+1 . The Diagram is identical to
Diagram 6.2, except for the indexing. Mega-round in P+1 starts from the outcome
of P , which is equal to $M , j0, , t %; continues with the moves w , T , b , and y ;
and ends with the moves E , +1 , and t+1 .
284 7 Games which reach local cardinals

Other than the indexical difficulties its easy to check that Lemma 6A.5 applies
in this context. Condition (1) in the lemma follows from Claim 7C.12; condition (2)
in the lemma follows from Claim 7C.10; condition (3) in the lemma follows from
Claim 7C.5; condition (4) in the lemma follows from condition (iv) above, at the start
of Section 7C (2); and condition (5) in the lemma follows from Claim 7C.9.
So far we verified that the moves described in condition (K3) are legal in
branch (M, , )(Y ). These moves could only be terminal through condition (I7) in
G
branch ends with a loss to II if M+1 is illfounded.
Section 6A. The condition states that G
But M+1 is wellfounded by Claim 7C.9. So the moves described in condition (K3)
are non-terminal. #

7C (3) Summary. We are working in Section 7C with the following objects:


a transitive model M of ZFC ;
an iteration strategy  for M;
a Woodin limit of Woodin cardinals and a -name Y in M;
branch (M, , )(Y );
branch for I in G
a winning strategy 
a regular tot U of successor length + 1 on M, consistent with ; and
a U-sequence $w,
 y%.
Definition 7B.17 breaks U and $w,  y% into strands along the branches of U . We assume
branch (M, , )(Y ), non-terminal,
here that each of these strands is a legal position in G
and played according to  branch . We assume further that the moves at the endpoint
of each strand are also consistent with  branch . Our work shows how to extend U in
a way that makes the resulting new strand, namely the strand leading to + 1, legal,
non-terminal, and consistent with  branch . Precisely we obtain:

Lemma 7C.14 (under assumptions (A1)(A5) listed earlier in Section 7C). There exists
an extension of U to a regular tot U+ of length + 1.2, so that defining P+1 with
reference to U+ yields the following conditions:
branch (M, , )(Y );
(1) P+1 is a legal position in G
branch (M, , )(Y ); and
(2) P+1 is non-terminal in G
branch .
(3) P+1 is played according to  #
This follows from Lemma 7C.7 if t is obstruction free over Q , and from Lemma
7C.13 if t is I-acceptably obstructed over Q . By Claim 7C.6 there are no other
cases. In the first case U+ is defined so that the U + -predecessor of + 1 is , and
P+1 continues P . The second case is more involved. U+ is defined to that the U + -
predecessor of +1 is the ordinal corresponding to the critical point of the I-acceptable
obstruction given by the case assumption. P+1 in this case continues P , and does so
with a leap.
7D Construction 285

7D Construction
Let M be a transitive model of ZFC . Let  be an iteration strategy for M. Let
be a Woodin limit of Woodin cardinals in M. Let A M be a W -name for a set
of sequences of reals of length . W as usual denotes Woodins algebra defined in
Section 4B. The objects above correspond to the settings in Section 7A. We work with
them fixed for the rest of the section.
Definition 7D.1. For expository simplicity fix some filter G which is W -generic/M.
Define Y M to be the canonical W -name for the set of -sequences t M[G] so
that z(t) A[G].
z(t) in Definition 7D.1 is the real part of t, given by Definition 4A.21. Recall that it
is equal to $t () | dom(t) and is a relative successor%. This is a sequence of reals
numbers. Using the fact that is a Woodin limit of Woodin cardinals it is easy to check
that there are relative successors below . So z(t) is a sequence (of reals) of length .
The clause z(t) A[G] in Definition 7D.1 is therefore meaningfulit is not trivially
false.
Let ini be the formula of Definition 5G.2. The next lemma begins the proof of the
main result described in Section 7A.
Lemma 7D.2. Suppose that:
(1) is countable in V; and
(2) ini [, Y ] holds in M.
local (M, , , A).
Then (in V) player I has a winning strategy for G
local (M, , , A). We
Proof. Fix an imaginary opponent willing to play for II in G
describe how to play for I, and win. The description as usual takes the form of a
construction. We work to construct:
(A) an ordinal ;
(B) a regular tot U+ of length + 0.2 on M; and
(C) a U-sequence $w , y | K U %, where U denotes U+  .
local (M, , , A) through the assignment:
We associate to these objects a position z in G
(D) z = y1+ +1 for each so that + 1 < 1 + .
In other words z is simply the sequence:

$y | < is zero or a successor ordinal%.

This is a sequence of reals by condition (C2) in Section 7B. We shall verify at the end
local (M, , , A) which
of the construction that this sequence of reals forms a run of G
286 7 Games which reach local cardinals

is won by player I. To verify victory by I we shall need an iteration tree U witnessing


condition (P) in Section 7A. Let us already here say that the tree we intend to use is
simply the merge of U+ .
The only help we get from the imaginary opponent during the construction is in the
creation of z. In other words the imaginary opponent only participates in the creation
of y , and only for which is either zero or a successor ordinal. The production of all
the other objects involved in conditions (B) and (C) falls squarely on us. We handle
it using the tools of Chapter 6. Corollary 6G.2 and Remark 6G.3 tell us that (under
the assumptions of the current lemma) player I has a winning strategy in the game
branch (M, , )(Y ). Fix, for the rest of the proof, a strategy 
G branch which witnesses
this. We intend to heave much of the burden of the construction on  branch .
We construct subject to the following conditions:
branch (M, , )(Y ), non-terminal, and
(1) for each < , P is a legal position in G

played according to branch ;

(2) U+ is consistent with the iteration strategy ;

(3) if < is a successor or zero then w , y , T , and b are consistent with


branch [P ];


(4) if < is a standard limit in U then w , T , b , and y are consistent with


branch [P ];


(5) if < is either zero, a successor, or a standard limit in U, then U + 1.2 is


obtained from U + 1 through an application of Lemma 7C.14; and

(6) if < is a phantom limit in U then T is the iteration tree which consists
entirely of padding, b is the unique branch through it, and E =undefined.

Conditions (2)(6) in fact completely determine each step of the construction:


If < is zero or a successor then w , y , T , and b are determined through
condition (3) by a collaborative effort which involves  branch [P ], the imaginary op-
ponent, and the iteration strategy . (The imaginary opponent participates in the cre-
ation of y , and  chooses b . Everything else is done by  branch .) E and the
U + -predecessor of + 1 are subsequently determined through condition (5) by an ap-
plication of Lemma 7C.14.
If < is a standard limit in U then w , T , b , and y are determined through
branch [P ] and the iteration strat-
condition (4) by a collaborative effort which involves 
egy . ( chooses b . Everything else is done by  branch .) E and the U + -predecessor
of + 1 are again determined through condition (5) by an application of Lemma 7C.14.
If < is a phantom limit in U then T , b , and E are precisely specified by
condition (6). The U + -predecessor of + 1 in this case is . There is no need to define
w and y since does not belong to K U .
7D Construction 287

Finally, if is a limit then the branch { | U } which is used to give rise to


the direct limit model M is picked by  subject to condition (2).
branch [P ]
Condition (1) is an inductive condition, needed to allow the references to 
in conditions (3) and (4). Notice that it holds trivially for = 0. For as long as the
condition continues to hold we can continue the construction. If ever it fails we must
end. In other words the construction ends at the first ordinal so that condition (1)
fails for = , if such an ordinal is ever reached.
Claim 7D.3. Let be a limit ordinal smaller than 1V . Suppose that condition (1) holds
for all < . Then the condition holds also for .

Proof. For a limit ordinal , P is equal to the increasing union U P . By assump-
tion each Pis legal in G branch (M, , )(Y ), non-terminal, and played according to
branch . So
 P is certainly branch (M, , )(Y ) and played according to
legal in G
U 
branch . It remains to check that it is non-terminal. Now
 U P , being a position of
limit length, can only be terminal through one of the conditions (I3) and (I4) in Sec-
tion 6A. But neither condition holds here. Condition (I3) does not hold since M , being
a model on a tot consistent with the iteration strategy , is wellfounded. Condition (I4)
does not hold since by assumption is smaller than 1V , and therefore so is the order
type of { | U }. #

Claim 7D.4. Let be either zero, a successor ordinal, or a standard limit in U. Suppose
that condition (1) holds for . Then the condition holds also for + 1.

Proof. For as in the claim, U + 1.2 is obtained from U + 1 through condition (5)
of the construction, or in other words through an application of Lemma 7C.14. The
conclusion of that lemma precisely secures condition (1) for + 1. #

The last two claims imply that if the construction ends before 1V , it must be at a
successor ordinal which immediately follows a phantom limit.
Claim 7D.5. The construction ends before reaching 1V .

Proof. Suppose for contradiction that the construction reaches 1V . Let r = { | U


1V }. Each of the positions P , r, is legal in G branch (M, , )(Y ) and played

according to  branch . It follows that the increasing union
r P is a legal position in

Gbranch (M, , )(Y ), and played according to branch . Now r, being cofinal in 1V , has

branch (M, , )(Y ).
order type 1V . It follows that r P is a position of length 1V in G
It is therefore lost by player I through condition (I4) in Section 6A. But this contradicts
branch is a winning strategy for player I.
the fact that  #

Remark 7D.6. By Claim 7D.5 only countable iteration trees actually appear in the
construction. But to prove the claim we had to allow for the possibility of a tree of
length 1V + 1. So the claim uses the fact that  acts not only on countable iteration
trees but also on iteration trees of length 1V .
288 7 Games which reach local cardinals

This kind of phenomenon, where an iteration strategy which acts also on trees of
length 1V is needed to in fact rule out the case of trees of length 1V , is typical in the
study of large cardinals. The classic example is the comparison argument of countable
mice in inner model theory, where it is always shown at the end that the comparison
has countable length, but where the possibility of an iteration tree of length 1 must be
allowed before it can be ruled out.
Let be the first ordinal for which condition (1) fails. We know from the last three
claims that such an ordinal exists, that it is smaller than 1V , and that it is a successor
ordinal immediately following a phantom limit in U.
Constructing up to subject to conditions (2)(6) we precisely obtain the objects
listed in conditions (B) and (C) above. Through the assignment of condition (D) we
obtain also a position
z = $y1+ +1 | + 1 < 1 + %
local (M, , , A). We now work through a series of claims to verify that z is a
in G
complete run of G local (M, , , A), and won by player I.
branch (M, , )(Y ), and won by player I
Claim 7D.7. P is a terminal position in G
through condition (P2) in Section 6A.

Proof. Let = 1. is then a phantom limit in U. From condition (6) it follows


that E =undefined. The U + -predecessor of = + 1 is therefore equal to , and
P extends P (by one mega-round).
Let = lh(P ). Since is a phantom limit in U, Claim 7C.2 implies that
mega-round of G branch (M, , )(Y ) following P is played subject to the trivial
rules of the phantom limit case in Section 6A. So P is simply the position which
extends P with one, trivial phantom mega-round. Such an extension is legal in
branch (M, , )(Y ) following P , and trivially consistent with 
G branch [P ]. So P

is legal in Gbranch (M, , )(Y ) and played according to branch . Since condition (1)
nonetheless fails for , P must be terminal in G branch (M, , )(Y ). The only way
a position ending a phantom mega-round could be terminal is through condition (P2)
in Section 6A. P must be won by player I through that condition, since it is played
according to Is winning strategy  branch . #

Following the notation of Section 7B let M and j, : M M denote the


models and embeddings of the tot U+ . U+ is a tot of length + 0.2 on M. It leads
to a final model M , and to a final iteration embedding j0, : M M . Through
Definition 7B.14, or more precisely the extension of that definition to U+ , we may talk
about t , an annotated position over M .
Claim 7D.8. There exists some G so that: G is j0, (W )-generic/M ; and t belongs
to j0, (Y )[G].

Proof. This is simply the payoff in condition (P2) of Section 6A, translated to the
current context using the fact that the outcome of P is equal to $M , j0, , t %. #
7E The main theorem 289

Fix from now on some G which witnesses Claim 7D.8.


Claim 7D.9. z belongs to j0, (A)[G].
Proof. Recall that z is equal to $y1+ +1 | + 1 < 1 + %. By Claim 7B.16, z is
precisely equal to z(t ). The current claim now follows from the fact that t belongs
to j0, (Y )[G], and from the relationship between Y and A given by Definition 7D.1.
#
Claim 7D.9 completes the proof of Lemma 7D.2. It shows that z, which was
produced with the collaboration of the imaginary opponent, is a complete run of
local (M, , , A) and won by player I. The objects which witness Is victory are the iter-
G
ation tree merge(U+ ) and the generic filter G given by Claim 7D.8. # (Lemma 7D.2)
Remark 7D.10. For the sake of clarity let us take note of the use of the assumptions
of Lemma 7D.2 during its proof. The assumptions were used in the appeal to Corol-
lary 6G.2, which (combined with Remark 6G.3) provided the most crucial element of
branch .
the construction, namely the strategy 

7E The main theorem


Let M be a transitive model of ZFC . Let  be an iteration strategy for M. Let be a
Woodin limit of Woodin cardinals in M. Let A and B be W names, each naming a set
of sequences of reals of length . The following theorem, previewed in Section 7A, is
local
the main result of the current chapter. It is phrased with reference to the games G
and H local defined in Section 7A. We already did most of the work needed for its proof.
All thats left now is to assemble the various parts.
Theorem 7E.1. Suppose that is countable in V. Then at least one of the following
three cases holds:
local (M, , , A);
(1) player I has a winning strategy in G
local (M, , , B); or
(2) player II has a winning strategy in H
(3) there exists some G V which is W -generic/M, and there exists some sequence
of reals $z | < % M[G], so that $z | < % belongs to neither A[G] nor
B[G].
Moreover M can distinguish this. More precisely, there are formulae and so that: if
M |= [, A] then case (1) holds; if M |= [, B] then case (2) holds; and otherwise
case (3) holds.
Proof. For expository simplicity fix some G which is W -generic/M. Let Y M
be the canonical name for the set of -sequences t M[G] so that z(t) A[G]. Let
Z M be the canonical name for the set of -sequences t M[G] so that z(t) B[G].
Let ini and ini be the two formulae of Definition 5G.2.
290 7 Games which reach local cardinals

Case 1. If ini [, Y ] holds in M. In this case an application of Lemma 7D.2 shows that
player I has a winning strategy in G local (M, , , A). So case (1) of the theorem holds.
#
Case 2. If ini [, Z] holds in M. In this case an argument which precisely mirrors the
development which led to case 1 (including the entire development of Chapter 6 and
Section 7D) shows that player II has a winning strategy in H local (M, , , B). So case
(2) of the theorem holds. #
Case 3. If both ini [, Y ] and ini [, Z] fail in M. In this case an application of
Corollary 5G.3 produces, in some generic extension of M, a -sequence t and some
filter G so that:
(i) G is W -generic/M;
(ii) t belongs to M[G]; and
(iii) t belongs to neither Y [G] nor Z[G].
Let D be the collection of maximal anti-chains of W in M. Each anti-chain in D
is a subset of M& which belongs to M, and has size strictly less than in M by
Lemma 4B.19. So each anti-chain in D is an element of M&. It follows from this
and from the inaccessibility of in M that the cardinality of D in M is at most .
Theorem 7E.1 assumes that is countable in V. So:
(iv) D is countable in V.
Using the last condition one can reflect the existence of objects t and G which satisfy
conditions (i)(iii), from an arbitrary generic extension of M, to V. So we may without
loss of generality assume that t and G belong to V.
Since is an inaccessible limit of Woodin cardinals, the set of relative successors
below has order type precisely . Let $ | < % enumerate this set in increasing
order. Following the notation of Definition 4A.21 let z = t ( ) for each < .
$z | < % is a sequence of reals. It is precisely equal to the real part of t, namely to
z(t).
From condition (ii) it follows that:
(v) $z | < % belongs to M[G].
From condition (iii) and the use of z(t), A, and B in the definition of Y and Z above it
follows that:
(vi) $z | < % belongs to neither A[G] nor B[G].
So case (3) of Theorem 7E.1 holds. #
The three cases above complete the proof of Theorem 7E.1. The formulae and
which witness the final part of the theorem are defined in the natural way, using the
formulae ini and ini , and using the definitions of Y and Z at the start of the proof.
# (Theorem 7E.1)
7E The main theorem 291

It is clear from the construction in Chapter 6, its use in Sections 7B and 7C, and
ultimately the proof of Lemma 7D.2, that the winning strategies in cases 1 and 2
are produced definably in  and a given real coding M& . In fact the strategies
are definable from the real and the restriction of  to just countable iteration trees.
These are the only trees that come up in construction; trees of length 1 only come
up through the proof of Claim 7D.5. Certainly then the winning strategies belong to
L(R, {countable itereation trees}).
This can be refined further. For a model P let  P be the restriction of  to iteration
trees which belong to P and are countable in P . Given a strategy in a long game on
natural numbers, let P be the restriction of to positions which belong to P and are
countable in P . The strategy is locally definable from  above a real u if it satisfies
the following property: let P be any model of ZFC so that u belongs to P , and so that
 P belongs to P ; then P belongs to P . The following remark is then clear from the
constructions which go into the proof of Theorem 7E.1:

Remark 7E.2. Condition (1) in Theorem 7E.1 can be strengthened to say that the
winning strategy is locally definable from  over a given real which codes M&; and
similarly with condition (2).

Let us also note for the record that  need only apply to iteration trees on M&. This
is clear from the constructions, and can also be derived abstractly from Theorem 7E.1
by assuming, without loss of generality, that M has no extenders above .

Exercise 7E.3. For each Woodin cardinal M let otrs() denote the order type of the
set { ( + 1) W | is a relative successor}. Show that Theorem 7E.1 holds with
condition (3) strengthened to include also the following statement: For each Woodin
cardinal < of M, the sequence $z | < otrs()% belongs to a generic extension of
M via the poset col(, ), if is not a Woodin limit of Woodin cardinals in M, and via
the poset W if it is.

Exercise 7E.4. This exercise reaches a parallel of Theorem 7E.1 for the case that
is a Woodin cardinal in M but not a Woodin limit of Woodin cardinals. Fix such
a . Suppose that M& + 1 is countable in V. Let A, B M be names for sets of
sequences of reals of length otrs() (see Exercise 7E.3). Define G local (M, , , A) and
H local (M, , , A) following the specifications in Section 7A, only replacing the poset
W in the payoff conditions (P) and (Q) with the poset col(, ). Prove that at least
one of the following three cases holds:
local (M, , , A).
(1) Player I has a winning strategy in G

local (M, , , B).


(2) Player II has a winning strategy in H

(3) There exists some G V which is col(, )-generic/M, and there exists some
sequence of reals z = $z | < otrs()% M[G] which belongs to neither A[G]
nor B[G].
292 7 Games which reach local cardinals

Moreover, the sequence z is such that for each Woodin cardinal < of M, the
restriction zotrs() belongs to a generic extension of M via the poset col(, ),
if is not a Woodin limit of Woodin cardinals in M, and via the poset W if it is.
Show further that M can distinguish this. Precisely, show that there are formulae and
so that: if M |= [, A] then case (1) holds; if M |= [, B] then case (2) holds;
and otherwise case (3) holds.
Exercise 7E.5. Suppose that M has no extenders which overlap Woodin cardinals. (In
particular there are no Woodin limits of Woodin cardinals in M.) Repeat Exercise 7E.4
in this case, without using any of the results in Chapters 4, 5, and 6. (This will not lead
to any additional results, but you will gain some experience.) You will have to define,
and work with, your own pullback operation. Since M has no extenders overlapping
Woodin cardinals, the settings here are a great deal simpler than the general settings
in Chapter 4. For your pullback operation you will only need the successor case from
that chapter, and a substantially simplified version of the case of relative limits, more
in the spirit of Chapter 2 than Chapter 3, without any records and with no reference to
obstructions.

7F Determinacy
Theorem 7E.1 yields an array of determinacy results. Roughly speaking it allows
handling games which end at the first cardinal of some inner model built relative to the
play. We illustrate this by presenting a specific application, with L as the inner model.
Let C R<1 be given. Glocal (L, C) is played as follows: Players I and II alternate
playing natural numbers as in Diagram 7.4 to produce reals z . They continue until
reaching the first which is uncountable in L[z | < ]. At that point the game ends.
Player I wins iff $z | < % belongs to C.

I z0 (0) z0 (2) ...... z (0) z (2) ......


II z0 (1) ... z (1) ...

Diagram 7.4. The game Glocal (L, C).

L[z | < ]
The first which is uncountable in L[z | < ] must in fact be equal to 1 .
We therefore refer to the games Glocal (L, ) as games which end at 1 in L of the
L[z | < ]
play. We view 1 as a local cardinal, in that it is a cardinal of the inner
model L[z | < ]. The notation Glocal (L, C) comes from this view.
Technically runs of length 1 in Glocal (L, C) are won by player II. But even very mild
large cardinal assumptions imply that the first which is uncountable in L[z | < ]
is always countable in V. The large cardinals we work with here are more than enough
7F Determinacy 293

for this, as the next claim demonstrates. So in our context all runs of Glocal (L, C) are
countable.

Claim 7F.1. Suppose that there exists an iterable class model M with a cardinal so
that: (a) is a Woodin limit of Woodin cardinals in M; and (b) is countable in V.
Let $u | < 1V % be a sequence of reals of length 1V . Then there is some , strictly
smaller than 1V , so that is uncountable in L[u | < ].

Proof. Let A M be the canonical W -name for the set of all sequences of reals of
length . Let B name the empty set. Apply Theorem 7E.1 with these names. Neither
case (2) nor case (3) of the theorem can occur. So it must be that I has a winning strategy
in G local (M, , , A). Let be such a strategy. Play against by making the moves
z (2i + 1) = u (i) for II. must secure victory for I at some strictly smaller than 1V .
Using Claim 7A.2 it is easy to see that this is uncountable in L[u | < ]. #

We should note that the large cardinal assumption in the claim is in fact a huge
overkill. It is enough to assume that is a Woodin cardinal (rather than a Woodin limit
of Woodin cardinals). One can then obtain the conclusion of the claim using Woodins
standard extender algebra, where extender axioms are defined using all extenders, rather
than just ones which overlap Woodin cardinals.
Our intention next is to prove the determinacy of Glocal (L, C) for payoff sets C
which are  (<2 11 ) in the codes. Let us first be more precise on the way we code
runs.

Definition 7F.2. Let be a countable ordinal and let $z | < % be a sequence of


reals of length . By a code for $z | < % we mean a sequence $xn | x < % in R
so that:

(1) x0 is a wellordering of a subset of {0};

(2) the order type of x0 is precisely ; and

(3) for every n dom(x0 ), xn is equal to z where < is the order type of n in
the wellordering x0 .

Definition 7F.3. Let  be a pointclass. A set C R<1 is  in the codes just in case
that the set C of $xn | n < % R which code elements of C belongs to .

Theorem 7F.4. Suppose that there exists an iterable class model M with a cardinal
so that: (a) is a Woodin limit of Woodin cardinals in M; and (b) is countable in V.
Then the games Glocal (L, C) are determined for all C which are  (<2 11 ) in the
codes.

Remark 7F.5. The existence of M and satisfying the assumptions in Theorem 7F.4
follows from the existence, in V, of a sharp for a Woodin limit of Woodin cardinals, see
Theorem 16 in Appendix A.
294 7 Games which reach local cardinals

Proof of Theorem 7F.4. Fix M and satisfying the assumptions of the theorem. Let 
be an iteration strategy for M.
Fix C R<1 which is  (<2 11 ) in the codes. We work to show that
Glocal (L, C) is determined.
Let C be the set of $xn | n < % R which code elements of C. C then belongs
to  (<2 11 ). By Martin [24] there is a number k < and a 1 formula so that
for any x = $xn | n < %:
x C x ] |= [
L[ x , c0 , . . . , ck1 ] whenever c0 < < ck1
are Silver indiscernibles for x
Let (a, , c0 , . . . , ck1 ) be the formula a is a sequence of reals of length ; and
it is forced in col(, ) that there is a code x for a so that [ x , c0 , . . . , ck1 ] holds
in L[x ].
Let u0 < < uk1 be uniform Silver indiscernibles. For expository simplicity
fix some G which is W -generic over M. W as usual stands for Woodins extender
algebra defined in Section 4B.
Working in M[G] let A be the set of sequences $z | < % of reals so that
M[G] |= [$z | < %, , u0 , . . . , uk1 ]
where is the least ordinal which satisfies the condition:
() is uncountable in L[z | < ].
Note that such an ordinal must exist, since is uncountable in M[G]. Let A M be
the canonical W -name for A.
Let B be the set of sequences $z | < % M[G] of reals so that
M[G] |= [$z | < %, , u0 , . . . , uk1 ]
where is the least ordinal which satisfies the condition () above. Let B M be
the canonical W -name for B.
Claim 7F.6. Every sequence of reals $z | < % in M[G] belongs either to A[G] or
to B[G].
Proof. Immediate. A and B name complementary sets. #
Apply Theorem 7E.1 with the names A and B. The last claim rules out case (3) of
the theorem. The remaining possibilities are either: (1) player I has a winning strategy
local (M, , , A); or (2) player II has a winning strategy in H
in G local (M, , , B).

Case 1. If I has a winning strategy in G local (M, , , A). It is easy to see that this
strategy induces a winning strategy for I in Glocal (L, C). The key point is that any run
$z | < % of G local (M, , , A) which is won by I must, by Claim 7A.2, either reach
or pass a which is uncountable in L[z | < ]. In other words it must either equal
or extend a complete run of Glocal (L, C). The definition of A above is such that this
run of Glocal (L, C) must be won by player I. #
7F Determinacy 295

Case 2. If II has a winning strategy in H local (M, , , B). Then this strategy induces
a winning strategy for II in Glocal (L, C). Again the key point is the fact that any run
local (M, , , B) which is won by II must either equal or extend a complete run of
of H
Glocal (L, C). The definition of B above is such that this run of Glocal (L, C) must be
won by player II. #

We already saw, through a use of Theorem 7E.1, that one of the two cases above
must hold. So Glocal (L, C) is determined. # (Theorem 7F.4)

To progress beyond the games in Theorem 7F.4 it is convenient to phrase the ending
condition in terms of descriptive set theory.
Let $z | < % be a sequence of reals of length . let x = $xn | n < % code
this sequence in the sense of Definition 7F.2. In particular then x0 is a wellordering of
{0}, of order type precisely . Given a function f : let fx be the real a
defined by setting a(n) equal to the unique k so that the order type of k in x0 is precisely
f (n). fx then codes the function f , modulo the wellordering x0 .
Let  be a pointclass. We say that f : belongs to [z | < ] just in
case that, for every code x of $z | < %, the real fx belongs to the pointclass (x ).
We say that is countable in [z | < ] just in case that there exists a function f
from onto , which belongs to [z | < ].

Remark 7F.7 (assuming every real has a sharp). If  is the pointclass  (<2 11 )
then, by Martin [24], the reals which belong to  are precisely the reals in L, and
similarly for any x R the reals which belong to (x) are precisely the reals in L[x].
It follows in this case that is countable in [z | < ] if and only if it is countable
in L[z | < ].

Let C R<1 be given. Define Glocal (, C), the game ending at 1 in 
of the play, in the obvious way, taking the definition of Glocal (L, C) and changing
uncountable in L[z | < ] to not countable in [z | < ]. Notice then that
the games Glocal (, ) become stronger, meaning longer, as  increases.
Theorem 7F.4 states the determinacy of the games Glocal (, C) for the pointclass
 =  (<2 11 ) and sets C in . The proof of Theorem 7F.4 involves nothing more
than an application of Theorem 7E.1. That theorem in turn is general enough that it can
be used in the context of pointclasses larger than  (<2 11 ). More specifically it
can be applied whenever the given model M, and its  iterates, can correctly compute
membership in sets which belong to .
Let us be more precise. Let A be a set of reals. Let M be a model, let  be an
iteration strategy for M, let be a Woodin cardinal in M, and suppose that M& + 1 is
countable in V. Let A M be a col(, )-name for a set of reals. The name A captures
A over $M, % just in case that for every g V which is col(, )-generic over M,
A[g] is precisely equal to A M[g]. A fully captures A over $M, , % just in case that
for any (countability preserving) -iteration embedding j : M M , j (A) captures
A over $M , j ( )%. A pointclass  is captured over $M, , % just in case that every
296 7 Games which reach local cardinals

set of reals in  is fully captured by some name over $M, , %. These definitions are
all due to Woodin.
It is clear, using Shoenfield absoluteness, that the pointclass 12 is captured over
$M, , % whenever M is a (wellfounded) class model. The results of Martin [24] give
the same for the pointclass  (<2 11 ), and this really was our tool in deriving
Theorem 7F.4 from Theorem 7E.1. Thus working in general settings we obtain:
Exercise 7F.8. Let  be a pointclass, either -parameterized (see Moschovakis [26,
p. 36]) or the union of countably many -parameterized pointclasses. Suppose that
there is an iterable model M of ZFC , a cardinal in M, and an iteration strategy 
for M so that: (a) is a Woodin limit of Woodin cardinals in M; (b)  is captured
by $M, , %; and (c) M& + 1 is countable in V. Let C be  in the codes. Then
Glocal (, C) is determined.
Hint. Replace the names A and B in the proof of Theorem 7F.4 with names capturing
(the set of codes for) C and its complement. You will have to add an argument in case 1
local (M, , , A) reach 1 in  of the play, and similarly in case 2.
to show that runs of G
For this you will have to argue that for any -iteration embedding j : M M , any
sequence of reals $z | < % of length at most j ( ), and any function f :
which belongs to [z | < ], if $z | < % belongs to an extension of M by a
generic for j (W ), then f belongs to the same extension. Here too you will use the
fact that  is captured over $M, , %. #
The last exercise can be applied in specific settings by noticing that our determinacy
proofs generally show that iterable models with enough large cardinals capture point-
classes generated by game quantifiers. The exercise and the corollary below demonstrate
this in the context of the quantifiers associated to games of fixed countable length. Sim-
ilar results can be proved with the quantifiers associated to the games in Chapter 3, and
also with the quantifiers associated to the games here, in the current chapter.
Exercise 7F.9. Let 1 be a countable ordinal. Let M be an iterable class model and
let < be such that: (a) M sees that is countable; (b) is a Woodin cardinal in M; (c)
there are 1 + Woodin cardinals of M in the interval (, ); and (d) M& is countable
in V. Let  be an iteration strategy for M. Then the pointclass  (<2 11 ) is
captured over $M, , %.
Hint. Go over the results in Chapter 2 and check that for each g which is col(, )
generic over M and each real x M[g], M[g] can (uniformly) identify which player
wins a game of length with payoff in (<2 11 )(x). This will show that sets in
 (<2 11 ) are captured over $M, %. Then use the uniformity of the argument
to obtain full capturing over $M, , %. #
Corollary 7F.10. Let 1 be a countable ordinal. Suppose that there exists an
iterable class model M and < in M so that: (a) is a Woodin limit of Woodin
cardinals in M; (b) there are 1 + Woodin cardinals of M in the interval (, ); and
(c) M& is countable in V. Let  be the pointclass  (<2 11 ), and let C be 
in the codes. Then Glocal (, C) is determined. #
7F Determinacy 297

Exercise 7F.11. Formulate and prove a result similar to the last corollary, for cont
instead of  .

It is natural to ask whether the results above, on capturing and subsequently on


determinacy, can be extended to universally Baire sets. Exercise 2E.5 shows how
universally Baire sets can be captured over models which embed into a rank initial
segment of V. The argument there is sufficiently uniform that it would produce full
capturing over a model and an iteration strategy, provided that the iteration strategy
picked realizable branches (see the hint for Exercise 2E.6), that is branches whose
direct limits can be embedded into a rank initial segment of V. The weak iterability
necessary for the results in Chapter 2 was obtained through Theorem 12 in Appendix A,
and the theorem produced realizable branches through the relevant iteration trees. This
allowed us to bring the determinacy proved in Chapter 2 to bear on universally Baire
sets. The mild iterability necessary for the results in Chapter 3 was obtained through
Claim 14 in Appendix A, which similarly produced realizable branches through the
relevant trees. Again this allowed proving determinacy in the case of universally Baire
payoff. The results here on the other hand require full iterability, obtained through
Theorem 16 in Appendix A. This theorem is proved indirectly, using fine structure
among other things. It does not show that models which embed into initial segments
of V are iterable, let alone iterable by a strategy which picks realizable branches. It
therefore does not lead to applications of our techniques to universally Baire sets.
Still it is possible, assuming sufficient large cardinals, to obtain fully iterable models
which capture a given -universally Baire set and have a Woodin limit of Woodin
cardinals. The proof of this is due to Woodin, and combines the iterability proof in
Neeman [31] (the source of Theorem 16 in Appendix A) with techniques from the fine
structure theory associated to AD+ . Woodin used his models and Theorem 7E.1 to
obtain the following determinacy:

Theorem 7F.12 (Woodin). Suppose that there is a Woodin limit of Woodin cardinals,
and a proper class of inaccessible limits of Woodin cardinals above it. Let A be -
universally Baire and let  be the pointclass of recursive preimages of A. Let C be 
in the codes. Then Glocal (, C) is determined. #

Woodin obtained this theorem as part of his work on !-logic. He used it to charac-
terize the !-logic completeness of the 12 -in- consequences of the CH, where 
is the pointclass of -universally Baire sets, in terms of the determinacy of open length
1 games with  payoff.
Let us end with an application of Theorem 7F.4, also due to Woodin. We present
this application through exercises with extensive hints. Ultimately they lead to a proof
of the consistency of determinacy for all ordinal definable games of length 1 on natural
numbers.

Exercise 7F.13. Suppose that every real has a sharp. Let N be a countable model of
ZFC and let  be an iteration strategy for N . Prove that there exists a sequence of
reals $a | < % so that:
298 7 Games which reach local cardinals

(1) N belongs to the model P = L[a | < ], and N is countable in P ;

(2) 1P is precisely equal to ;


L[a | <]
(3) is minimal, in the sense that 1  = for each < ; and

(4)  P belongs to P .

 P in condition (4) is the restriction of  to iteration trees which belong to P and are
countable in P .

Hint. Construct $a | < % so that a0 is a real coding N, and for each > 0, a is a
real coding the restriction of  to countable iteration trees in L[a | < ]. Continue
L[a | <]
until reaching the first so that 1 = . #

Exercise 7F.14 (Woodin). Suppose that there exists an iterable class model M with
a cardinal so that: (a) is a Woodin limit of Woodin cardinals in M; and (b) is
countable in V. Prove that there is a class model P so that, in P , every definable (over
P , with no parameters) length 1P game on natural numbers is determined.

Hint. Let  be an iteration strategy for M. Let P = L[a | < ] be obtained by


Exercise 7F.13, applied to N = M& and the restriction of  to trees on N. Let (v) be
a formula in the language of set theory, with one free variable. Let G be the following
game: Players I and II alternate playing natural numbers in the usual fashion, to produce
a sequence of reals x = $x | < % (of length = 1P ). If P |= [ x ] then player I
wins; and otherwise player II wins. You will show that G is determined in P .
Let G be the following game: Players I and II alternate playing natural numbers
in the usual fashion, to produce reals z . They continue until reaching the first
which is uncountable in L[z | < ]. At that point the game ends. Player I wins if
L[z | < ] |= [zmain ], where zmain is the sequence $z | < and is even%; and
otherwise player II wins.
Runs of G thus give rise to a main part zmain , and an auxiliary part, consisting
of the reals z for odd . The auxiliary part allows each of the two players to put as
much information as she wants into the model L[z] consulted in the payoff condition.
By Theorem 7F.4, G is determined. Using Remark 7E.2 (and the comment which
follows it) get a winning strategy in G so that P belongs to P . Use this to produce
a strategy in G, which belongs to P and wins against all plays in P . #

Exercise 7F.15 (Woodin). Work under the assumptions of the previous exercise. Prove
that ZFC+all ordinal definable games of length 1 on natural numbers are determined
is consistent. In fact prove the consistency of ZFC+all length 1 games on natural
numbers, definable from real and ordinal parameters, are determined.

Hint. There are standard ways to derive determinacy for games with parameters from
determinacy for games without parameters. Ordinal parameters can be added by looking
at the least non-determined game, where least involves a definable wellordering
7F Determinacy 299

of all the ordinal definable games. Real parameters can be added by asking player I
to play, in her first moves, a real which defines a non-determined game, and then
continue to play in the game defined by that real. #
Appendix A
Extenders, generic extensions, and iterability

Let L be the language of set theory with an additional unary relation symbol K. Let
ZFC consist of the axioms of ZFC with replacement and comprehension extended to all
formulae in L . We work throughout the book with the language L and with models
of ZFC . The addition of the symbol K has to do with our treatment of extenders and
iteration trees in connection with generic extensions, to be explained below.

Extenders
Definition 1. By a (, ) extender we mean an object E = $Ea | a []< % so that:

(1) each Ea is a -complete measure on []|a| ;

(2) the measures $Ea | a []< % are compatible;

(3) E is normal;

(4) E is -complete; and

(5) K(E).

Conditions (1)(4) in Definition 1 follow MartinSteel [19, 1A] and we refer the
reader to that paper for more precise details. Condition (5) is phrased in the language L .
It places the demand that E belongs to the class {x | K(x)}. This class may encompass
the entire universe, or all objects in the universe which satisfy conditions (1)(4), and
in such cases condition (5) is vacuous. But there are cases where we have to make the
class more restrictive.
If E is an extender in M then we use Ult(M, E) to denote the ultrapower of M
by E and use iEM : M Ult(M, E) to denote the ultrapower embedding. We shall not
go into the details of the ultrapower construction here. The reader may find these in
MartinSteel [18]. Note that the construction must be adapted to L , but the adaptation
is trivial: for x = [f, a] Ult(M, E) set K Ult(M,E) (x) iff K M (f (u)) for Ea a.e. u.
Given a ZFC model M and an ordinal we use M& to denote VM . We say
that two models M and N agree to just in case that M& = N&. If E is a (, )
extender in M, and N agrees with M to + 1, then Ult(N, E) makes sense and we use
iEN : N Ult(N, E) to denote the corresponding embedding.
We use StrengthM (E), the strength of E in M, to denote the largest so that M
and Ult(M, E) agree to . E is -strong if its strength is at least .
302 A Extenders, generic extensions, and iterability

We use crit(E), the critical point of E, to denote the critical point of iEM , namely
the least ordinal so that iEM ( )  = . Following MartinSteel [19], Definition 1 has
a built-in restriction to short extendersthat is extenders measuring subsets of their
critical point and no more. The restriction comes in through condition (1) which forces
the critical point of a (, ) extender to be precisely , never less. A more general
definition can be found in Kanamori [11, 26], but we do not use these more general
extenders at all in this book. The built-in restriction to short extenders means that the
level of agreement needed between M and N to make sense of the ultrapower of N by
an extender E M is crit(E) + 1. The restriction also forces our extenders to always
be at or below superstrong, in the sense given by:

Fact 2. Let E be an extender in M and let be the critical point of E. Then


StrengthM (E) iEM ().

Let E be a (, ) extender of a model M. We say that E is countable in V if


and (2 )M are both countable in V. Literally this is equivalent to the statement
that E consists of countably-in-V many measures, and each of these measures is itself
countable in V.
An embedding h : N Q is said to preserve countability just in case that: (1) for
every which is inaccessible in N , if is countable in V then so is h( ); and (2) for
every N , if (2 )N is countable in V then so is (2h() )Q .
Clause (1) implies that if is Woodin in N and countable in V then h( ) is countable
in V. Clause (2) implies that if is Woodin in N and N& + 1 is countable in V then
Q&h() + 1 is countable in V. All the uses of preservation of countability in the book
are made through these two consequences.

Fact 3. Let M and N be models of ZFC . Let E be an extender of M. Suppose that M


and N agree past crit(E). Suppose that E is countable in V. Then iEN : N Ult(N, E)
preserves countability.

Claim 4. The family of embeddings which preserve countability is closed under com-
positions and countable direct limits. #

All the ultrapower embeddings we work with in this book match the settings of
Fact 3. It follows that these embeddings, and their compositions and countable direct
limitsfor example the ones obtained through iteration trees, see belowpreserve
countability.

Interpretation of K in generic extensions


Let M M be models of (some sufficiently large part of) ZFC . Let E = $Ea | a
[]< % be an extender in M. The fattening of E to M is the object E = $Ea | a
A Extenders, generic extensions, and iterability 303

[]< % defined by setting, for each a []< and each A M ,



1 if (A M)(A A Ea (A) = 1), and
Ea (A ) =
0 otherwise.

Let P be a forcing notion in M. Let G be P-generic/M. We adapt the definition


of a forcing extension to the language L by setting K M[G] (x) iff x is the fattening to
M[G] of an extender of M.

Remark 5. We say that K has a trivial interpretation in M just in case that {x M |


K M (x)} contains all x which satisfy conditions (1)(4) of Definition 1 in M. Note that
even if K has a trivial interpretation in M, it need not have a trivial interpretation in
M[G]. There may very well be objects E M[G] which satisfy conditions (1)(4)
in M[G] yet fail to be fattenings of extenders of M.

Claim 6. Let E be an extender in M with crit(E) > cardM (P). Let E be the fat-
tening of E to M[G]. Then E is an extender in M[G]. Moreover, Ult(M[G], E ) =
Ult(M, E)[G], and iEM[G]
extends iEM .

Proof. Condition (5) of Definition 1 holds for E trivially through our definition of
K M[G] . The argument showing that E satisfies conditions (1)(4) in M[G] is standard
using the fact that P is a small forcing relative to crit(E). The same standard argument
shows that Ult(M[G], E ) and Ult(M, E)[G] have the same universe, and that iEM[G]
)
extends iEM . Using this and os Theorem its easy to check further that K Ult(M[G],E =
K Ult(M,E)[G] . #

Claim 7. Let E be an extender in M with crit(E) > cardM (P). Let N be a model
of ZFC and suppose that N agrees with M past crit(E). (In particular P N, G is
P-generic/N , and N[G] agrees with M[G] past crit(E).) Let E be the fattening of E
to M[G]. Then Ult(N [G], E ) = Ult(N, E)[G] and iEN[G] extends iEN .

Proof. Similar to the proof of Claim 6. #

Claim 8. Every extender in M[G] is a fattening of an extender in M.

Proof. Immediate using the definition of K M[G] , and using condition (5) in Definition 1.
#

Claim 8 would fail if it werent for the addition of condition (5) to Definition 1 (and
implicitly the addition of K to our language).
304 A Extenders, generic extensions, and iterability

Iteration trees
An iteration tree T of length on M consists of a tree order T on ,1 and objects
$E | + 1 < % and $M , j, | T < % satisfying the following conditions:

(1) M0 = M.

(2) For each so that + 1 < , E is either equal to pad or else it is an extender
of M .

(3) If E is an extender of M then M +1 = Ult(M , E ) and j, +1 : M M +1


is the ultrapower embedding, where is the T -predecessor of +1. (Implicit
in this definition of M +1 and j, +1 is the demand that M and M agree past
crit(E ).) If E =pad then the T -predecessor of + 1 is , M +1 = M , and
j, +1 is the identity.

(4) For limit < , M is the direct limit of the system $M , j, | T < %, and
j, : M M for T are the direct limit embeddings.

(5) The remaining embeddings j, for T < are obtained through composition.

Notice that we allow padding in iteration trees. This provides some extra flexibility that
is often useful for indexing in our constructions. Of course the padding can always be
removed, possibly making the iteration tree shorter.
A cofinal branch through T is a set b , linearly ordered by T , downward closed
under T , and cofinal in . Given a cofinal branch b we use Mb to denote the direct limit
of the system $M , j, | T b% and use j,b : M Mb to denote the direct limit
embeddings. jb stands for j0,b , embedding M = M0 into Mb . We refer to Mb as the
direct limit along b, and to jb as the direct limit embedding along b. b is a wellfounded
cofinal branch just in case that Mb is wellfounded. These definitions are all taken from
MartinSteel [19], and we refer the reader to that paper for more details.

Remark 9. For the most part in this book we construct iteration trees of length . We
describe these trees informally, by specifying En and the T -predecessor of n + 1 for
each n < . These objects, and the starting model M, determine the tree completely.

Let P be a forcing notion in M and let G be P-generic/M. Let T be an iteration


tree on M[G] with tree order T , extenders E , models M , and embeddings j,
.

T is called a fattening of T just in case that T =T , M = M [G] for each


< lh(T ), E equals the fattening of E to M [G] for each so that + 1 < lh(T ),
j
and j, , for all T . Using Claims 7 and 8 we obtain the following:

Claim 10. Suppose that all the critical points used in T are above cardM (P). Then T
can be fattened to an iteration tree on M[G]. #
1 See Section 7B for the definition of a tree order.
A Extenders, generic extensions, and iterability 305

Claim 11. Every iteration tree on M[G] with critical points above cardM (P) is a
fattening of an iteration tree on M. In particular, if G collapses cardM (P) to then all
iteration trees on M[G] are fattenings of iteration trees on M. #

These two claims allow us to regard iteration trees on generic extensions of M as


iteration trees on M and vice versa. We switch between the two viewpoints as needed
throughout the book, and often use the same letter T to stand for both an iteration tree
on M and the fattened tree on M[G].

Iterability
The iteration trees constructed in this book are all normal and plus 2 in the sense of
MartinSteel [19]. These technical restrictions on the trees are needed for the following
theorem:

Theorem 12 (MartinSteel [19]). Let : M V be elementary where M is countable


and V is some rank initial segment of V.2 Let T be a normal plus 2 iteration tree
of length on M. Then there is a cofinal branch b through T , and an embedding
: Mb V , so that jb = . (Note that b is then a wellfounded branch, since
Mb embeds into V .) #

A weak iteration of M of length consists of objects M , T , b for < and


embeddings j, : M M for < < , so that:

(1) M0 = M.

(2) For each < , T is a normal, plus 2 iteration tree of length on M ; b is a


cofinal branch through T ; M +1 is the direct limit along b ; and j, +1 : M
M +1 is the direct limit embedding along b .

(3) For limit < , M is the direct limit of the system $M , j, | < < % and
j, : M M are the direct limit embeddings.

(4) The remaining embeddings j, are obtained by composition.

A weak iteration is thus a linear composition of blocks. Each block is generated by a


length iteration tree and a cofinal branch through it.
In the weak iteration game on M players good and bad collaborate to produce a
weak iteration of M, of length 1V . Bad plays the iteration trees T and good plays
the branches b . (These moves determine the iteration completely.) If ever a model
M , < 1 , is reached which is illfounded, then bad wins. Otherwise good wins.
M is weakly iterable if good has a winning strategy in the weak iteration game.
2 The actual universe V is expanded to the language L through the trivial assignment K V (x) for all x.
306 A Extenders, generic extensions, and iterability

Claim 13 (MartinSteel [19]). Let : M V be elementary where M is countable


and V is some rank initial segment of V. Then good has a winning strategy in the
weak iteration game on M.
Proof. Immediate through iterated applications of Theorem 12. The good player should
simply keep choosing branches given by Theorem 12, successively embedding each
M +1 into V , and preserving commutativity which is needed for the limits. #
A mild iteration of M of length consists of objects M , T , b , Q , k , F for
< and embeddings j, : M M for < < , so that:
(1) M0 = M.
(2) For each < , T is a normal, plus 2 iteration tree of length on M ; b is a
cofinal branch through T ; Q is the direct limit along b ; and k : M Q is
the direct limit embedding along b .
(3) F is either an extender of Q with crit(F ) < crit(k ), or equal to undefined.
If F =undefined then M +1 is equal to Q and j, +1 is equal to k . If F is an
extender then M +1 = Ult(M , F ) and j, +1 : M M +1 is the ultrapower
embedding.
(4) For limit < , M is the direct limit of the system $M , j, | < < % and
j, : M M are the direct limit embeddings.
(5) The remaining embeddings j, are obtained by composition.
A mild iteration is thus a linear composition of blocks. Each block is either generated
simply by a length iteration tree and a branch through it, or generated by a length
iteration tree, a branch through it, and an additional ultrapower. The two possibilities
are illustrated in Diagram 1.

j, +1

kk # k
Sk
Sk
/ M k SSSk / Q
F
M +1 / M kSkSkSkS / Q = M +1 /
M0
S k
T T

Diagram 1. A mild iteration.

In the mild iteration game on M players good and bad collaborate to produce a
mild iteration of M, of length 1V . Bad plays T and F for each , and good plays
the branches b . (These moves determine the iteration completely.) If ever a model
M , < 1 , is reached which is illfounded, then bad wins. Otherwise good wins.
M is mildly iterable if good has a winning strategy in the mild iteration game.
Claim 14. Let M, , T , and b be as in Theorem 12. Let Q be the direct limit along b
and let k : M Q be the direct limit embedding. Let F be an extender of Q with
crit(F ) < crit(k). Let M = Ult(M, F ) and let j : M M be the ultrapower
embedding. Then there is an embedding : M V so that j = .
A Extenders, generic extensions, and iterability 307

Proof. Let : Q V be given by Theorem 12. We have k = . It follows


that and agree through crit(k), and therefore past crit(F ). Using this agreement, a
standard argument allows copying the ultrapower of M by F to an ultrapower of V by
(F ). The copying argument produces the embedding illustrated in the left part of
Diagram 2 with the commutativity j = h where h is the ultrapower embedding
of V by (F ). Using and the -completeness of (F ) in V a standard argument
now produces .

/ Ult(V , (F ))
V
O cG cGG V
O hQQQQ
G GG QQQ

G GG QQQ
GG
QQQ
T kk G G T kk
k Q
Sk
k k
SSSS / Q S
kk
S /
M SSS k Q M
F < F <
M M
k

j j

Diagram 2. The proof of Claim 14 (left) and the end result (right).
#

Claim 15. Let : M V be elementary where M is countable and V is some rank


initial segment of V. Then good has a winning strategy in the mild iteration game
on M.
Proof. Immediate through iterated applications of Theorem 12, in cases where
F =undefined, and Claim 14 in cases where F is defined. #
In the ( full ) iteration game on M players good and bad collaborate to construct
an iteration tree T of length 1V +1 on M. bad plays all the extenders, and determines
the T -predecessor of +1 for each . good plays the branches { | T } for limit ,
thereby determining the limit models M . (Note that good is also responsible for the
final move, which determines MV .) If ever a model along the tree is reached which is
1
illfounded then bad wins. Otherwise good wins. M is ( fully) iterable if good
has a winning strategy in this game. An iteration strategy for M is a strategy for the
good player in the iteration game on M. Weak and mild iteration strategies are defined
similarly.
Notions of iterability including weak and full were introduced by MartinSteel [19].
Their results summarized above show how to obtain weakly and mildly iterable models.
Proofs of full iterability are substantially more complicated. They seem to require fine
structure, and to date they require also some limitation on the large cardinals involved.
There have been several results progressively pushing this limitation further up, starting
with the work of MartinSteel [19] which obtains iterability for models with a Woodin
cardinal. For our purposes in this book the following theorem of Neeman [31] reaches
a sufficiently large cardinal.
Theorem 16 (Neeman [31]). Suppose that there is a Woodin limit of Woodin cardinals
in V, say , and suppose that (V ) exists. Then there is a class model M of ZFC and
308 A Extenders, generic extensions, and iterability

some M so that: is a Woodin limit of Woodin cardinals in M; is countable in V;


and M is fully iterable. #

We conclude the appendix with a brief index to the uses of iterability in this book.
Let X be a large cardinal axiom. Suppose that X holds in some rank initial segment V
of V. Using Claim 13 it follows that there is a countable model M so that X holds in
M and M is weakly iterable. If in addition (V ) exists in V then one can expand the
argument and produce a class model M of ZFC so that:
M = L(VM ) for some M;

X holds in M (in fact in VM );

VM is countable in V; and
M is weakly iterable.

Models of this kind, with X standing for the existence of a certain number of Woodin
cardinals, suffice for the determinacy proofs in Chapter 2.
The constructions in Chapter 3 require mild iterability. Models satisfying the as-
sumptions in Chapter 3 can be obtained from sharps for appropriate large cardinals in V,
using Claim 15.
For the determinacy in Chapter 7 even mild iterability is not enough. The iteration
trees there are non-linear compositions of blocks, similar to the ones in mild iterations
except that F may be applied to M for , not only to M . Such trees are as
complicated as the ones in the full iteration game. The construction therefore demands
a model which is fully iterable. For the large cardinal assumption in Theorem 7F.4
the necessary model can be obtained assuming a sharp for a Woodin limit of Woodin
cardinals in V, using Theorem 16.
Bibliography

[1] John W. Addison and Yiannis N. Moschovakis. Some consequences of the axiom
of definable determinateness. Proc. Nat. Acad. Sci. U.S.A., 59:708712, 1968.

[2] David Blackwell. Infinite games and analytic sets. Proc. Nat. Acad. Sci. U.S.A.,
58:18361837, 1967.

[3] Douglas Burke and Ernest Schimmerling. Handwritten notes from a course taught
by W. Hugh Woodin at UC Berkeley in the Spring of 1990.

[4] Morton Davis. Infinite games of perfect information. In Advances in game theory,
pages 85101. Princeton University Press, Princeton, N.J., 1964.

[5] Qi Feng, Menachem Magidor, and Hugh Woodin. Universally Baire sets of reals.
In Set theory of the continuum (Berkeley, CA, 1989), Math. Sci. Res. Inst. Publ. 26,
pages 203242. Springer-Verlag, New York, 1992.

[6] Matthew Foreman, Menachem Magidor, and Saharon Shelah. Martins maximum,
saturated ideals, and nonregular ultrafilters. I. Ann. of Math. (2), 127(1):147,
1988.

[7] Matthew Foreman. Potent axioms. Trans. Amer. Math. Soc., 294(1):128, 1986.

[8] David Gale and Frank M. Stewart. Infinite games with perfect information.
In Contributions to the theory of games, vol. 2, Annals of Mathematics Studies 28,
pages 245266. Princeton University Press, Princeton, N. J., 1953.

[9] Leo Harrington. Analytic determinacy and 0 . J. Symbolic Logic, 43(4):685693,


1978.

[10] Thomas Jech. Set theory. The third millennium edition, revised and expanded.
Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2003.

[11] Akihiro Kanamori. The higher infinite. Large cardinals in set theory from their
beginnings. Perspectives in Mathematical Logic. Springer-Verlag, Berlin, 1994.

[12] Alexander S. Kechris and W. Hugh Woodin. Equivalence of partition properties


and determinacy. Proc. Nat. Acad. Sci. U.S.A., 80(6 i.):17831786, 1983.

[13] Alexander S. Kechris. Homogeneous trees and projective scales. In Cabal Seminar
7779 (Proc. Caltech-UCLA Logic Sem., 197779), Lecture Notes in Math. 839,
pages 3373. Springer-Verlag, Berlin, 1981.

[14] Kenneth Kunen. Saturated ideals. J. Symbolic Logic, 43(1):6576, 1978.


310 Bibliography

[15] Kenneth Kunen. Set theory. An introduction to independence proofs. Studies in


Logic and the Foundations of Mathematics 102. North-Holland Publishing Co.,
Amsterdam, 1980.

[16] Casimir Kuratowski. Sur les thormes de sparation dans la Thorie des ensem-
bles. Fund. Math., 26:183191, 1936.

[17] Menachem Magidor. Precipitous ideals and  14 sets. Israel J. Math., 35(12):
109134, 1980.

[18] Donald A. Martin and John R. Steel. A proof of projective determinacy. J. Amer.
Math. Soc., 2(1):71125, 1989.

[19] Donald A. Martin and John R. Steel. Iteration trees. J. Amer. Math. Soc., 7(1):173,
1994.

[20] Donald A. Martin. The axiom of determinateness and reduction principles in the
analytical hierarchy. Bull. Amer. Math. Soc., 74:687689, 1968.

[21] Donald A. Martin. Measurable cardinals and analytic games. Fund. Math.,
66:287291, 1969/1970.

[22] Donald A. Martin. Borel determinacy. Ann. of Math. (2), 102(2):363371, 1975.

[23] Donald A. Martin. Infinite games. In Proceedings of the International Congress


of Mathematicians (Helsinki, 1978), pages 269273, Helsinki, 1980. Acad. Sci.
Fennica.

[24] Donald A. Martin. The largest countable this, that, and the other. In Cabal seminar
7981, Lecture Notes in Math. 1019, pages 97106. Springer-Verlag, Berlin, 1983.

[25] Donald A. Martin. A purely inductive proof of Borel determinacy. In Recursion


theory (Ithaca, N.Y., 1982), Proc. Sympos. Pure Math. 42, pages 303308. Amer.
Math. Soc., Providence, RI, 1985.

[26] Yiannis N. Moschovakis. Descriptive set theory. Studies in Logic and the Foun-
dations of Mathematics 100. North-Holland Publishing Co., Amsterdam, 1980.

[27] Jan Mycielski and Hugo Steinhaus. A mathematical axiom contradicting the axiom
of choice. Bull. Acad. Polon. Sci. Sr. Sci. Math. Astronom. Phys., 10:13, 1962.

[28] Jan Mycielski and Stanisaw Swierczkowski. On the Lebesgue measurability and
the axiom of determinateness. Fund. Math., 54:6771, 1964.

[29] Itay Neeman. Optimal proofs of determinacy. Bull. Symbolic Logic, 1(3):327339,
1995.

[30] Itay Neeman. Large cardinals and the determinacy of long games. PhD thesis,
UCLA, 1996.
Bibliography 311

[31] Itay Neeman. Inner models in the region of a Woodin limit of Woodin cardinals.
Ann. Pure Appl. Logic, 116(13):67155, 2002.

[32] Itay Neeman. Optimal proofs of determinacy. II. J. Math. Log., 2(2):227258,
2002.

[33] Itay Neeman. Games of length 1 . To appear.


[34] Itay Neeman. An introduction to proofs of determinacy of long games. Proceed-
ings of Logic Colloquium 2001. To appear.

[35] Itay Neeman. The strength of continuously coded determinacy. In preparation.

[36] John C. Oxtoby. The Banach-Mazur game and Banach category theorem. In
Contributions to the theory of games, vol. 3, Annals of Mathematics Studies 39,
pages 159163. Princeton University Press, Princeton, N. J., 1957.

[37] Saharon Shelah and W. Hugh Woodin. Large cardinals imply that every reasonably
definable set of reals is Lebesgue measurable. Israel J. Math., 70(3):381394,
1990.

[38] Saharon Shelah. Proper forcing. Lecture Notes in Math. 940. Springer-Verlag,
Berlin, 1982.

[39] Jack H. Silver. Some applications of model theory in set theory. PhD thesis,
Berkeley, 1966.

[40] Jack H. Silver. Some applications of model theory in set theory. Ann. Math. Logic,
3(1):45110, 1971.

[41] John R. Steel. Long games. In Cabal Seminar 8185, Lecture Notes in Math. 1333,
pages 5697. Springer-Verlag, Berlin, 1988.
[42] John R. Steel. The well-foundedness of the Mitchell order. J. Symbolic Logic,
58(3):931940, 1993.
[43] Philip Wolfe. The strict determinateness of certain infinite games. Pacific J. Math.,
5:841847, 1955.

[44] W. Hugh Woodin. Aspects of determinacy. In Logic, methodology and philos-


ophy of science, VII (Salzburg, 1983), Studies in Logic and the Foundations of
Mathematics 114, pages 171181. North-Holland, Amsterdam, 1986.

[45] W. Hugh Woodin. The extender algebra. Presented during lectures on 12 deter-
minacy at the Cabal seminar, CaltechUCLA, April 1988.

[46] W. Hugh Woodin. Supercompact cardinals, sets of reals, and weakly homogeneous
trees. Proc. Nat. Acad. Sci. U.S.A., 85(18):65876591, 1988.
Index

A (auxiliary games map), 15, 23 cofinal branch through an iteration tree,


definition, 2123 304
Amix (mixed pivot games map), 43, 50 continuously coded length, 87
definition, 4344 countability preserving embeddings,
exact rules, 4446 302
first fifth of a mega-round, 222 countable extender, 302
useful positions, 48 critical point (of an extender), 302
Apiv (pivot games map), 28, 30
definition, 2830 (, )-pullback, see pullback
acceptable obstruction, 156, 157 -name
I-acceptable, 168 for a Woodin limit of Woodin
II-acceptable, 167, 168 cardinals, 162
agreement between models, 301 for W , 155
amenable witness to t -sequence
when = rdm(t) is a limit, 140 for a Woodin limit of Woodin
when = rdm(t) is a successor, cardinals, 162
139 for W , 155
annotated position, 137 determinacy
auxiliary games map, 23 games ending at 1 in L of the
mirrored, 38 play, 293
avoids C and D (-sequence which) games ending at 1 in  of the
play, 296, 297
for a Woodin limit of Woodin
games of continuously coded
cardinals, 194
length, 119, 120, 127, 128
for W , 176
games of fixed countable length,
53, 73, 85, 86
B (mirrored auxiliary games map), 37
games with homogeneously Suslin
Bpiv (mirrored pivot games map), 38 payoff, 86, 128
Back(, )(C ), see pullback games with universally Baire
Back I and Back II , 167 payoff, 83, 85, 127, 297
Baire, see universally Baire set ordinal definable games of length
basic axiom, 146 1 (consistency result), 298
basic identity, 141142 12 sets, 39
branch through a tree order, 271
e(), 135
captured pointclass, 295 early end
clear annotated position, 151 branch ,
in a limit mega-round of G
code (and -code), 138139 213
code (in Chapter 3) for a -position, 88 in a successor mega-round of
cofinal branch through a tree order, 271 skip , 218
G
314 Index

elastic type, 18 cont , 124126


G
even branch, 15, 27
Gfixed , 52
in runs of Amix , 47 local , 268269
G
exceeding type, 18 skip
G
expansion of a reduced position, 249 definition, 218219
extender, 301 early end in a successor
countable, 302 mega-round, 218
critical point, 302 expansion of a reduced position,
fattening, 302 249
short, 302 position, 249
strength, 301 reduced position, 249
with respect to H , 19 skip, 219
extender algebra, 148 G(t)
chain condition, 148 definition, 150
definition, 143148 genericity, 151
generics, 151 games ending at 1 in L of the play
history, 8, 154 definition, 292
extender axiom, 147 determinacy, 293
extension (and a-extension) for steps, games ending at 1 in  of the play
177 definition, 295
extension (and n-extension) for witnesses, determinacy, 296, 297
137, 140 games of continuously coded length
extension of a tot to length + 0.2, 276 definition, 87
determinacy, 119, 120, 127, 128
fattening games of fixed countable length
of an extender, 302 definition, 51
of an iteration tree, 304 determinacy, 53, 73, 85, 86
first fifth of a mega-round of Rmix [x], generic run, 15, 23
222 in mirrored auxiliary games, 38
full iterability, 307308 generic strategies map, 27
functor, 156 mirrored, 38
genericity iteration, 43, 154
 (game quantifier), 83, 85, 127
branch
G cont , 126
H
definition, 209214 fixed , 53
H
phantom limit case, 212
Hlocal , 269
standard limit case, 212214 height of an identity, 142
successor case, 210211 branch , 215
history of a position in G
early end in a limit mega-round, hopeful -sequence wrt Y , 162
213
history of a position, 215 identity, 142
leap, 213 indiscernibles, see local indiscernibles
outcome of a position, 215 inference (on identities), 143144
position, 215 interpretation of K
Index 315

in generic extensions, 302 minimal obstruction, 152


in ultrapowers, 301 mirrored
iseq (identities sequence), 151 auxiliary games map, pivot games
realization, 151 map, and strategies map,
iterability, 305308 3738
full, 307 pullback, 167174
mild, 306 mixed pivot, 43, 47
uses of, 308 residing above , 50
weak, 305 useful, 48
iteration game mixed pivot games map, 50
full, 307 mixed pivot strategies map, 50
mild, 306
weak, 305 -name, see -name
iteration strategy, 307 nice annotated position, 176
iteration tree, 304 nice -sequence, 183
fattening, 304 nice iteration tree, 27
nice, 27 normal iteration tree, 29
normal, 29 -position, 87
padding, 304 -run, 88
using only extenders from below
, 29 obstruction, 151
acceptable, 156, 157
jumpM (), 20 minimal, 152
obstruction free annotated position, 151
K, 301 odd branch, 15, 28
interpretation, see interpretation in runs of Amix , 47
of K One Step Lemma, 20
K U , 273 , see u v
(, n)-type, see type otrs(), 291
outcome of a position in G branch , 215
L, 136
L(t ()), see L(w) padding (in iteration trees), 304
L(w), 136 phantom limit
leap, 213 in a regular tot, 273
length of a skip frame, 250 in G branch , 212
<2 11 sets, 51 ini (, C), 175, 208
linear strand, see strand 12 determinacy, 39
local indiscernibles, 20 pivot, 15, 28
locally definable strategy, 291 in mirrored auxiliary games, 38
mixed, see mixed pivot
Martins auxiliary game, 7475 residing above , 36
related facts, 75, 80 pivot games map, 30
merge of a tot, 272 mirrored, 38
mild iterability, 306307 pivot strategies map, 35
316 Index

mirrored, 38 piv (pivot strategies map), 28, 35


position in G branch , 215 construction, 3035

position in Gskip , 249 properties, 3536
preservation of countability (for embed- skip, 219
dings), 302 skip frame, 250
projm (u), 17 skipping game, 218
ini (, D), 175, 208 standard limit
pullback, 155 in a regular tot, 273
compositions and summary, in G branch , 212
166167 , see x  y
mirrored, 174 step, 177
from a relative limit, 158162 strand, 276
mirrored, 169171 strength (of an extender), 301
from a relative successor, 165 with respect to H , 19
mirrored, 173174 StretchE (u), 18
from a Woodin limit of Woodin subtype, 17
cardinals, 162165 suitable annotated position, 156
mirrored, 172173 I-suitable, 168
II-suitable, 167
rdm(t), 137
supernice -sequence, 201
real part of an annotated position, 141
realizable branch through an iteration
t[G], 154
tree, 83
t, w, see code
red(P , f ), 251
skip , 249 t, w, y, 138
reduced position in G
gen (mirrored generic strategies map),
regular tot, 272, 277
38
relative domain, 135, 137
piv (mirrored pivot strategies map), 38
relative limit, 135
tot (tree of trees), 271
relative successor, 135
consistent with , 272
residing above
extended to length + 0.2, 276
mixed pivot, 50
regular, 272
pivot, 36
tree order, 270
root, 47
trunc(P, b, n), 227
route, 251
truth value of an identity, 142143
saturated annotated position, 177178 type, 16
saturated -sequence, 183 domain, 16
seals Y (ordinal which), 154 elastic, 18
-sequence, see -sequence, see U-sequence exceed, 18
short extender, 302 of x0 , . . . , xn1 in V , 17
gen (generic strategies map), 27 projection, 17
mix (mixed pivot strategies map), 48, realizable, 17
50 stretch, 18
construction, 4849 sub-, 17
properties, 4950 u , 17
Index 317

u (in connection with types), 17 weak iterability, 305306


u v, 161 wellfounded branch through
U-sequence, 275 an iteration tree, 304
universally Baire set, 82 witness, 136137
propagation of u.B. representations amenable, limit , 140
through the projective hierarchy, amenable, successor , 139
85 n-extension, 140
under applications of cont , 127 Woodin cardinal, 19
under applications of  , 84
under applications of  , 85 x  y, 139
used in a witness, 136
useful runs and positions in Amix , 48 z(t), 141
ZFC (see also K, interpretation of K),
W , 135 301
W (see also extender algebra), 148

S-ar putea să vă placă și